1 Plenary: Akshay Venkatesh

A fairy tale: see Langlands Elephant. The point of this talk is to see the entire elephant!

In the analogy between number fields and 3-manifolds, automorphic forms are on the number field side – what is the manifold analog? Some history:

Let \(X = V(x^3+y^3+z^3) \subseteq {\mathbb{P}}^2_{/ {{\mathbb{C}}}}\); there are automorphisms

These should all act on \(H^*(X)\) for any such \(H^*\), but conjugation is quite discontinuous.

Let \(X \in {\mathsf{Mfd}}\), how does one compute \(H^*_{\mathrm{sing}}(X; {\mathbb{Z}})\)? Reduce to a linear-algebraic problem by triangulating and forming a chain complex of simplices. However, for \(X \mathrel{\vcenter{:}}=\operatorname{Spec}R, R\mathrel{\vcenter{:}}={\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{2} } \right] }\), \begin{align*} H^1_\text{ét}(X; C_2) &= R^{\times}/{ {R^{\times}}_{\scriptscriptstyle \square} }= \left\{{\pm 1,\pm 2}\right\} \cong C_2{ {}^{ \scriptscriptstyle\times^{2} } } \\ H^2_\text{ét}(X; C_2) &= {\mathsf{Quat}}{\mathsf{Alg}}_{/ {R}} = \left\{{ \operatorname{Mat}_{2\times 2}(R), R[i,j,k]/\left\langle{i^2=j^2=k^2=-1}\right\rangle }\right\} .\end{align*} so \(H^2\) classifies quaternion algebras over \(R\). So computing this is very different to the case of manifolds! Also note that these are not dual on-the-nose, since \(H^1,H^2\) have different orders.

Comparing duality for number rings vs 3-manifolds. For us, a number ring will be

For simplicity, we’ll take \(R = {\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }, X = \operatorname{Spec}R\). For \(M\in{\mathsf{Ab}}{\mathsf{Grp}}\) a \(p{\hbox{-}}\)torsion group (where we need the order to be invertible in \(R\)), we’ll consider \begin{align*} H^i(X; M) \mathrel{\vcenter{:}}= H^i_\text{ét}(\operatorname{Spec}{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }; M) .\end{align*} There is a LES involving \(M {}^{ \vee }\mathrel{\vcenter{:}}=\mathop{\mathrm{Hom}}_{\mathsf{Grp}}(M, S^1) = \mathop{\mathrm{Hom}}_{\mathsf{Grp}}(M, \mu_{p^\infty})\) where \(\mu_{p^\infty}\) are the \(p{\hbox{-}}\)power roots of 1:

Link to Diagram

On the other hand, let \(X\in {\mathsf{Mfd}}^3\) be a manifold with boundary, then there is a LES

Link to Diagram

So \(X \approx {\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }\) is a (nonorientable!) 3-manifold in this analogy, with \({{\partial}}X \approx { {\mathbb{Q}}_p }\), which is now like a 2-manifold. The nonorientable assumption here is related to the need to twist the Galois actions on the scheme side. How to realize this: delete a tubular neighborhood of the knot, then

Relating to automorphic forms: for \(G={\operatorname{SL}}_2\),

We would like some association that works similarly: \begin{align*} {\mathsf{Mfd}}^3 &\to { \mathsf{Vect} }_{/ {k}} \\ M &\mapsto {\mathcal{A}}_M .\end{align*}

A non-example is \(M\mapsto H_{\mathrm{sing}}^*(M; {\mathbb{C}})\), which behaves nothing like \({\mathbb{Z}}\mapsto {\mathcal{A}}_{\mathbb{Z}}\). E.g. it has the wrong type of functoriality: the map \({\mathbb{Z}}\to {\mathbb{Z}} { \left[ \scriptstyle {\sqrt 2} \right] }\) is like a branched double cover, but for manifolds there are maps both ways and here it is difficult to find wrong-way maps. Moreover the corresponding spaces of automorphic forms \({\mathcal{A}}_{{\mathbb{Z}}}, {\mathcal{A}}_{{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }}\) differ by far more than just a part coming from \({ {\mathbb{Q}}_p }\).

Note also that \(H^*(M \cup N) \cong H^*(M) \oplus H^*(N)\), but for automorphic forms we have \({\mathcal{A}}_{{\mathbb{Z}}\oplus {\mathbb{Z}}} \cong {\mathcal{A}}_{\mathbb{Z}}\otimes{\mathcal{A}}_{\mathbb{Z}}\), where the source is like a degenerate quadratic extension.

A heavily studied piece of the analogy: \begin{align*} {\operatorname{ {\text{Automorphic forms}} }}\rightleftharpoons{\mathsf{TQFT}} ,\end{align*} coming from the Kapustin-Witten 2006, where a (4-dimensional) TQFT is a essentially a monoidal functor: \begin{align*} {\mathsf{TQFT}}^3 = [({\mathsf{Bord}}^3, {\textstyle\coprod}) \to ({ \mathsf{Vect} }_{/ {k}} , \otimes_k)] .\end{align*} .

See Atiyah TQFT, section 2.

2 Closing: Akshay Venkatesh

Automorphic forms using TFQTs as a metaphor. We’ll consider \({\mathsf{TQFT}}_4 = [({\mathsf{Bord}}^3, {\textstyle\coprod}), ({ \mathsf{Vect} }_{/ {{\mathbb{C}}}} , \otimes_{\mathbb{C}})]\) where \({\mathsf{Bord}}^3\) is the category whose objects are 3-manifolds and morphisms \(M\to N\) are 4-manifolds \(W\) with \({{\partial}}W = M {\textstyle\coprod}N\). These are meant to extract invariants of 4-manifolds that are amenable to cut-and-paste arguments. The correspondences:

Fix \(G\in{\mathsf{Fin}}{\mathsf{Grp}}\), and a correspondence

An informal definition of extended TQFTs, in particular \({\mathsf{TQFT}}_4\):

This should yield

Last time we said \(\operatorname{Spec}{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }\) is like a 3-manifold with boundary \({ {\mathbb{Q}}_p }\), which is like a 2-manifold. A philosophy is to put all places on the same footing – note that we haven’t included the places at \(p\) and \(\infty\) here, so really we should have \({{\partial}}\operatorname{Spec}{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] } = \operatorname{Spec}{\mathbb{R}}, \operatorname{Spec}{ {\mathbb{Q}}_p }\), and \({{\partial}}\operatorname{Spec}{\mathbb{Z}}= \operatorname{Spec}{\mathbb{R}}\). So our new picture should be:

(todo finish)

2.1 Automorphic Forms as \({\mathsf{TQFT}}_4\)

What should the automorphic form correspondence be in this analogy? Our 3-dimensional objects:

The 2-dimensional objects:

Really for a TQFT, we should assign something to objects in the category of its boundary – here e.g. the vector space \({\mathcal{A}}_{{\mathbb{Z}}}\) is an object in \({\mathsf{Rep}}G({ {\mathbb{Q}}_p })\). Idea: functions on \({\mathsf{Bun}}_G\) are hard to deal with, e.g. Hecke operators turn into infinite sums. Make things robust to passing to algebraic closures by passing from functions to sheaves!

Last time: thinking of \(\operatorname{Spec}{\mathbb{Z}}\) as a 3-manifold, \begin{align*} \operatorname{Spec}{\mathbb{Z}}= \operatorname{Spec}{ {\mathbb{Z}}_p }{ \displaystyle\coprod_{\operatorname{Spec}{ {\mathbb{Q}}_p }} } \operatorname{Spec}{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] } \approx M^3{ \displaystyle\coprod_{Z^2} }N^3 .\end{align*} We want the following: \begin{align*} \mathop{\mathrm{Hom}}_{G({ {\mathbb{Q}}_p })}({\mathcal{A}}_{{ {\mathbb{Z}}_p }}, {\mathcal{A}}_{{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }}) =_? {\mathcal{A}}_{{\mathbb{Z}}} .\end{align*}

We regard \({\mathcal{A}}_{\mathbb{Z}}\) as elements of \({\mathcal{A}}_{{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }}\) which are also unramified at \(p\), and by Frobenius reciprocity this should yield \begin{align*} {\mathcal{A}}_{\mathbb{Z}}= \mathop{\mathrm{Hom}}_{G({ {\mathbb{Q}}_p })} \qty{\left\{{ \text{Functions on } \dcosetr{ G({ {\mathbb{Q}}_p })}{G({ {\mathbb{Z}}_p })} , {\mathcal{A}}_{{\mathbb{Z}}{ \left[ { \scriptstyle \frac{1}{p} } \right] }} }\right\}} ,\end{align*} which should encode a Hecke algebra at \(p\).

What is the Langlands correspondence in this language? What should an “arithmetic TQFT” be?

Let \(O\) be a category of arithmetic ring and \(A_O\) be a category or vector space, and call an associated \(O\to A_O\) an arithmetic field theory. Let \(X_{/ {{\mathbb{F}}_p}}\) be a smooth projective curve and \(G=\operatorname{GL}_n\). The Langlands correspondence here (due to Drinfeld and Lafforgue) yields \begin{align*} \left\{{\substack{ \text{Cuspidal functions on $n{\hbox{-}}$dimensional} \\ \text{vector bundles on } X }}\right\} &\rightleftharpoons \left\{{\substack{ \text{Functions on $n{\hbox{-}}$dimensional} \\ \text{irreducible Galois reps} }}\right\}\\ T_x\curvearrowright x &\rightleftharpoons\operatorname{Frob}_x \curvearrowright? \end{align*} We regard the LHS \(A\) as “automorphic forms.”

This suggests the following viewpoint on the Langlands correspondence: we are only seeing one level, and there is a second arithmetic field theory \(B^{(G {}^{ \vee })}\) built out of Galois representations of the Langlands dual \(G {}^{ \vee }\), so \({\mathbb{Z}}\) yields a vector space and \({ {\mathbb{Q}}_p }\) yields a category, and an equivalence of arithmetic field theories \(A^{(G)} \rightleftharpoons B^{(G {}^{ \vee })}\). Often \(B\) is a category of coherent sheaves. This should package local, global, and geometric Langlands into a single theory!

The abstract correspondence between automorphic forms and Galois reps isn’t so useful; the real utility comes from matching structures and numerical invariants on both sides, e.g. Fourier coefficients, Rankin-Selberg or doubling integrals, the \(\Theta\) correspondence, etc which all match with something on the Galois side (usually an \(L{\hbox{-}}\)function). This yields a panoply of matching invariants! E.g. for \(E/{\mathbb{Q}}\) an elliptic curve, \begin{align*} L(\operatorname{Sym}^2 E, 1) = \prod_p {p^2 \over (1-1/p) {\sharp}E({\mathbb{F}}_{p^2})} \in \pi\cdot \mathrm{Area}(E_{\mathbb{C}}) {\mathbb{Q}} ,\end{align*} where the area is of the fundamental parallelogram of \(E\), which is hard to prove without automorphic forms. How can we interpret this in terms of TQFTs?

Consider numerical invariants of automorphic forms and Galois reps landing in \({\mathbb{C}}\). Let \({\mathcal{O}}\) be a 3-dimensional ring of integers over \(X\).

The numerical invariants of Galois reps should be elements of \(B_{\mathcal{O}}^{(G {}^{ \vee })}\), and numerical invariants of automorphic forms should come from \(A_{\mathcal{O}}^{(G)}\) where given \(P\), one takes \(\phi\mapsto {\left\langle {P},~{\phi} \right\rangle}\). To find matching invariants, we want to match elements in \(A_{\mathcal{O}}\) to elements in \(B_{\mathcal{O}}\). More ambitiously, we can ask for matching boundary conditions in \(A^{(G)}\) and \(B^{(G {}^{ \vee })}\).

A boundary condition in \({\mathsf{TQFT}}_4\) is a coherent assignment:

See Kasputin’s 2010 ICM address for a nice overview.

Joint work with David Ben-Zvi, Sakellaridis, an informal summary:

For the next generation of number theorists: why are there such similarities between TQFTs and automorphic forms? This is something deep that we barely understand at all.

Extending to 1-dimensional objects: these should be 2-categories which are categorical reps of a loop group.

3 Ellen Eischen, Automorphic Forms on Unitary Groups, Talk 1

Overall plan:

Today:

If the previous talk was a “fairy tale,” this will flesh out the “based on a true story” part!

3.1 Motivation from modular forms

Consider \(\zeta(2k)\) for \(k\in {\mathbb{Z}}_{\geq 0}\); known to Euler as \begin{align*} \zeta(2k) = (-1)^k \pi^{2k} {2^{2k-1} \over (2k-1)!} \qty{-{ {B_{2k}} \over 2k} } = (-1)^k \pi^{2k} {2^{2k-1} \over (2k-1)!} \zeta(1-2k) ,\end{align*} where \(B_{2k}\) is the \(2k\)th Bernoulli number, whose exponential generating function is \begin{align*} {ze^z\over e^z-1} = \sum_{k\geq 0} B_k {z^k\over k!} \in {\mathbb{Q}}{\left[\left[ z \right]\right] } .\end{align*}

Proving rationality of \(\zeta(2k)\) (up to powers \(\pi^n\)) involves the normalized Eisenstein series \begin{align*} G_{2k}(q) = \zeta(1-2k) + 2\sum_{k\geq 1} \sigma_{2k-1}(n) q^n, \quad q\mathrel{\vcenter{:}}= e^{2\pi i z}, \quad \sigma_k(n) \mathrel{\vcenter{:}}=\sum_{d\divides n} d^k ,\end{align*} and one can use similar techniques to prove rationality for

All of these correspond to Artin \(L{\hbox{-}}\)functions \(L(s, \rho)\) for \(\rho\) a Galois representation. Dimensions \(n=1\) and (partially) \(n=2\) are handled class field theory. Can we generally show special values are algebraic? And if so, what do these values mean?

More generally, one can ask about algebraicity or rationality of special values of \(L{\hbox{-}}\)functions attached to modular forms. Our first tool for constructing such things will be Rankin-Selberg convolution. Why care about special values: Kummer used congruences for \(\zeta\), checking if \(p\divides { \operatorname{cl}} (K)\) is equivalent to checking if \(p\) divides numerators of Bernoulli numbers, which can be used to prove special cases of Fermat. Picked up later for Iwasawa theory, controls behavior of towers of towers of cyclotomic extensions in \({\mathsf{G_K}{\hbox{-}}\mathsf{Mod}}\).

Conjectures

It might seem like \(\operatorname{GL}_n\) for \(n\geq 3\) is the next step, but this turns out to be too general! Even \({\operatorname{SL}}_n\) in these ranges is difficult. Instead we’ll move to unitary groups, where we’ll have Shimura varieties to work with.

3.2 Unitary groups

Fix \(K\in {\mathrm{CM}}\mathsf{Field}\), so \(K/K^+/{\mathbb{Q}}\) with \(K^+/{\mathbb{Q}}\) totally real and \(K/K^+\) quadratic imaginary, and \(V\in { \mathsf{Vect} }_{/ {K}}\) with a nondegenerate Hermitian pairing \({\left\langle {{-}},~{{-}} \right\rangle}\), which can be extended linearly to \(V_R \mathrel{\vcenter{:}}= V\otimes_{K^+} R\) for any \(R\in\mathsf{Alg}_{/ {K^+}}\).

The general unitary group is the algebraic group \(G \mathrel{\vcenter{:}}=\operatorname{GU}(V, {\left\langle {{-}},~{{-}} \right\rangle})\) which is defined for each \(R\in \mathsf{Alg}_{/ {K^+}}\) as \begin{align*} R \mapsto \left\{{g\in \operatorname{GL}_{K_R}(V_R) {~\mathrel{\Big\vert}~}{\left\langle {gv},~{gw} \right\rangle} = \nu {\left\langle {v},~{w} \right\rangle} \text{ for some } \nu\in R}\right\} .\end{align*} The unitary group is the subgroup for which \(\nu = 1\) is enforced.

If \(R={\mathbb{R}}\), choose an ordered basis for \(B\) to define the signature \begin{align*} {\left\langle {v},~{w} \right\rangle} = vA { {}^{t}{ (w) } }, \qquad A = { \begin{bmatrix} {\one_a} & {0} \\ {0} & {-\one_b} \end{bmatrix} },\qquad \operatorname{sig}(A) \mathrel{\vcenter{:}}=(a, b) .\end{align*}

For the remainder of today, assume \(K^+ = {\mathbb{Q}}\).

3.3 Automorphic forms on unitary groups, connections to modular forms

On the modular form side:

  1. \(f: {\mathfrak{h}}\to {\mathbb{C}}, f(z) = (cz+d)^{-k}f(\gamma z)\), holomorphic at cusps, etc
  2. \begin{align*} \phi_f: {\operatorname{SL}}_2({\mathbb{R}})\to {\mathbb{C}}, {\operatorname{SL}}_2({\mathbb{R}})\curvearrowright{\mathfrak{h}} .\end{align*} transitively fixing \(i\), \begin{align*} \phi_f(g) = j(g, i)^{-k}f(gi) ,\end{align*} and \begin{align*} \phi_f: \dcosetl{\Gamma}{G({\mathbb{R}})} \to {\mathbb{C}}\\ \phi_f(g(\operatorname{rot}(\theta)) = e^{ki\theta} \phi_f(g) \qquad \operatorname{rot}(\theta) \mathrel{\vcenter{:}}= \left[\begin{array}{cc} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{array}\right] ,\end{align*} extend to \begin{align*} \phi: \dcosetl{\Gamma Z(G)}{G({\mathbb{R}})} \to {\mathbb{C}} \end{align*} for \(G=\operatorname{GL}_2, {\operatorname{SL}}_2, \operatorname{GL}_2^+\), etc
  3. Adelic interpretation: \(\operatorname{GL}_2({\mathbb{A}}) = \operatorname{GL}_2({\mathbb{Q}}) \operatorname{GL}_2^+({\mathbb{R}}) \tilde K\) where \(\tilde K \mathrel{\vcenter{:}}=\prod_p K_p\)is a compact open subgroups of \(\operatorname{GL}_2({ {\mathbb{Q}}_p })\) with determinant \({ {\mathbb{Z}}_p }^{\times}\) and equal to \(\operatorname{GL}_2({ {\mathbb{Q}}_p })\) for all but finitely many places.

On the automorphic side:

  1. Replace \({\mathfrak{h}}\) with \(G/{\mathcal{K}}_\infty = {\operatorname{U}}_{n, m}({\mathbb{R}})/ {\operatorname{U}}_n{ {}^{ \scriptscriptstyle\times^{2} } }\), a quotient by a compact.
  2. Writing \(G\mathrel{\vcenter{:}}=\operatorname{GU}(n, m)\) for a form of signature \((n,m)\), replace with \(\dcosetl{\Gamma Z(G)}{G}\) where \(G\supseteq{\mathcal{K}}_\infty \mathrel{\vcenter{:}}= U(n){ {}^{ \scriptscriptstyle\times^{2} } }\), and analogously \(\dcosetl{\Gamma Z(G)}{G({\mathbb{R}})} \to {\mathbb{C}}\).
  3. For \(G({\mathbb{A}}_f) = {\textstyle\coprod}_i G({\mathbb{R}})^{-1}? {\mathcal{K}}\).

An automorphic form on \({\operatorname{U}}_{n, n}\) is a holomorphic function \(f\in {\mathfrak{h}}_n\to V\) where \(\rho\curvearrowright V\) is a representation of \(\operatorname{GL}_n({\mathbb{C}}){ {}^{ \scriptscriptstyle\times^{2} } }\) where \begin{align*} f(z) = \rho(cz+d, { {}^{t}{ (\mkern 1.5mu\overline{\mkern-1.5muc\mkern-1.5mu}\mkern 1.5mu) } } z + \mkern 1.5mu\overline{\mkern-1.5mud\mkern-1.5mu}\mkern 1.5mu)^{-1}f( \gamma z)\qquad \forall \gamma \in \Gamma = { \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} }\in {\operatorname{U}}_{n, n}({\mathcal{O}}_K) \end{align*} where \begin{align*} \gamma z = (az+b)(cz+d)^{-1}, \qquad {\mathfrak{h}}_n \mathrel{\vcenter{:}}=\left\{{z\in \operatorname{Mat}_n({\mathbb{C}}) {~\mathrel{\Big\vert}~}i({ {}^{t}{ (\mkern 1.5mu\overline{\mkern-1.5muz\mkern-1.5mu}\mkern 1.5mu) } } - z) > 0}\right\} .\end{align*}

One thing we haven’t mentioned yet: modular forms as sections of line bundles over modular curves, so moduli of elliptic curves with level structure, and the generalized setup will be vector bundles over (unitary) Shimura varieties.

4 Ellen Eischen, Talk 2

Today: more on automorphic forms, and approaches to studying \(L(s, \pi)\) for \(\pi\) a cuspidal representation of a unitary group. Note that we’ve been taking the adelic approach, see Wee Tek’s talk for how to define \(L{\hbox{-}}\)functions in this setting.

Several perspectives on automorphic forms on unitary groups:

Today: sections of a vector bundle over certain moduli spaces.

See Shimura’s first paper on Rankin-Selberg convolutions! See also two papers by Siegel’s student that define the generalizations \({\mathfrak{h}}_n\).

4.1 Modular Forms

Let \({\mathcal{M}}\) be a modular curve (parameterizing curves with some level structure) and let \(\xi\to {\mathcal{M}}\) be the universal elliptic curve. Write \(\Omega_{\xi/{\mathcal{M}}}^1\) for the relative differentials, and define \begin{align*} \omega \mathrel{\vcenter{:}}=\pi_* \Omega^1_{\xi/{\mathcal{M}}} .\end{align*} A modular form is section of a tensor power of \(\omega\), i.e. an element of \(H^0({\mathcal{M}}; \omega{ {}^{ \scriptstyle\otimes_{}^{k} } })\).

We can regard a modular form as a rule \((E, \omega) \mapsto F(E, \omega)\in {\mathbb{C}}\) that transforms like \begin{align*} F(E, \lambda \omega) = \lambda^{-k} F(E, \omega) \qquad\forall \lambda\in {\mathbb{C}}^{\times} .\end{align*} Equivalently, a rule \(\tilde F\) that maps \(E\) to some \(\omega\in \Omega_{E/{\mathbb{C}}}\), e.g.  \begin{align*} \tilde F(E) = F(E, \omega)\omega{ {}^{ \scriptstyle\otimes_{}^{k} } } .\end{align*}

Connecting this with last time: \begin{align*} (E, \omega) \rightleftharpoons\Lambda_{(E, \omega)} \rightleftharpoons{\mathbb{Z}}+ \tau {\mathbb{Z}}\\ F(E, \omega) \longrightarrow\cdots \longrightarrow f_F(\tau) ,\end{align*} i.e. such a rule can be regarded as a function on lattices.

Similarly, automorphic forms arise as global sections of a vector bundle over a unitary Shimura variety \({\mathcal{M}}\) parameterizing abelian varieties with

One can similarly identify \begin{align*} {\mathcal{M}}({\mathbb{C}}) \cong \dcoset{G({\mathbb{Q}})}{G({\mathbb{A}})}{{\mathcal{K}}\cdot {\mathcal{K}}_\infty} ,\end{align*} which will be a finite disjoint unions of copies of symmetric spaces (e.g. \({\mathfrak{h}}_n\)) for \({\operatorname{U}}_{a, b}\).

Write \(\underline{A}\) for an abelian variety with some extra structure, one can then also view an automorphic form as a function \begin{align*} F(\underline{A}, \ell) = \rho(tg)^{-1}F(\underline{A},\ell) \qquad \forall G\in \operatorname{GL}_a \times \operatorname{GL}_b ,\end{align*} where \(\ell = (\ell_+, \ell_-)\) is an ordered basis for \(\Omega_{A/{\mathbb{C}}}\) that decomposes according to the signature.

Let \({\mathcal{A}}\xrightarrow{\pi} {\mathcal{M}}\) be the universal family and define the sheaf \(\omega \mathrel{\vcenter{:}}=\pi_* \Omega_{{\mathcal{A}}/{\mathcal{M}}}\); one can then build a sheaf of automorphic forms \(\omega^p\) in much the same way. This reformulates the notion of an automorphic form in terms of lattices and functions on symmetric spaces like \({\mathfrak{h}}_n\).

4.2 ?

Goal for today: introduce an approach to studying certain \(L{\hbox{-}}\)functions using the doubling method. Note that the example of looking at the constant term of an Eisenstein series from yesterday turns out to be deceptively simple, hence a different approach today.

Let - \(f(q) = \sum_{k\geq 1} a_kq^k\) be a weight \(k\) cusp form and - \(g(q) = \sum_{k\geq 0} b_kq^k\) be a weight \(\ell\) modular form,

where \(a_k, b_k \in { \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Q} \mkern-1.5mu}\mkern 1.5mu }\). The Rankin-Selberg product is \begin{align*} D(s,f,g) = \sum_{n\geq 1} a_n b_n n^{-s} .\end{align*}

Shimura proved that \begin{align*} { D(m,f,g) \over {\left\langle {f},~{f} \right\rangle}_{{\mathrm{Pet}}} } \in \pi^k { \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Q} \mkern-1.5mu}\mkern 1.5mu }\qquad \text{for } \ell < k,\, {k+\ell-2\over 2} < m < k .\end{align*}

This prove relies on the realization \begin{align*} D(k-1-r, f, g) = c\pi^k {\left\langle {\tilde f},~{\, g\delta_{ \lambda}^{(r)} E } \right\rangle}_{{\mathrm{Pet}}} ,\end{align*} where

Note that the weight-raising operator doesn’t preserve holomorphicity, which can be bad for algebraicity results, but it turns out that the result is “almost holomorphic.”

4.3 Proving Algebraicity: A Recipe

A general strategy:

Note that each step is highly nontrivial, and in some contexts, some steps haven’t even been completed yet. The third step often involves working one place at a time. Even having all three may not be enough, sometimes the results one gets aren’t amenable to algebraic/geometric study and are instead only good for analytic purposes.

4.4 Doubling

Setup:

Next time: we’ll introduce the doubling integral after pairing with an Eisenstein series.

5 Ellen Eischen, Talk 3

Some references on doubling:

5.1 Eisenstein Series

We’ll continue with the previous setup. Let \(P \leq H\) be parabolic preserving \(V^\Delta \mathrel{\vcenter{:}}=\left\{{(v,v) {~\mathrel{\Big\vert}~}v\in V}\right\}\). We get a decomposition \(W = V^\Delta \oplus V_\Delta\) where \(V_\Delta = \left\{{(v, -v) {~\mathrel{\Big\vert}~}v\in V}\right\}\). Then \begin{align*} P = \left\{{ { \begin{bmatrix} {A} & {*} \\ {0} & {{ {}^{t}{ (\mkern 1.5mu\overline{\mkern-1.5muA\mkern-1.5mu}\mkern 1.5mu^{-1}) } }} \end{bmatrix} } {~\mathrel{\Big\vert}~}A\in \operatorname{GL}_n}\right\} .\end{align*}

Given a Hecke character \begin{align*} \chi: \dcosetl{K^{\times}}{{\mathbb{A}}_K^{\times}}\to {\mathbb{C}} ,\end{align*} view this as a character of \(P\) via \begin{align*} P\to \operatorname{GL}_n \xrightarrow{\operatorname{det}} {\mathbb{A}}_K^{\times}\to {\mathbb{C}}\\ { \begin{bmatrix} {A} & {*} \\ {0} & {{ {}^{t}{ (\mkern 1.5mu\overline{\mkern-1.5muA\mkern-1.5mu}\mkern 1.5mu^{-1}) } }} \end{bmatrix} } \mapsto A\mapsto \operatorname{det}(A) \xrightarrow{\chi} \chi(\operatorname{det}(A)) .\end{align*}

For \(s\in {\mathbb{C}}\), let \begin{align*} f_{S, \chi} = \operatorname{Ind}_{P({\mathbb{A}})}^{H({\mathbb{A}})} (\chi \cdot {\left\lvert {{-}} \right\rvert}^{-s} ) = \left\{{f: H({\mathbb{A}})\to {\mathbb{C}}{~\mathrel{\Big\vert}~}f(ph) = \chi(p) {\left\lvert {p} \right\rvert}^{-s} f(h)}\right\} .\end{align*} Define the Siegel Eisenstein series for \(g\in H\) by \begin{align*} E_{F_s, \chi}(g) = \sum_{\gamma \in \dcosetl{ P({\mathbb{Q}}) }{ H({\mathbb{Q}}) } } f_{s, \chi}(\gamma g) .\end{align*}

5.2 Doubling Integral

Setup:

Define a zeta integral \begin{align*} Z(\phi, \tilde\phi, f_{s, \chi}) \mathrel{\vcenter{:}}=\int_{ \dcosetl{(G\times G)({\mathbb{Q}})}{(G\times G)({\mathbb{A}})} } E_{f_{s, \chi}} (g_1, g_2) \phi(g_1) \tilde \phi(g_2) \chi^{-1}(\operatorname{det}g_2) \,dg _1 \,dg _2 ,\end{align*} which is against an appropriately normalized Haar measure.

Most analytic properties of \(E\) will carry over to \(Z\), e.g. the functional equation and meromorphic continuation – this is a common theme! In the case of \(G = \operatorname{GU}_1\) or \(\operatorname{GU}_n\) a definite unitary group, one can express \(Z\) as a finite sum.

\begin{align*} Z(\phi, \tilde\phi, f_{s, \chi}) = \int_{G({\mathbb{A}})} f_{s, \chi}((g, 1)) {\left\langle {\pi(g) \phi},~{\tilde\phi} \right\rangle} \,dg , \qquad {\left\langle {\phi},~{\tilde\phi} \right\rangle} = \int_{\dcosetl{G({\mathbb{Q}})}{G({\mathbb{A}})}} \phi(g)\tilde\phi(g) \,dg ,\end{align*} where this pairing is \(G{\hbox{-}}\)invariant which is unique up to a constant multiple.

If there is a restricted tensor product representation \begin{align*} \pi = \bigotimes^{\operatorname{res}}_v \pi_v, \qquad \tilde \pi = \bigotimes^{\operatorname{res}}_v \tilde \pi_v, \end{align*} where \(\operatorname{Ind}(\chi{\left\lvert {{-}} \right\rvert}^v)= \bigotimes O_v\) with

then there is an Euler product decomposition \begin{align*} Z(\phi,\tilde\phi, f_{s,\chi}) = \prod_v Z_v(\phi_v, \tilde\phi_v, f_v) \end{align*} where \begin{align*} Z_v(\phi_v, \tilde\phi_v, f_v) \mathrel{\vcenter{:}}=\int_{G({\mathbb{Q}}_v)} f_v((g, 1)) {\left\langle {\pi_v(g) \phi_v },~{ \tilde\phi_v} \right\rangle} \,dg .\end{align*}

By the uniqueness of the invariant pairing, there must exist a decomposition \begin{align*} {\left\langle {\phi},~{\tilde\phi} \right\rangle} = \prod_v {\left\langle {\phi_v},~{\tilde\phi_v} \right\rangle} .\end{align*}

Shimura computed coefficients in many cases.

Outline proof of the theorem: we’ll analyze the orbits of \(G\times G\curvearrowright X\mathrel{\vcenter{:}}=\dcosetl{P}{H}\). Setup:

Idea: write \(E\) as a sum and rearrange, then reduce to a known computation. Reexpress it as \begin{align*} E_{f, \chi} (h) &= \sum_{\scriptscriptstyle [\gamma] \in \dcoset{P({\mathbb{Q}})} {H({\mathbb{Q}})} {(G\times G)({\mathbb{Q}})} } \qty{ \sum_{ \scriptscriptstyle \dcosetl{(G\times G)({\mathbb{Q}})} {(G\times G)({\mathbb{Q}})} } f_{s, \chi}( \gamma h)} \\ &= \sum_{\scriptscriptstyle [\gamma] \in \dcoset{P({\mathbb{Q}})} {H({\mathbb{Q}})} {(G\times G)({\mathbb{Q}})} } \qty{ \sum_{ \scriptscriptstyle \dcosetl{(G\times G)({\mathbb{Q}})} {(G\times G)({\mathbb{Q}})} } \int_{ \scriptscriptstyle \dcosetl{(G\times G)({\mathbb{Q}})} {[G\times G]({\mathbb{A}})} } f_{s, \chi}((g, h)) \phi(g) \tilde\phi(h) \chi^{-1}\operatorname{det}(h) \,dg \,dh }\\ &\mathrel{\vcenter{:}}= \sum_{\scriptscriptstyle [\gamma] \in \dcoset{P({\mathbb{Q}})} {H({\mathbb{Q}})} {(G\times G)({\mathbb{Q}})} } I(\gamma) ,\end{align*} where \begin{align*} I(\gamma) \mathrel{\vcenter{:}}= \int_{\dcosetl{[G\times G]^\gamma ({\mathbb{Q}})} {[G\times G]({\mathbb{A}})} } f_{s, \chi}(\gamma(g, h)) \phi(g)\tilde\phi(g) \chi^{-1}\operatorname{det}(g)\,dg \,dh .\end{align*}

We’ll have to analysis the \(\gamma = 1\) and \(\gamma\neq 1\) cases separately, they’re quite different:

Rewrite the stabilizer in a more convenient way: \begin{align*} [G\times G]^{\gamma} &= \left\{{(g,h) \in G\times G{~\mathrel{\Big\vert}~}P\gamma(g, h) = P\gamma}\right\} \\ &= \left\{{(g, h)\in G\times G{~\mathrel{\Big\vert}~}\gamma(g, h)\gamma^{-1}\in P}\right\} ,\end{align*} so \begin{align*} [G\times G]^1 = P \cap(G\times G) = \left\{{(g,g){~\mathrel{\Big\vert}~}g\in G}\right\} \mathrel{\vcenter{:}}= G^\Delta .\end{align*} Thus \begin{align*} f_{s, \chi}(1\cdot(g, h)) &= f_{s, \chi}(g, h) \\ &= f_{s, \chi}((h,h) \cdot (h^{-1}g, 1)) \\ &= \chi(\operatorname{det}h) f_{s,\chi}(h^{-1}g, 1) \end{align*} and \begin{align*} I(1) = \int_{\dcosetl{ G^\Delta({\mathbb{Q}}) } {(G\times G)({\mathbb{A}})} } f_{s, \chi}(h^{-1}g, 1) \phi(g)\tilde\phi(g) \,dg \,dh .\end{align*} We have an identification \begin{align*} G\times G &\rightleftharpoons G^\Delta \times(G\times 1) \rightleftharpoons G\times G \\ (g,h) &\rightleftharpoons(h,h)\cdot(h^{-1}g, 1) \rightleftharpoons(h,h^{-1}g) \mathrel{\vcenter{:}}=(g,g_1) ,\end{align*} which we can use to write \begin{align*} I(1) &= \int_{G({\mathbb{A}})} \int_{\dcosetl{G({\mathbb{Q}})} {G({\mathbb{A}})} } f_{s,\chi}(g_1, 1) \pi(g_1) \pi(h) \tilde\phi(h)\,dh \,dg _1 \\ &= \int_{G({\mathbb{A}})} f_{s, \chi}(g_1, 1) {\left\langle {\pi(g_1) \pi},~{\tilde\phi} \right\rangle} \,dg _1 .\end{align*}

All other orbits \([\gamma]\neq [1]\) decompose to products including terms of the form \begin{align*} \int_{\dcosetl{N_i({\mathbb{Q}})} {N_i({\mathbb{A}})} } \phi_i(n\cdot g)\,dn \end{align*} for \(i=1,2\), \(\phi_1 = \phi, \phi_2 = \tilde \phi\) and \(N_i\) unipotent radicals of a parabolic subgroup of \(G\) which is nontrivial for at least on term – however, these are cuspidal, so such integrals vanish (essentially by definition), making the entire thing vanish.

This falls into step 1 of the overall strategy – we found a pairing. So the next question is step 2: can we choose \(f_{s, \chi}, \phi,\tilde\phi\) so that we nice multiples of Langlands \(L{\hbox{-}}\)functions \(L(s,\pi,\chi)\)? This will rely on reducing to computations to local integrals that were computed by Godement and Jacquet for \(\operatorname{GL}_n\).

Next: pulling back automorphic forms to smaller groups, what does this look like for \(n=1\)?

6 Ellen Eischen, Talk 4: Revisiting the Doubling Method for \(n=1\)

6.1 Reducing to Finite Sums

Goal: see what happens if we do the doubling method in the following setup. Let

Spoiler: we’ll get an expression for the \(L{\hbox{-}}\)function \(L(s, \chi)\) for a Hecke character \(\chi: \dcosetl{K^{\times}}{{\mathbb{A}}_K^{\times}} \to CC^{\times}\) as a finite sum of values \(E_\chi(A) \chi(A)\) for some elliptic curves \(A\) with CM by \({\mathcal{O}}_K\), and we’ll obtain an algebraicity result.

Note that \begin{align*} \operatorname{GU}_{1, 1} \cong \dcosetr{\operatorname{GL}_2\times \mathop{\mathrm{Res}}_{K/{\mathbb{Q}}} {\mathbb{G}}_m }{{\mathbb{G}}_m} ,\end{align*} and the associated symmetric space consists of copies of the upper half plane \({\mathfrak{h}}_1\). The associated modular form is a modular form, possibly with mild additional conditions on each component.

Reminder of the doubling method: we had an integral \begin{align*} Z(s,\chi,\phi,\tilde\phi) = \int_{\dcosetl{G{ {}^{ \scriptscriptstyle\times^{2} } } ({\mathbb{Q}})} {{\mathbb{G}}{ {}^{ \scriptscriptstyle\times^{2} } }({\mathbb{A}})} } E_{f_S, \chi}(g, h) \phi(g) \tilde \phi(h) \chi^{-1}(\operatorname{det}h) \, dg\, dh .\end{align*}

Some properties:

There is a diagram:

Link to Diagram

Moreover there is an embedding:

Link to Diagram

These induce embeddings of corresponding Shimura varieties:

Recall that the adelic points of our quotients are \({\mathbb{C}}{\hbox{-}}\)points of unitary Shimura varieties, and \({\mathcal{M}}_{\operatorname{GU}_{1,1}}({\mathbb{C}}) = {\textstyle\coprod}\dcosetl{\Gamma_K}{{\mathfrak{h}}_1}\) where we mod out by some level. Any \(z\in {\mathfrak{h}}= {\mathfrak{h}}_1\) corresponds to some \({\mathbb{C}}{ {}^{ \scriptscriptstyle\times^{2} } }/\left\langle{(z\mkern 1.5mu\overline{\mkern-1.5mua\mkern-1.5mu}\mkern 1.5mu + \mkern 1.5mu\overline{\mkern-1.5mub\mkern-1.5mu}\mkern 1.5mu, za + b)}\right\rangle\) where \(a,b\) are in some \({\mathcal{O}}_K\) lattice. Note the similarity to \({\mathbb{C}}/\left\langle{{\mathbb{Z}}+ \tau {\mathbb{Z}}}\right\rangle\) for elliptic curves.

An upshot is that there are three special things in this case:

Let \(Z(s, \chi) \mathrel{\vcenter{:}}= Z(s, \chi, \phi,\tilde\phi)\). We can choose \(f_{S,\chi}\) such that \begin{align*} Z(s,\chi) = c L(s, \chi) \end{align*} for \(c\) a scalar, i.e. they differ by a multiple. This expresses \(L(s, \chi)\) as a finite sum of values of \(E(s,\chi) \cdot \chi({-})\) for \(E\) an automorphic form on \({\operatorname{U}}_{1, 1}\), so a special kind of modular form. There is a variant of Damerell’s formula, which expresses \(L(s, \chi)\) as such a finite sum where \(E\) is an Eisenstein series in a space of Hilbert modular forms.

6.2 Rationality Properties for Eisenstein Series

We can obtain an Eisenstein series on \({\mathfrak{h}}= {\mathfrak{h}}_1\) of the form \begin{align*} \sum_{(c,d)\in \Lambda} {\chi(d) \over (cz+d)^k (cz+d)^s} \end{align*} where \(\Lambda\) is an appropriate \({\mathcal{O}}_K\) lattice, and for certain characters will converge for \(\Re(s) + k > 2 = 2n\). This will have rational Fourier coefficients, and is holomorphic for \(s=0\). As in the case of modular forms, there is a \(q{\hbox{-}}\)expansion (or more generally in other signatures, a Fourier-Jacobi expansion) principle:

Automorphic forms on \({\operatorname{U}}_{n,n}\) are determined by their \(q{\hbox{-}}\)expansions.

In particular, if the coefficients of the \(q{\hbox{-}}\)expansion are contained in \(R\), then \(f\) is in fact defined over \(R\). Kai-Wen Lan proved a more general version of this principle for \({\operatorname{U}}_{a,b}\) with any signature, and showed that algebraic \(q{\hbox{-}}\)expansions and analytic (i.e. Fourier) expansions agree.

So things look good for \(s=0\)!

What about \(s\neq 0\), i.e. when the Eisenstein series is not holomorphic?

We use Mass-Shimura differential operators \(\delta^{(r)}_K\) to relate \(E\) at \(s\neq 0\) to \(E\) at \(s=0\), where here \(\delta\) raises weights by \(2r\). For \(F\) a modular form defined over \({ \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Q} \mkern-1.5mu}\mkern 1.5mu }\), Shimura proved the following: \begin{align*} { (\delta_k^{(r)} F)(A) \over \Omega^{k+2r} } \in { \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Q} \mkern-1.5mu}\mkern 1.5mu } \end{align*} for each CM point \(A\). These operators have incarnations in \({\operatorname{U}}_{n, m}\) and there are analogous algebraicity results. In fact, \begin{align*} E(z, -r, \chi) = c (-4\pi y)^r \delta_k^{(r)} E(z, 0,\chi) .\end{align*} where \(c\) is a nice rational factor. Combining these results yields \begin{align*} {L(r, \chi) \over \Omega^{k+2r}}\in {\mathbb{Q}} .\end{align*}

A word about this operator: \begin{align*} \delta_k f = {1\over 2\pi i} \qty{{k\over 2iy} + {\frac{\partial }{\partial z}\,} } f = {1\over 2\pi i}y^{-k} {\frac{\partial }{\partial z}\,} (y^k f), \qquad \delta_k^{(r)} = \delta_k \circ \delta_k \circ \cdots \circ \delta_k .\end{align*} Katz’s idea: reexpress this operator geometrically over a moduli space of elliptic curves, or more generally AVs, in terms of the Gauss-Manin connection and the Kodaira morphism, and a splitting \begin{align*} H^1_{\mathrm{dR}} = \omega \oplus H^{0, 1} \end{align*} which preserves algebraicity at CM points.

7 Wee Teck Gan: Automorphic forms and the theta correspondence (Talk 1)

Goal: reformulating the Ramanujan-Petersson conjecture in terms of representation theory.

7.1 The Ramanujan-Petersson conjecture

Let \(f: {\mathfrak{h}}\to {\mathbb{C}}\) be a holomorphic cusp form of weight \(k\) and level 1. Suppose \(f\) is an eigenvector for the Hecke operator \(T_p\), then \(f\) has a Fourier expansion \begin{align*} f(z) = \sum_{k\geq 1} a_k(f) q^n, \qquad q\mathrel{\vcenter{:}}= e^{2\pi i z} ,\end{align*} which can be normalized so that \(a_1(f) = 1\). The remaining coefficients are then the Hecke eigenvalues, so \begin{align*} T_p f = a_p(f) f .\end{align*}

\begin{align*} {\left\lvert {a_p(f) } \right\rvert} \leq 2p^{k-1\over 2} .\end{align*} This was proved by Deligne as a consequence of the Weil conjectures. There is an analog for Maass forms, which involves the hyperbolic Laplacian, which similarly bounds Fourier coefficients.

Error terms come from the cusp forms here. There is a bridge that takes holomorphic modular forms and Maass forms to the world of automorphic forms.

Setup: let \(k\in \mathsf{Field}_{/ {{\mathbb{Q}}}} , v \in {\operatorname{Places}}(K)\) so that \(k_v\in\mathsf{Loc}\mathsf{Field}\). Define the adeles as \({\mathbb{A}}\mathrel{\vcenter{:}}=\prod^{\operatorname{res}}k_v\) which admits a diagonal embedding \(k\hookrightarrow{\mathbb{A}}\) with \(\dcosetl{k}{{\mathbb{A}}}\) compact. Let \(G\in{\mathsf{Alg}}{\mathsf{Grp}}_{/ {k}}\) be reductive, e.g. \({\operatorname{SL}}_n, {\operatorname{U}}_n\), we then similarly have \begin{align*} G(k) \hookrightarrow G({\mathbb{A}}) = \prod^{\operatorname{res}}_v G(k_v) \end{align*} with \(\left\{{k_v}\right\}\) an open compact subgroup. Write \([G] = \dcosetl{G(k)}{G({\mathbb{A}})}\), and note the there is a right action \([G] \curvearrowleft G({\mathbb{A}})\).

7.2 Automorphic Reps

An automorphic form on \(G\) is a function \(f: [G]\to {\mathbb{C}}\) satisfying

Write \({\mathcal{A}}(G)\) for the vector space of automorphic forms on \(G\). Note that this carries at left \(G({\mathbb{A}})\) action: \begin{align*} (g_0 . f)(g) = f(g g_0) .\end{align*}

The finiteness condition will guarantee that \(f\) will come from the kernel of a differential operator, e.g. the CR equations for holomorphy. Requiring \(K{\hbox{-}}\)finiteness only gives an action on finite adeles.

An automorphic representation is an irreducible representation of \({\mathcal{A}}(A)\).

7.3 Cusp Forms

A form \(f\in {\mathcal{A}}(G)\) is cuspidal iff for all parabolic subgroups \(P\) with \(P=MN\), the constant term of \(f\) along \(N\) is zero, where the constant term is defined as \begin{align*} f_N(g) = \int_{[N]} = f(ng) \,dn .\end{align*} This yields a subspace of cusp forms \({\mathcal{A}}_{ \mathrm{cusp} }(G) \leq {\mathcal{A}}(G)\) which is stable under the \(G({\mathbb{A}})\) action.

One can take a character \(\psi: [N]\to {\mathbb{C}}^{\times}\), then there is a \((N, \psi){\hbox{-}}\)Fourier coefficient of \(f\): \begin{align*} f_{N, \psi}(g) = \int_{[N]} \mkern 1.5mu\overline{\mkern-1.5mu\psi(n)\mkern-1.5mu}\mkern 1.5mu \cdot f(ng) \,dn .\end{align*}

Uniform moderate growth and being cuspidal imply that \(f\in {\mathcal{A}}_{ \mathrm{cusp} }(G)\) rapidly decays at \(\infty\), i.e. faster than \(1/p\) for any polynomial, so that \(f\in L^2\): \begin{align*} \int_{[G]} {\left\lvert {f} \right\rvert}^2 < \infty .\end{align*} So define the Hilbert space of square-integrable automorphic forms \begin{align*} {\mathcal{A}}_2(G) \mathrel{\vcenter{:}}=\left\{{f\in {\mathcal{A}}(G) {~\mathrel{\Big\vert}~}{\left\lVert {f} \right\rVert}_{L^2}<\infty }\right\} .\end{align*} There is a containment \begin{align*} {\mathcal{A}}_{ \mathrm{cusp} }(G) \subseteq {\mathcal{A}}_2(G) \subseteq {\mathcal{A}}(G) ,\end{align*}

where there are a decomposition into irreducible reps

A main question for automorphic representations: for which \(\pi\) is \(m(\pi) > 0\)? I.e. which representations occur as cuspidal or \(L^2\) reps? Moreover, what do all of the irreducible reps of \(G({\mathbb{A}})\) look like?

Recall that since \(G({\mathbb{A}}) = \prod^{\operatorname{res}}_v G(k_v)\), we expect a representation \(\pi\) of \(G({\mathbb{A}})\) to break up as \(\pi = \bigotimes^{\operatorname{res}}_v \pi _v\) with

7.4 Unramified Reps

There is a containment \(G_v \supseteq K_v\) where \(G_v\) is unramified, i.e. quasi-split (so has a Borel) and split by an unramified extension of \(k_v\), and \(K_v\) is a hyper-special subgroup, which is a maximal compact. This yields \(G_v \supseteq B_v\supseteq T_v N_v\), and there is a bijection \begin{align*} {\mathsf{Irr}}{\mathsf{Rep}}(G_v)(K_v{\hbox{-}}\text{unramified}) &\rightleftharpoons \left\{{\text{Unramified characters of }T_v}\right\}/W \\ I(\chi) = \operatorname{Ind}_{B_v}^{G_v} \chi &\mapsfrom \chi ,\end{align*} where we mod out by a Weyl group action \(W\). Note that \(I(\chi)\) is the unique unramified subquotient.

There is a further correspondence \begin{align*} \left\{{\text{Unramified characters of }T_v}\right\}/W \\ &\underset{ {\text{Langlands}} }{\rightleftharpoons} \left\{{\text{Semisimple conjugacy classes in } G {}^{ \vee }({\mathbb{C}})}\right\} \\ \chi &\mapsto S_\chi ,\end{align*} so there is some semisimple conjugacy class associated to characters \(\chi\).

Thus for \(\pi\in {\mathcal{A}}_{ \mathrm{cusp} }(G)\) with \(\pi = \bigotimes^{\operatorname{res}}_v \pi_v\), one gets a collection \(\left\{{S_{\pi_v} {~\mathrel{\Big\vert}~}v\not\in S}\right\} \subseteq G {}^{ \vee }\). For \(R: G {}^{ \vee }\to \operatorname{GL}_N({\mathbb{C}})\), we can form an \(L{\hbox{-}}\)function \begin{align*} L^S(s, \pi R) \mathrel{\vcenter{:}}=\prod_{v\not\in S} L(s, \pi_v, R),\qquad L(s, \pi_v, R) \mathrel{\vcenter{:}}={1\over \operatorname{det}1-q_v^{-s} R( S_{\pi_v} ) }, \quad q\mathrel{\vcenter{:}}=? .\end{align*} These generalize Hecke \(L{\hbox{-}}\)functions and those attached to modular forms.

7.5 Tempered Reps

A character of the torus \(\chi: T_v\to {\mathbb{C}}^{\times}\) yields \(\pi_\chi\) a \(K_v{\hbox{-}}\)unramified irrep. Say \(\pi_\chi\) is tempered iff \(\chi\) is unitary, i.e. it factors as \(\chi: T_v\to S^1\) so that \({\left\lvert {\chi} \right\rvert} = 1\). Tempered reps naturally occur as regular representations.

Note that tempered reps are weakly contained in \(L^2(G_v)\), but not e.g. the trivial representation of \({\operatorname{SL}}_2\) is not in \(L_2({\mathbb{R}})\), but \({\operatorname{SL}}_2({\mathbb{R}})\) does not have finite volume. In general, the trivial representation is not tempered unless the group is compact.

Let \(\pi = \bigotimes^{\operatorname{res}}_v \subseteq {\mathcal{A}}_{ \mathrm{cusp} }(G)\) for \(G\) quasi-split (or split), then \(\pi_v\) is tempered for almost all \(v\).

This conjecture is false! There is a counterexample for \(G = {\operatorname{SP}}_4\), and a goal for this course is to construct a counterexample for \(G = {\operatorname{U}}_3\).

If \(\pi \subseteq {\mathcal{A}}_{ \mathrm{cusp} }(G)\) and \(\pi\) is globally generic (a certain big enough Fourier coefficient), then \(\pi_v\) is tempered for almost all \(v\).

7.6 Unitary groups

Let

Then define the unitary group as \begin{align*} {\operatorname{U}}(V) \mathrel{\vcenter{:}}=\mathop{\mathrm{Aut}}(V, {\left\langle {{-}},~{{-}} \right\rangle}) .\end{align*}

There are some invariants:

Henceforth we’ll take \(V\) to be Hermitian and \(W\) to be skew-Hermitian.

8 Wee Teck Gan (Talk 2)

Correction from last time: \begin{align*} {\left\langle {v_2},~{v_1} \right\rangle} = {\varepsilon}{\left\langle {v_1},~{v_2} \right\rangle}^c .\end{align*}

Notation from last time:

Over \(p{\hbox{-}}\)adic fields, \({\operatorname{disc}}(V)\) determines \(V\). By composing with a quadratic character \(w_{E/F}\), we obtain \begin{align*} {\operatorname{disc}}\circ w_{E/F} = (V \xrightarrow{{\operatorname{disc}}} F^{\times}/\operatorname{Nm}(E^{\times})) \xrightarrow{w_{E/F}} \left\langle{\pm 1}\right\rangle ,\end{align*} so there are exactly two classes of Hermitian vector spaces of a given dimension, which we’ll denote \(V^+, V^-\).

Over a real field, this is not enough – one also needs the signature \(\operatorname{sig}(V) = (p, q)\) where \(p+q=n\), in which case \begin{align*} {\operatorname{disc}}(V_{p, q}) = (-1)^q (-1)^{p\choose 2} .\end{align*} For \(E/K\) an extension of number fields, there is a local-global principle: \begin{align*} {\mathsf{Herm}}{ \mathsf{Vect} }_{/ {K}} &\hookrightarrow\prod_{v\in {\operatorname{Places}}(K)} {\mathsf{Herm}}{ \mathsf{Vect} }_{/ {K_v}} \\ V &\mapsto \left\{{V\otimes_K K_v}\right\}_{v\in {\operatorname{Places}}(K)} .\end{align*}

We’ll call spaces in the image of this correspondence coherent.

\(V\) is coherent iff for almost every place \(v\), \begin{align*} V\text{ is coherent} \iff {\varepsilon}(V_v) = 1 { \text{a.e.} }\text{ and } \prod_v {\varepsilon}(V_v) = 1 .\end{align*}

Let \(k\) be a \(p{\hbox{-}}\)adic field.

In rank 1:

In rank 2:

In rank \(2n\):

In rank \(2n+1\):

8.1 Howe-PS: Counterexample to the Ramanujan-Petersson Conjecture

Let \(\dim W = 3\), so \(U(W) = {\operatorname{U}}_3\), then \(\mathop{\mathrm{Res}}_{E/K}(W)\in { \mathsf{Vect} }^{\dim = 6}_{/ {k}}\). The trace to \(K\) yields a symplectic form: \begin{align*} \omega({-}, {-}) \mathrel{\vcenter{:}}=\operatorname{Tr}_{E/k}{\left\langle {{-}},~{{-}} \right\rangle}_W .\end{align*}

There is an embedding \(U(W) \hookrightarrow{\mathsf{Sp}}(\mathop{\mathrm{Res}}_{E/k}(W))\), so \({\operatorname{U}}_3 \hookrightarrow{\mathsf{Sp}}_6\) There is a simple something: \begin{align*} \Omega \subseteq {\mathcal{A}}_2({\mathsf{Sp}}({-})) ,\end{align*} which we’ll call theta functions. Note that \(ZU(W) = E^1 \mathrel{\vcenter{:}}=\left\{{x\in E^{\times}{~\mathrel{\Big\vert}~}\operatorname{Nm}(x) = 1}\right\}\). Consider \(i^* \Omega \subseteq {\mathcal{A}}(U(W))\); there is a central character decomposition \begin{align*} i^*( \Omega) = \bigoplus _{\chi} \Omega_{\chi} ,\end{align*} where the sum is over automorphic characters of \(E^1\).

\(\Omega_\chi\) is an irreducible cuspidal representation, with at most one exception \(\chi\), and this \(\Omega_\chi\) produced a counterexample for the RP conjecture.

A complication: the theta functions don’t live on \({\mathsf{Sp}}_6\), but rather on a double cover, and this leads to many technicalities.

Howe-PS produces a correspondence: \begin{align*} \left\{{\text{Automorphic characters on } E^1\cong {\operatorname{U}}_1}\right\} &\rightleftharpoons\left\{{\text{Automorphic reps of } {\operatorname{U}}_3}\right\} \\ \chi &\mapsto \Omega_{\chi} .\end{align*}

How can one produce an injective map \begin{align*} {\mathsf{Irr}}(G) \hookrightarrow{\mathsf{Irr}}(H) ?\end{align*}

Recall that \begin{align*} {\mathsf{Irr}}(G\times H) = \left\{{\pi \otimes\sigma {~\mathrel{\Big\vert}~}\pi \in {\mathsf{Irr}}(G), \sigma \in {\mathsf{Irr}}(H)}\right\} .\end{align*} The idea to produce this map: find \((G\times H){\hbox{-}}\)reps \(\Omega\) and produce a subset \begin{align*} \Sigma_{ \Omega} = \left\{{(\pi, \sigma) {~\mathrel{\Big\vert}~}\mathop{\mathrm{Hom}}_{G\times H}(\Omega, \pi\otimes\sigma) \neq 1}\right\} \subseteq {\mathsf{Irr}}(G) \times {\mathsf{Irr}}(H) .\end{align*}

Is the correspondence \(\Sigma_{\Omega}\) a graph?

There is a decomposition \begin{align*} \Omega_{G\times H} &= \bigoplus _\pi \bigoplus _\sigma m(\pi, \sigma) \pi\otimes\sigma \\ &= \bigoplus _\pi \qty{\bigoplus _\sigma m(\pi, \sigma) \sigma }\otimes\pi \\ &\mathrel{\vcenter{:}}=\bigoplus _\pi \Theta(\pi) \otimes\pi .\end{align*} Is \(\Theta(\pi)\) an irreducible rep, or zero? If so, this produces a map \begin{align*} \Theta: {\mathsf{Irr}}(G) \to {\mathsf{Irr}}(H)\cup\left\{{0}\right\} .\end{align*}

Upshot: one needs \(\dim \Omega\) to be small. Suppose \(G\times H \to E\), take the smallest non-trivial representation \(\Omega\) of \(E\) and pull it back to \(G\times H\). If \(G\times H \subseteq E\), this can be done by restriction.

The theta correspondence is an instance of all of these ideas.

8.2 The Theta Correspondence

Let

This yields a map of the form \(G\times H\to E\): \begin{align*} {\operatorname{U}}(V) \times {\operatorname{U}}(W) \to {\mathsf{Sp}}(V\otimes_E W) .\end{align*} What is \(\Omega\)? To get small enough weights, one needs to pass to the metaplectic cover \({\operatorname{Mp}}\).

8.3 Metaplectic Groups and Weil Reps

For \(\psi: F\to {\mathbb{C}}^{\times}\) a nontrivial character:

Link to Diagram

Here \(\left\{{\omega_\psi}\right\}\) is the smallest infinite-dimensional representation of \({\operatorname{Mp}}\) and referred to as the Weil representation.

On where this comes from: QM. One looks at the Heisenberg group, uses the Stone-von-Neumann theorem, see 2.3 and 2.4 in the notes.

One needs a lift of the following form:

Link to Diagram

By Xudle, \(\tilde i\) exists and is determined by a pair of characters \((\chi_V, \chi_W)\) of \(E^{\times}\) such that

Such a \(\chi_V\) gives \(\tilde {\operatorname{U}}(W) \to {\operatorname{Mp}}\), and similarly for \(W\).

Set \begin{align*} \Omega_{V, W, \chi_V, \chi_W, \psi} \mathrel{\vcenter{:}}=\tilde\iota_{\chi_V, \chi_W}^*(\omega_\psi) ,\end{align*} which has properties described in the lecture notes.

For \(\pi \in {\mathsf{Irr}}{\operatorname{U}}(V)\), define \begin{align*} \Theta(\pi) \mathrel{\vcenter{:}}={\operatorname{coinv}}_{{\operatorname{U}}(V)}(\Omega \otimes\pi {}^{ \vee }) .\end{align*}

Note that there is a \({\operatorname{U}}(W)\) action on both sides. Moreover, \begin{align*} \mathop{\mathrm{Hom}}( {\operatorname{coinv}}_G (\Omega\otimes\pi {}^{ \vee }), {\mathbb{C}}) \cong \mathop{\mathrm{Hom}}_G(\Omega\otimes\pi {}^{ \vee }, {\mathbb{C}}) \cong \mathop{\mathrm{Hom}}_G(\Omega, ?) .\end{align*}

Define \(\theta(\pi)\) to be the maximal semisimple quotient of \(\Theta(\pi)\). This is a finite length semisimple rep.

Is \(\theta(\pi)\) zero or not?

9 Wee Teck Gan (Talk 3)

Last time: we describe the Howe-PS correspondence \begin{align*} \text{Automorphic characters of } {\operatorname{U}}_1 &\rightleftharpoons \text{Automorphic reps of } {\operatorname{U}}_3 \\ \chi &\mapsto \Omega_\chi .\end{align*} A correction: it’s not true that \(\Omega_\chi\) is cuspidal except for at most one \(\chi\); instead if can be cuspidal for many \(\chi\). We defined \(\Omega, \Theta(\pi)\) with a \({\operatorname{U}}(W)\) action, and Howe duality which took \(\Theta(\pi)\neq 0\) to a unique irreducible quotient \(\theta(\pi)\). Thus \(\Theta: {\mathsf{Irr}}{\operatorname{U}}(V)\hookrightarrow{\mathsf{Irr}}{\operatorname{U}}(W) {\textstyle\coprod}\left\{{0}\right\}\) is injective away from the zero locus.

When is \(\Theta(\pi)\neq 0\)?

Let \(\dim W\) be odd, and label \(W_r^{\varepsilon}= 2r+1\). We know all skew-Hermitian spaces of a particular dimension, so we obtain towers:

Link to Diagram

Note that \(W_{r+1}^+ = W_r^+ \oplus {\mathbb{H}}\).

Which \(\theta_r^{\varepsilon}(\pi)\) are nonzero?

Thus reduces checking infinitely many nonzero conditions to just computing the values of these two numbers. We can reduce this further to just checking one number by the following:

\begin{align*} \dim W_{r_0^+(\pi)}^+ + \dim W_{r_0^-(\pi)} = 2\dim V + 2 .\end{align*}

If \(\dim W^+ + \dim W^- = 2\dim V\), then for any \(\pi\in {\mathsf{Irr}}{\operatorname{U}}(V)\), exactly one of \(\Theta_{W^+}(\pi)\) or \(\Theta_{W^-}(\pi)\) is nonzero.

Take \({\operatorname{U}}_1\times {\operatorname{U}}_1 = {\operatorname{U}}(V) \times {\operatorname{U}}(W_0)\) where \({\operatorname{U}}(V) = E^1\), and let \(\chi\in {\mathsf{Irr}}E^1\). Then \begin{align*} \dim W_{r^+(\chi)}^+ + \dim W_{r^-(\chi)}^- = 4 ,\end{align*} These two dimensions are numbers in \(\left\{{1,3}\right\}\), and exactly one of \(\theta_0^{\pm}(\chi)\) is nonzero, and for \(r>0\) we have \(\theta_r^{\varepsilon}(\chi) \neq 0\). Which \(\theta_0^{\varepsilon}(\chi)\) are nonzero?

\begin{align*} \theta_{V, W_0, \psi}(\pi) \neq 0 \iff {\varepsilon}(v) {\varepsilon}(W_0) = {\varepsilon}_E\qty{ {1\over 2}, \chi_E \chi_W^{-1}, \psi(\operatorname{Trace}_{E/F} (\delta-1)) } \end{align*} where \({\varepsilon}_E\) is the local epsilon factor defined in Tate’s thesis. Here \(\chi_E\) is the composite character \(\chi_E(x) = \chi\qty{x\over x^?}\) defined by

Link to Diagram

The \(\delta\in E_0^{\times}\) appears because a Hermitian space depends on a choice of a traceless element.

Applying Howe-PS to \({\operatorname{U}}_1\times {\operatorname{U}}_3\): let \(V = \left\langle{1}\right\rangle = V_0^+\) and \(\chi \in {\mathsf{Irr}}E^1 = {\mathsf{Irr}}{\operatorname{U}}(V)\). Since \(\dim W^{\varepsilon}= 3\), \(\Omega^{\varepsilon}\) is semisimple and decomposes as \begin{align*} \Omega^{\varepsilon}= \bigoplus _{\chi\in{\mathsf{Irr}}E^1} \chi \otimes\Theta^{\varepsilon}(\chi) .\end{align*}

In fact, \(\Theta^{\varepsilon}(\chi) \hookrightarrow\operatorname{Ind}_B^{{\operatorname{U}}(W)}\qty{ \chi_v {\left\lvert {{-}} \right\rvert}^{- {1\over 2} } \otimes\chi }\) where \(B = \operatorname{diag}(a, b, (a^c)^{-1}) + N^+\) (upper triangular) with \(a\in E^{\times}\) and \(b\in E^1\).

9.1 Global Setting

For \(K\in \mathsf{Field}_{/ {{\mathbb{Q}}}}\), writing \(\theta = \prod_v \theta_v\), one might hope for a map \({\mathsf{Irr}}{\operatorname{U}}(V)({\mathbb{A}}) \to {\operatorname{U}}(W)({\mathbb{A}})\). Instead, we’ll want a map \begin{align*} \theta: \left\{{\text{Automorphic reps of } {\operatorname{U}}(V)}\right\} &\to \left\{{\text{Automorphic reps of } {\operatorname{U}}(W)}\right\} ,\end{align*} i.e. a concrete way to transfer functions from a space \(X\) to a space \(Y\). If \(K\in C(X\times Y)\), we can define \begin{align*} T_K: C(X) &\to C(Y) \\ T_k(f)(y) &\mathrel{\vcenter{:}}=\int_X K(x, y) f(x) \,dx ,\end{align*} so \(K\) acts like a matrix. In our case, we’ll want a lift

Link to Diagram

Here \(\Omega = \tilde{\iota}^* W_\psi\). For \(\pi \in {\mathcal{A}}_{ \mathrm{cusp} }({\operatorname{U}}(V))\), we have a map

Link to Diagram

This yields \begin{align*} w_\psi \otimes\pi &\to {\mathcal{A}}({\operatorname{U}}(V)) \\ \phi \otimes f &\mapsto \theta(\phi, f), \qquad \theta(\phi, f)(g) \mathrel{\vcenter{:}}=\int_{[{\operatorname{U}}(V)]} \theta(\phi)(g, h) \mkern 1.5mu\overline{\mkern-1.5muf(h)\mkern-1.5mu}\mkern 1.5mu\, dh .\end{align*} So define the global theta lift of \(\pi\) as \begin{align*} \Theta(\pi) \mathrel{\vcenter{:}}=\left\langle{\theta(\phi, f) {~\mathrel{\Big\vert}~}\phi\in w_\phi, f\in \pi}\right\rangle \subseteq {\mathcal{A}}({\operatorname{U}}(W)) .\end{align*}

If \(\Theta(\pi) \subset {\mathcal{A}}_2({\operatorname{U}}(W))\) is a proper subset, then \(\Theta(\pi)\) is either zero or isomorphic to \(\bigotimes_v \theta(\pi_v)\).

Let \(\pi \subseteq {\mathcal{A}}_{ \mathrm{cusp} }({\operatorname{U}}(V))\),

  1. There exists a smallest \(r_0 = r_0^{\varepsilon}(\pi)\) such that \(\Theta_{r_0}^{\varepsilon}(\pi) \neq 0\). In this case, \(\Theta_{r_0}^{\varepsilon}(\pi) \subseteq {\mathcal{A}}_{ \mathrm{cusp} }({\operatorname{U}}(W))\).

  2. For all \(r > r_0\),\(\Theta_r^{\varepsilon}(\pi)\neq 0\) and is noncuspidal, i.e. not contained in \({\mathcal{A}}_{ \mathrm{cusp} }({\operatorname{U}}(W))\).

  3. For all \(r\geq \dim V\) in the stable range, \(0\neq \Theta_r^{\varepsilon}(\pi) \subseteq {\mathcal{A}}_2({\operatorname{U}}(W))\). Note that being nonzero follows from 1 and 2.

10 Wee Teck Gan (Talk 4)

Take \(V = \left\langle{1}\right\rangle\) and \(W_r^{\varepsilon}= W_0^{\varepsilon}\oplus {\mathbb{H}}^r\) which has dimension \(2r+1\), and let \(\chi \in {\mathcal{A}}({\operatorname{U}}(V))\). We know \(\Theta_r^{\varepsilon}(\chi)\neq 0\) for all \(r>0\), which is the stable range. Note that \(\Theta_r^{\varepsilon}(\chi) \subseteq {\mathcal{A}}_2( {\operatorname{U}}(W_r^{\varepsilon}))\), i.e. these are square-integrable. What happens when \(r=0\)?

\(\Theta_{W_0^{\varepsilon}}(\chi) \neq 0 \iff\) several conditions hold:

Note that \begin{align*} {\varepsilon}_v &= {\varepsilon}(1/2, \chi_E \chi_{W, v}^{-1}, \phi(\operatorname{Trace}?)) \\ 1 = \prod_v {\varepsilon}_v &= {\varepsilon}(1/2, \chi_E \chi_W^{-1}) .\end{align*}

For \(\psi\in W_\phi\), we produce \(\Theta(\phi)\) and obtain an integral \begin{align*} \Theta(\phi, \chi)(g) = \int_{[{\operatorname{U}}(V)]} \phi(g,h) \chi(h)^{-1}\, dh \in {\mathcal{A}}({\operatorname{U}}(W_0)) .\end{align*} Is this function nonzero for some \(\phi\)? There isn’t a good notion of Fourier expansion here, so one instead computes \({\left\langle {\Theta(\phi, \chi)},~{\Theta(\phi, \chi)} \right\rangle}\). Write \(V^\square = V \oplus -V\), where \(-V\) is \(V\) with the form negated.

Link to Diagram

One can them map \({\operatorname{U}}(V^\square) \to {\operatorname{U}}(W_0)^\square\); this diagram is called the doubling see-saw. Combining this with Siegel-Weil associates to the above inner product the doubling zeta integral \(Z(0, \phi, \chi)\). By Ellen’s lectures, this reduces to computing the central value of an \(L{\hbox{-}}\)function, \(cL(1/2, \chi_E \chi_W^{-1})\), up to a fudge factor \(c\). The process is the Rallis inner product formula: \begin{align*} {\left\langle {\Theta(\phi,\chi)},~{\Theta(\phi,\chi)} \right\rangle} \leadsto Z(0,\phi,\chi)\leadsto L(1/2, \chi_E \chi_W^{-1}) .\end{align*}

10.1 Howe-PS

Setup:

We know that the local components are contained in non-tempered principal series, i.e.  \begin{align*} \Theta_W(1)_v \hookrightarrow\operatorname{Ind}_?^{{\operatorname{U}}(W_v)} {\left\lvert {{-}} \right\rvert}_v^+ \otimes 1_v? .\end{align*} It only remains to check that happens when this is not cuspidal. If it is not, then \(\Theta_{W_0}(1) \neq 0\), so pick 2 places \(v_1, v_2\) of \(K\) and swap the signs on \(W_{0, v_i}\) to produce \(W_0'\), and run the above argument on \(W' = W_0' \oplus {\mathbb{H}}\).

10.2 Arthur’s Conjecture

Goal: classify constituents of \({\mathcal{A}}_2(G)\), i.e. describe this as a \(G({\mathbb{A}}){\hbox{-}}\)module. We’ll make a basic hypothesis (global Langlands for \(\operatorname{GL}_n\)) that there exists a group \(L_F\) (thought of as \({ \mathsf{Gal}} (\mkern 1.5mu\overline{\mkern-1.5muF\mkern-1.5mu}\mkern 1.5mu/F)\)) such that there is a bijection \begin{align*} {\mathsf{Irr}}{\mathsf{Rep}}^{\dim = n} L_F \rightleftharpoons{\mathsf{Rep}}_{ \mathrm{cusp} }\operatorname{GL}_n ,\end{align*} where for all \(v\) there is a Weil-Deligne group \(L_{F_v} \approx { \mathsf{Gal}} (\mkern 1.5mu\overline{\mkern-1.5muF\mkern-1.5mu}\mkern 1.5mu_v/F_v)\) with a map \(L_{F_v} \hookrightarrow L_F\).

Two adelic representations \(\pi = \bigotimes_v \pi_v\) and \(\pi' = \bigotimes_v \pi_v'\) are nearly equivalent iff \(\pi_v \cong \pi_v'\) for almost all places \(v\).

We can decompose into near-equivalence classes \({\mathcal{A}}_w(G) = \bigoplus_\psi {\mathcal{A}}_{\psi}\), where \(\psi: L_F\times {\operatorname{SL}}_2 \to {}^{L}{G}\) is a map to the Langlands \(L\) group \({}^{L}{G} = G {}^{ \vee }\rtimes{ \mathsf{Gal}} (\mkern 1.5mu\overline{\mkern-1.5muF\mkern-1.5mu}\mkern 1.5mu/F)\), such that

This has something to do with elliptic \(A{\hbox{-}}\)parameters.

Given \(\psi\), how can we describe \({\mathcal{A}}_\psi\)?

From \(\psi\) we’ll obtain

For all \(v\), we should have a finite set of unitary reps of \(G(F_v)\), \begin{align*} \prod_{\psi_v} = \left\{{\pi_{\eta_v} {~\mathrel{\Big\vert}~}\eta_v \in {\mathsf{Irr}}S_{\psi_v}}\right\} ,\end{align*} i.e. for almost all \(v\), \(\pi_{1_v}\) is irreducible unramified with Satake parameters \begin{align*} \psi_v\qty{\operatorname{Frob}_v, \operatorname{diag}(q_v^{1\over 2} , q_v^{-{1\over 2} })} \in {}^{L}{G} .\end{align*}

The key point: if \(\psi({\operatorname{SL}}_2) = 1\), then \(\pi_{1_v}\) is tempered. If not, \(\psi_{1_v}\) is non-tempered.

This explains how rigidity obstructs the Ramanujan-Petersson conjecture?

Set \(\pi_\psi = \bigotimes_v \pi_{\psi_v}\) and let \begin{align*} {\mathcal{A}}_\psi = \bigoplus_{\eta \in {\mathsf{Irr}}S_{\psi, ?}} m_\eta \pi_\eta, \qquad m_\eta = \dim \mathop{\mathrm{Hom}}_{S_\psi}({\varepsilon}_\psi, \eta) .\end{align*} To define \({\varepsilon}_\psi\), define a map \begin{align*} (L_F\times \Omega_2)\times S_{\psi} \xrightarrow{\psi\times \operatorname{id}} {}^{L}{G} / Z(G {}^{ \vee })^{\Gamma_F} \curvearrowleft{ \operatorname{Ad} }{\mathfrak{g}} {}^{ \vee } \end{align*} where \({\mathfrak{g}} {}^{ \vee }= \mathsf{Lie}(G {}^{ \vee }) = \bigoplus _{i\in I} \rho_i \otimes S_{r_i} \otimes\eta_i\) for some index set \(I\), and \(S_r\) are \(r{\hbox{-}}\)dimensional irreps of \({\operatorname{SL}}_2\). Set \(T \subseteq I\) to be the indices such that \(r_i\) is even, \(\eta_i\) is orthogonal, and \(\rho_i\) is symplectic, and \({\varepsilon}(1/2, \phi_i) = -1\). Then define \begin{align*} {\varepsilon}_\psi: S_\psi &\to \left\langle{\pm 1}\right\rangle \\ s &\mapsto \prod_{i\in T} \eta_i(s) .\end{align*}

For \(\psi({\operatorname{SL}}_2) = \left\{{1}\right\}\), \({\varepsilon}_\psi = 1\) and \(T = \emptyset\).

10.3 Specializing to \({\operatorname{U}}_n\)

Fix \(G={\operatorname{U}}_n\), let \(E/F\) be an extension, and \begin{align*} G {}^{ \vee }= \operatorname{GL}_n({\mathbb{C}}) {~\trianglelefteq~} {}^{L}{G} = \operatorname{GL}_n({\mathbb{C}}) \rtimes{ \mathsf{Gal}} (E/F) .\end{align*} An \(L{\hbox{-}}\)parameter is a map \begin{align*} \psi: L_F \times {\operatorname{SL}}_2 \to {}^{L}{G} \end{align*} where the subset \(L_E \times {\operatorname{SL}}_2\) maps to \(G {}^{ \vee }\). By an email comment of Benedict Gross, \(\psi\) is determined by this restriction. Not every such map extends, but conjugate self-dual reps of sign \((-1)^{n-1}\) will.

For \({\operatorname{U}}_3\), \({ \left.{{\psi}} \right|_{{L_E}} }: L_E\times {\operatorname{SL}}_2 \to \operatorname{GL}_3({\mathbb{C}})\) which decomposes as \({ \left.{{\psi}} \right|_{{L_E}} } = \mu \oplus \chi \otimes S_2\) such that \(\chi\) are not characters of \(L_E\), but rather automorphic characters of \(\dcosetl{E^{\times}}{{\mathbb{A}}_E^{\times}}\) with \({ \left.{{\mu}} \right|_{{{\mathbb{A}}_F^{\times}}} } = 1\) and \({ \left.{{\chi}} \right|_{{{\mathbb{A}}_F^{\times}}} } = \omega_{E/F}\). For \(S_\psi = \mu_2 \xrightarrow{\Delta} \prod_v S_{\psi_v}\), we have \(S_{\psi_v} = \mu_2\) if \(v\) is inert in \(E\) and 1 otherwise. Then \({\varepsilon}_{\psi}: \mu_2 \to \left\langle{\pm 1}\right\rangle\) which is trivial when \({\varepsilon}(1/2, \chi\mu^{-1}) = 1\) and nontrivial if this is \(-1\). So \(\prod_{\psi_v} = \left\{{\pi^+_v, \pi_v^-}\right\}\) if \(v\) is inert, and just \(\left\{{\pi_v^+}\right\}\) otherwise, meaning \begin{align*} m(\pi^{{\varepsilon}}) = \begin{cases} 1 & \prod_v {\varepsilon}_v = {\varepsilon}(1/2) \\ 0 & \text{otherwise}. \end{cases} .\end{align*} For almost every \(v\), \(\pi_{1_v} = \pi_v^+\). Something about \(\operatorname{Ind}_{B_v}^{U_?} \chi {\left\lvert {{-}} \right\rvert}_v^{-{1\over 2}} \otimes\tilde \mu\). Something about Howe-PS.

11 Aaron Pollack: Modular forms on exceptional groups (Lecture 1)

Plans for lectures:

  1. What is \({\mathsf{G}}_2\), what are modular forms on it?
  2. Fourier expansions of modular forms on \({\mathsf{G}}_2\).
  3. Examples and theorems about modular forms on \({\mathsf{G}}_2\).
  4. Beyond \({\mathsf{G}}_2\), possibly \({\mathbf{E}}_8\).

First generalize modular forms to modular functions: let \(f:{\mathfrak{h}}\to {\mathbb{C}}\) be a modular form of level \(\Gamma\) and weight \(\ell>0\). Define \begin{align*} \phi_f: {\operatorname{SL}}_2({\mathbb{R}}) &\to {\mathbb{C}}\\ \phi_f(g) &\mathrel{\vcenter{:}}= j(g, z)^{-\ell} f(gz) \\ \\ j\qty{ g = { \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} },z} &\mathrel{\vcenter{:}}= cz+d .\end{align*}

Some properties:

  1. Growth: \(\phi_f\) is of moderate growth.
  2. Invariance: \(\phi_f(\gamma g) = \phi_f(g)\) for all \(\gamma\in \Gamma \leq {\operatorname{SL}}_2({\mathbb{Z}})\).
  3. Equivariance on a compact: \(k_\theta \mathrel{\vcenter{:}}={ \begin{bmatrix} {\cos \theta} & {-\sin \theta} \\ {\sin \theta} & {\cos \theta} \end{bmatrix} }\in {\operatorname{SO}}_2({\mathbb{R}})\) satisfies \(\phi_f(g k_\theta) = e^{-i\ell \theta} \phi_f(g)\)
  4. Operator equation: \(D_{{\mathrm{CR}}} \phi_f \equiv 0\) where we decompose the complexified Lie algebra \begin{align*} {\mathfrak{sl}}_3({\mathbb{R}}) \otimes_{\mathbb{R}}{\mathbb{C}}\cong (k_0 \otimes{\mathbb{C}}) + (\phi_0\otimes{\mathbb{C}}) \end{align*} as antisymmetric and symmetric parts, then \(p_0\otimes{\mathbb{C}}= {\mathbb{C}}X_+ + {\mathbb{C}}X_-\) where \(X_{\pm} = { \begin{bmatrix} {1} & {\pm i} \\ {\pm i} & {-1} \end{bmatrix} }\), and \(D_{{\mathrm{CR}}} \phi_f \mathrel{\vcenter{:}}= X_- f\).

Conversely, if \(\phi: {\operatorname{SL}}_2({\mathbb{R}}) \to {\mathbb{C}}\) satisfies these properties, then \(f(z) = j(g_z, w^\ell)\phi(g_z)\) where \(g_z \cdot c = z\) is well-defined, holomorphic, weight \(\ell\), level \(\Gamma\) modular forms.

11.1 Modular forms on \({\mathsf{G}}_2\)

Recall that \({\mathsf{G}}_2\) is a simple noncompact Lie group of dimension 14, with maximal compact \(K = ({\operatorname{SU}}_2\times {\operatorname{SU}}_2)/\left\langle{\pm I}\right\rangle\). Write the first factor as \({\operatorname{SU}}_2^l\) for “long” and the second as \({\operatorname{SU}}_2^s\) for “short,” then the root system looks like the following:

There is an action of \(K\) on \(V_\ell \mathrel{\vcenter{:}}=\operatorname{Sym}^\ell({\mathbb{C}}^2)\otimes\one\), and the diagonal acts trivially.

Suppose \(\Gamma\leq {\mathsf{G}}_2\) is a congruence subgroup, so \(\Gamma = {\mathsf{G}}_2({\mathbb{Q}}) \cap K_f\) where \(K_f \subseteq {\mathsf{G}}_2(M_f)\), and let \(\ell \in {\mathbb{Z}}_{>0}\),
A modular form of weight \(\ell\) and level \(\Gamma\) is a map \(\phi: {\mathsf{G}}_2 \to V_\ell\) such that

  1. Growth: \(\phi\) has moderate growth.
  2. Invariance: \(\phi( \gamma g) = \phi(g)\) for all \(\gamma\in \Gamma\).
  3. Equivariance on a compact: \(\phi(gk) = k^{-1}\phi(g)\) for all \(k\in K\).
  4. Operator equation: \(D_\ell \phi = 0\).

Equivalently, a map \(\phi: \dcosetl{{\mathsf{G}}_2({\mathbb{Q}})}{{\mathsf{G}}_2({\mathbb{A}})}\) satisfying similar conditions.

The upshot: modular forms on \({\mathsf{G}}_2\) have a classical Fourier expansion and Fourier coefficients, which appear very arithmetic.

11.2 What is \({\mathsf{G}}_2\)?

Todos:

We’ll define a \(C_3{\hbox{-}}\)graded Lie algebra over \({\mathbb{Q}}\): \begin{align*} {\mathfrak{g}}_2 = {\mathfrak{sl}}_3[0] + V_g({\mathbb{Q}})[1] + V_3 {}^{ \vee }({\mathbb{Q}})[2] ,\end{align*} where \({\mathfrak{sl}}_3\) are the traceless matrices as usual and \(V_3\) is the 3-dimensional standard representation of \({\mathfrak{sl}}_3\). The grading will mean that \([x,y]\) will land in degree \({\left\lvert {x} \right\rvert} + {\left\lvert {y} \right\rvert}\). The bracket is defined as follows: \begin{align*} [\phi, \phi'] &\mathrel{\vcenter{:}}=\phi\phi' - \phi'\phi & \phi,\phi'\in {\mathfrak{sl}}_3 \\ [\phi, v] &\mathrel{\vcenter{:}}=\phi(v), & v\in V_3\\ [\phi, \delta] &\mathrel{\vcenter{:}}=\phi(\delta),&\delta\in V_3 {}^{ \vee } .\end{align*}

\begin{align*} \bigwedge\nolimits^3 V_3= \one \implies \bigwedge\nolimits^2 V_3= V_3 {}^{ \vee } \implies \bigwedge\nolimits^2 (V {}^{ \vee }) = V_3 .\end{align*} Fix a basis \(V_3 = \left\langle{v_1, v_2, v_3}\right\rangle\) and \(V_3 {}^{ \vee }= \left\langle{\delta_1, \delta_2, \delta_2}\right\rangle\) its dual basis, then

Moreover,

\begin{align*} [v, v'] &= 2 v\vee v' \in \bigwedge\nolimits^2 V_3 \cong V_3 {}^{ \vee }\\ [\delta, \delta'] &= 2\delta\vee\delta' \in \bigwedge\nolimits^2 V_3 {}^{ \vee }\cong V_3 \\ [\delta, v] &= 3v\otimes\delta - \delta(v)\one \in {\mathfrak{sl}}_3 ,\end{align*} noting that the last is traceless and \(3v\otimes\delta\in V_3\otimes V_3 {}^{ \vee }\cong \mathop{\mathrm{End}}(V_3)\). All other brackets are determined by antisymmetry and linearity

The algebra \({\mathfrak{g}}_2\) as defined above is a simple Lie algebra, i.e. the Jacobi identity holds and there are no nontrivial ideals. Moreover \begin{align*} \mathop{\mathrm{Aut}}({\mathfrak{g}}_2) = \left\{{g\in \operatorname{GL}({\mathfrak{g}}_2) {~\mathrel{\Big\vert}~}[gx, gy] = g[x, y]\, \forall x, y \in {\mathfrak{g}}_2}\right\} \end{align*} and \({\mathsf{G}}_2 \cong \mathop{\mathrm{Aut}}^0({\mathfrak{g}}_2)\) is the connected component.

Note: a similar procedure can be used to define all of the exceptional groups, see notes.

What is the root diagram for \({\mathfrak{g}}_2\)? Let \({\mathfrak{h}}\leq {\mathfrak{sl}}_3\) be the diagonal elements, i.e.  \begin{align*} {\mathfrak{h}}= \left\{{\sum_{1\leq i\leq 3} \alpha_i E_{ii} {~\mathrel{\Big\vert}~}\sum \alpha_i = 0}\right\} ,\end{align*} and let \(r_1, r_2, r_3: {\mathfrak{h}}\to {\mathbb{Q}}\) be such that \begin{align*} r_j \sum_{1\leq i\leq 3} \alpha_i E_{ii} = \alpha_j ,\end{align*} i.e. projection onto the \(j\)th component. Note that \(\sum r_i = 0\).

What are the weights of \({\mathfrak{h}}\) on \({\mathfrak{g}}_2\)? Since \({\mathfrak{g}}_2 = {\mathfrak{sl}}_3 + V_3 + V_3 {}^{ \vee }\), the actions are:

This yields a root diagram:

On the differential operator: take the Cartan involution \begin{align*} \Theta: {\mathfrak{g}}_2\otimes{\mathbb{R}}\to {\mathfrak{g}}_2 \otimes{\mathbb{R}} .\end{align*}

Explicitly,

Define

Suppose \(\phi: {\mathsf{G}}_2\to V_\ell = \operatorname{Sym}^{2\ell}({\mathbb{C}}^2)\otimes\one\) such that \(\phi(gk) = k^{-1}\phi(g)\) for all \(k\in K\). Let \(\left\{{X_\alpha}\right\}\) be a basis of \(p\), \(\left\{{X_\alpha {}^{ \vee }}\right\}\) basis of \(p {}^{ \vee }\), then \begin{align*} \tilde D_\ell \phi = \sum_\alpha X_\alpha p \otimes X_\alpha {}^{ \vee }\in V_\ell \otimes p {}^{ \vee } .\end{align*} where \(X_\alpha \phi\) is the derivative of the right regular action, i.e. if \(X\in p_0\), \begin{align*} (X_p)(g) = {\frac{\partial }{\partial t}\,} \phi(g\exp(tx))\Big|_{t=0} .\end{align*} Then \begin{align*} V_\ell \otimes\phi {}^{ \vee } &= (S^{2\ell} \otimes\one) \times V_\ell \boxtimes\operatorname{Sym}^3(V_2) \\ &= (S^{2\ell + 1} + S^{2\ell - 1})\boxtimes S^3(V_2) \\ &\xrightarrow{{\operatorname{pr}}} S^{2\ell -1}(V_\ell) \boxtimes S^3(V_3) .\end{align*}

This relates

12 Aaron Pollack, Talk 2

Last time: modular forms on \({\mathsf{G}}_2\). Note that \({\mathsf{G}}_2\) over \(K\) does not have a \({\mathsf{G}}_2{\hbox{-}}\)invariant complex structure, while \({\operatorname{SL}}_2({\mathbb{R}})/{\operatorname{SO}}_2 = {\mathfrak{h}}\) has an \({\operatorname{SL}}_2({\mathbb{R}}){\hbox{-}}\)invariant complex structure.

Today: let \(f(z) = \sum_{k\geq 0} a_f(k)q^k\) of weight \(\ell\) where \(\phi_f(g) = j(g, i)^{-\ell} f(gi)\) where \(\phi_f: {\operatorname{SL}}_2({\mathbb{R}}) \to {\mathbb{C}}\). Define \begin{align*} W_n: {\operatorname{SL}}_2({\mathbb{R}}) &\to {\mathbb{C}}\\ g &\mapsto j(g, i)^{-\ell} \exp(2\pi i n (gi)) .\end{align*}

Some properties:

What will happen: we’ll define \(\phi: \dcosetl{\Gamma}{{\mathsf{G}}_2}\to V_\ell\) where \(V_\ell = \operatorname{Sym}^2({\mathbb{C}}^2)\otimes\one\) which admits an action by \(K = {\operatorname{SU}}_2\times {\operatorname{SU}}_2/\pm I\). In this case, we’ll essentially have \(\phi \approx \sum_{f\in ?} a_\phi(f) W_f(g)\) where the \(a_\pi(f) \in {\mathbb{C}}\) are Fourier coefficients and \(W_f\) satisfies similar properties.

Recall that \({\mathfrak{g}}_2 = {\mathfrak{sl}}_3 + V_3 + V_3 {}^{ \vee }\), spanned by \(\left\{{E_{ij}}\right\}, \left\{{v_1,v_2,v_3}\right\}, \left\{{\delta_1, \delta_2, \delta_3}\right\}\) respectively.

Note that

We want to define a Fourier expansion along the unipotent radical of \(P\).

Some facts:

What are the characters of \(N\)? Suppose

Define \begin{align*} \phi_w(g) \mathrel{\vcenter{:}}=\int_{[N]} \psi^{-1}({\left\langle {w},~{\mkern 1.5mu\overline{\mkern-1.5mun\mkern-1.5mu}\mkern 1.5mu} \right\rangle}) \phi(ng) \,dn ,\end{align*} where \(\mkern 1.5mu\overline{\mkern-1.5mun\mkern-1.5mu}\mkern 1.5mu\) is the image of \(n\) in \(N/Z\) which we identify with \(W\) via the exponential. Similarly define \begin{align*} \phi_Z(g) = \int_{[Z]} \phi(zg)\,dz,\qquad \phi_N(g) = \int_{[N]} \phi(zg)\,dz .\end{align*} Then \begin{align*} \phi_Z(g) = \phi_N(g) + \sum_{w\in W({\mathbb{Q}})} \phi_w(g) ,\end{align*} and we’ll produce a refinement.

\begin{align*} \phi_Z(g) \equiv 0 \implies \phi(g) \equiv 0 .\end{align*}

12.1 Generalized Whittaker Functions

Suppose \(\phi: \dcosetl{{\mathsf{G}}_2({\mathbb{Q}})}{{\mathsf{G}}_2({\mathbb{A}})} \to V_\ell\) is a modular form of weight \(\ell\). These satisfy

Call such functions satisfying these properties general Whittaker functions of type \((w ,\ell)\).

We’ll show that such functions are uniquely determined up to a scalar multiple, i.e. for some explicit \(W_w\), \begin{align*} \phi_w(g) = \lambda W_w(g) .\end{align*} From this, we’ll obtain a Fourier expansion for \(\phi\) a modular form of weight \(\ell\): \begin{align*} \phi_Z(g) = \phi_N(g) + \sum_{w\neq 0} a_{\phi}(w) W_w(g) .\end{align*}

Identify \(W\) as a space \(B\) of binary cubics under \begin{align*} W &\to B \\ w \mathrel{\vcenter{:}}= a E_{12} + {b\over 3} v_1 + {c\over 3} \delta_3 + d E_{23} &\mapsto f_w\mathrel{\vcenter{:}}= au^3 + bu^2v + cuv^2 + dv^3 .\end{align*} For \(w\in W({\mathbb{R}})\setminus\left\{{0}\right\}\), for \(m\in \operatorname{GL}_2({\mathbb{R}})\) define \begin{align*} \beta_w(m) \mathrel{\vcenter{:}}={\left\langle {w},~{m\cdot (u-? v)^3} \right\rangle} ,\end{align*} which will appear in Fourier expansions.

TFAE:

If \(w\) satisfies these properties, say \(w\) is positive semidefinite and write \(w\geq 0\).

For \(m\in \operatorname{GL}_2({\mathbb{R}}) = M({\mathbb{R}})\) and \(w\geq 0\) PSD, \begin{align*} W_w(m) = {\left\lvert {\operatorname{det}w} \right\rvert}\operatorname{det}(w)^\ell \sum_{-\ell \leq v\leq 0} \qty{{\left\lvert {\beta_w(m)} \right\rvert} \over \beta_w(m) }^v K_v({\left\lvert {\beta_w(m)} \right\rvert}) {x^{\ell+v} y^{\ell -v} \over (\ell+v)! (\ell-v)!} .\end{align*} where \(x,y\) are a fixed basis of \(V_\ell\), and \(K_v\) is a classical \(K{\hbox{-}}\)Bessel function \begin{align*} K_v(y) \mathrel{\vcenter{:}}={1\over ?}\int_0^N e^{-{y(t+t^{-1})\over 2 }} t^v {\,dt\over t} ,\end{align*} which diverges at \(y=0\).

These functions \(W_w: M({\mathbb{R}}) \to V_\ell\) extend uniquely to \({\mathsf{G}}_2\to V_\ell\), viz

Suppose \(w\neq 0\) and \(F\) is a generalized Whittaker function of type \((w, \ell)\). Then

Consequently, if \(\phi\) is a modular form on \({\mathsf{G}}_2\) of weight \(\ell\), there exist \(a_\phi(w)\in {\mathbb{C}}\) with \begin{align*} \phi_Z(g) = \phi_N(g) + \sum_{w\geq 0 \text{ integral}} a_\phi(w) W_w(g) .\end{align*} Moreover, \(\phi_N\) can be explicitly described in terms of modular forms of weight \(3\ell\) on \(\operatorname{GL}_2\).

The terms \(a_\phi(w)\) are by definition the Fourier coefficients of \(\phi\).

Gan-Gross-Savim used a multiplicity 1 result of Wallach to define the Fourier coefficients without using the explicit function \(W_w(s)\).

13 Aaron Pollack, Talk 3: Examples of (and theorems about) modular forms on \({\mathsf{G}}_2\)

13.1 Degenerate Eisenstein Series

Recall \({\mathsf{G}}_2\supseteq P\) a Heisenberg parabolic, with \(P = MN\) where \(M\cong \operatorname{GL}_2\). Write \(\nu\) for the composition \(P\to M \xrightarrow{\operatorname{det}} \operatorname{GL}_1\). Suppose \(\ell > 0\) is even, and recall that \(V_\ell = \operatorname{Sym}^2({\mathbb{C}}^2)\otimes\curvearrowleft K \leq {\mathsf{G}}_2\) for \(K\) a maximal compact. Let \begin{align*} f_{\ell, \infty} (g; s) = \operatorname{Ind}_{P({\mathbb{R}})}^{{\mathsf{G}}_2({\mathbb{R}})} {\left\lvert {\nu} \right\rvert}^s \otimes V_\ell \end{align*} be defined by

\begin{align*} f_{\ell, \infty}(pg^j s) &= N(\mu)^s f_{e\ll, \infty} (g; s) \quad \forall p\in P({\mathbb{R}}) \\ f_{\ell,\infty}(gk; s) &= k^{-1}f_{\ell, \infty}(g) \quad \forall k\in K .\end{align*}

By the Iwasawa decomposition \(C_3({\mathbb{R}}) = P({\mathbb{R}}) K\), \(f\) is uniquely determined one we set \begin{align*} f_{\ell, \infty}(1) = x^\ell y^\ell \in V_\ell = \left\langle{x^?, x^?, \cdots, y^{2\ell}}\right\rangle .\end{align*} Let \(f_?\) be a flat section in \(\operatorname{Ind}_{P({\mathbb{A}}_f)}^{{\mathsf{G}}_2({\mathbb{A}}_f)}({\left\lvert {\nu} \right\rvert}^3)\), and let \(f_g(g, s) = f_?(gf, s) f_{\ell,\infty}(g_?; s) \in {\mathsf{G}}_2({\mathbb{A}})\). Define \begin{align*} E_\ell(g,f,s) = \sum_{\gamma\in \dcosetl{P({\mathbb{Q}})}{ {\mathsf{G}}_2({\mathbb{Q}})} } f_\ell(\gamma g, s) .\end{align*} If \(\Re(s) > 3\), set \(E_\ell(g) \mathrel{\vcenter{:}}= E_\ell(g, f, s=\ell+1)\).

If \(\ell>0\) is even and \(\ell\geq 4\), then \(E_\ell(g)\) is a quaternionic modular form on \({\mathsf{G}}_2\) of weight \(\ell\).

\(f_{\ell, ?}(g, s=\ell+1)\) is annihilated by \(D_\ell\), so \(E_\ell(g)\) is as well by absolute convergence.

If \(\pi\) is a cuspidal automorphic representation of \(\operatorname{GL}_2 = M\) associated to a holomorphic weight \(3\ell\) modular form which is cuspidal,

If \(\phi\) is a level 1 quaternionic modular form on \({\mathsf{G}}_2\),

There is a bijection \begin{align*} \left\{{\text{Integral binary cubic forms}}\right\}/\operatorname{GL}_2({\mathbb{Z}}) \rightleftharpoons \left\{{\text{Cubic rings}}\right\}{_{\scriptstyle / \sim} } ,\end{align*} where cubic rings are free rank 3 \({\mathbb{Z}}{\hbox{-}}\)algebras. Thus if \(\ell>0\) is even, \(\phi\) is a level 1 weight \(\ell\) modular form on \({\mathsf{G}}_2\), and \(A\) is a cubic ring, there is a well-defined map \(a_\phi(A) = a_\phi(w)\) if \(A\rightleftharpoons f_w\).

If \(f_w\) is nondegenerate, the cubic ring \(A(f_w)\) associated to \(f_w\) is totally real \(\iff f\) is positive semidefinite.

Suppose \(A\) is the maximal order in a totally real cubic etale \({\mathbb{Q}}{\hbox{-}}\)algebra \(E\). There exists a constant \(c_\ell\in {\mathbb{C}}\), independent of \(A\), such that \begin{align*} a_{E_\ell}(A) = c_\ell \zeta_E(1-\ell) ,\end{align*} where \(\ell\) is even. The LHS are Fourier coefficients of modular forms on \({\mathsf{G}}_2\).

It is not known that \(c_\ell\) is nonzero.

An open question: \(E(g, f_\pi)\) is Eisenstein, can anything be said about its Fourier coefficients.

13.2 Cusp Forms

Suppose \(\ell\geq 16\) is even. There exist nonzero cusp forms on \({\mathsf{G}}_2\) of weight \(\ell\), all of whose Fourier coefficients are algebraic integers.

Steps:

  1. Start with a holomorphic Siegel modular form \(f\) on \({\operatorname{SP}}_4\) of weight \(\ell\), so \(f\) has Fourier coefficients in \({ \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Z} \mkern-1.5mu}\mkern 1.5mu }\)

  2. Take a \(G{\hbox{-}}\)lift of \(f\) to \({\operatorname{SO}}_{4,4}\) to obtain \(G(f)\), and define \begin{align*} \Theta(f)(g) = \int_{[{\operatorname{SP}}_4]} \theta(g, h) \mkern 1.5mu\overline{\mkern-1.5muf(h)\mkern-1.5mu}\mkern 1.5mu \, dh ,\end{align*} then \(\Theta\) on \({\operatorname{SO}}_{4,4}\times {\operatorname{SP}}_4\) is a \(\theta\) function.

  3. There is a good theory of quaternionic modular forms on \({\operatorname{SO}}_{4, n}\), so choose \(\theta(g,h)\) such that \(\Theta(f)\) is a one of weight \(\ell\) (and cuspidal).

  4. Express the Fourier coefficients of \(\Theta(f)\) in terms of classical Fourier coefficients of \(f\), showing that the Fourier coefficients of \(G(f)\) are in \({ \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Z} \mkern-1.5mu}\mkern 1.5mu }\).

  5. Use \({\mathsf{G}}_2 \xhookrightarrow{\iota}{\operatorname{SO}}_{4,4}\) and pullback to obtain \(i^*(\Theta(f))\), which is still cuspidal and has Fourier coefficients that are sums of the original coefficients, so still in \({ \mkern 1.5mu\overline{\mkern-1.5mu \mathbb{Z} \mkern-1.5mu}\mkern 1.5mu }\).

There is an explicit dimension formula for the level 1 cuspidal quaternionic modular forms of weight \(\ell\). In particular, the smallest is a level 1 cusp form of weight 6.

Suppose \(\phi\) is a level 1 cuspidal quaternionic modular form on \({\mathsf{G}}_2\) associated to a cuspidal automorphic representation \(\pi\) on \({\mathsf{G}}_2({\mathbb{A}})\). Suppose that the Fourier coefficient \(a_\phi({\mathbb{Z}}{ {}^{ \scriptscriptstyle\times^{3} } }) \neq 0\), then

  1. The complete standard \(L{\hbox{-}}\)function of \(\pi\) has a functional equation: \begin{align*} \Lambda(\pi, \text{std}, s) = \Lambda(\pi, \text{std}, 1-s) .\end{align*}

  2. There exists a Dirichlet series for this \(L{\hbox{-}}\)function expressing the Fourier coefficients in terms of an \(L{\hbox{-}}\)function: \begin{align*} \sum_{T \subseteq {\mathbb{Z}}{ {}^{ \scriptscriptstyle\times^{3} } },n\geq 1} {a_\phi({\mathbb{Z}}+ nT) \over [{\mathbb{Z}}{ {}^{ \scriptscriptstyle\times^{3} } } : T]^{s-\ell+1} }n^{-s} = a_p({\mathbb{Z}}{ {}^{ \scriptscriptstyle\times^{3} } }) {L(\pi, \text{std}, s-z\ell+1) \over \zeta(s-2\ell + 2)^2 \zeta(2s-4\ell+2) } .\end{align*}

Carry out a refined analysis of a Rankin-Selberg integral (due to Gurevich-Segal).

13.3 A Theorem

There is a theory of half-integral weight modular forms on \({\mathsf{G}}_2\). These have a good notion of Fourier coefficients taking values in \({\mathbb{C}}/\left\{{\pm 1}\right\}\).

Suppose \(R \subseteq E\) is a cubic ring in a totally real cubic field. Let \({\partial}_R\) be the different, and let \(Q_R\) be the square roots of \({\partial}_R^{-1}\) in the narrow class group of \(E\). Say \((I, \mu)\) is balanced

Note that if \(R\) is the maximal order, \((I, \mu)\) is balanced iff \(I^2\mu = {\partial}_R^{-1}\). Define an equivalence relation by \begin{align*} (I, \mu) \sim (I', \mu') \iff \exists \beta\in E^{\times}, I' = \beta I, \mu' = \beta^{-2}\mu \end{align*} and set \(Q_R\) to be the balanced pairs mod equivalence.

There exists a weight \(1/2\) modular form \(\theta'\) on \({\mathsf{G}}_2\) whose Fourier coefficients include the numbers \(\pm {\left\lvert {Q_R} \right\rvert}\) for \(R\) even monogenic, i.e.  \begin{align*} R = {\mathbb{Z}}[y]/ \left\langle{y^3 + cy^2 + by+a, a,b,c\in 2{\mathbb{Z}}}\right\rangle .\end{align*}

Define \(\Theta\) on \(\tilde {\operatorname{F}}_4\)? Then let \(\Theta'\) be a pullback along \({\mathsf{G}}_2\to {\operatorname{F}}_3\).

14 Aaron Pollack, Talk 4: Beyond \({\mathsf{G}}_2\)

Upshot for today: there exist groups \({\mathsf{G}}_3, {\operatorname{F}}_4, {\mathbf{E}}_{n, 4}\) for \(n=6,7,8\) where \({\mathsf{G}}_2\) is split and the \({\mathbf{E}}\) groups are rank 4 over \({\mathbb{R}}\). These admit modular forms with Fourier expansions and coefficients similar to the \({\mathsf{G}}_2\) story. We’ll define these exceptional groups today.

14.1 Exceptional Algebras

Let be a composition algebra over \(k\), where \(\operatorname{ch}k = 0\), with a multiplication \(C{ {}^{ \scriptstyle\otimes_{k}^{2} } }\to C\) which is not necessarily commutative or associative. There exists a norm map \(n_C: C\to K\) given by a nondegenerate quadratic form with \(n_C(xy) = n_C(x) n_C(y)\).

Let \(J_C = H_3(C)\) be Hermitian \(3\times 3\) matrices with coefficients in \(C\), so \begin{align*} J_C = \left\{{ \begin{bmatrix} c_1 & x_3 & x_2^* \\ x_3^* & c_2 & x_1 \\ x_2 & x_1^* & c_3 \end{bmatrix} {~\mathrel{\Big\vert}~}c_i\in k, c_i\in C }\right\} .\end{align*} This has dimension \(3+3C\) over \(k\).

For \(C=k, H_3(K)\) are symmetric \(3\times 3\) matrices and there is a determinant map \begin{align*} \operatorname{det}: J_C &\to k \\ X&\mapsto c_1c_2 c_3 - \sum c_i n_C(x_i) + \operatorname{Trace}_C(x_1(x_2 x_3)) .\end{align*}

If \(C=k\) this is the usual determinant. Note that \(M_J' = \left\{{g\in \operatorname{GL}(J_C) {~\mathrel{\Big\vert}~}\operatorname{det}(gX) = \operatorname{det}(X) \forall X\in J_C}\right\}\) has positive dimension, and is thus infinite, making it an interesting algebra.

Idea: there exists a group \(G_{J_C}\) such that

All have a good notion of quaternionic modular forms and Fourier expansions/coefficients.

A degree map \(J_C \xrightarrow{\deg} k\) commuting with \(x\mapsto x^3\) and \(\operatorname{det}\) recovers \(G_{J_C = k} = {\mathsf{G}}_2\). Recall \begin{align*} {\mathfrak{g}}_2 = {\mathfrak{sl}}_{2, \ell}[0] + {\mathfrak{sl}}_{2, s}[0] + V_2 \otimes W[1] \end{align*} which is \(C_2{\hbox{-}}\)graded; we’ll mimic this to construct \({\mathfrak{g}}_{J_C}\).

14.2 Freudenthal Construction

Let \(J = J_C/{\mathbb{Q}}\) and \(k={\mathbb{Q}}\) and \begin{align*} W_J = {\mathbb{Q}}\bigoplus J \oplus J {}^{ \vee }\oplus {\mathbb{Q}} .\end{align*} There is a symplectic form \begin{align*} {\left\langle {{\left[ {a,b,c,d} \right]}},~{{\left[ {a,b,c,d} \right]}} \right\rangle} = ad' - (b,c') - (c, b') - dc' .\end{align*} There is a degree 4 polynomial map \(q: W_J\to {\mathbb{Q}}\). Define \begin{align*} H_J^1 = \left\{{g\in \operatorname{GL}(W_J) {~\mathrel{\Big\vert}~}{\left\langle {gw},~{gw'} \right\rangle} = {\left\langle {w},~{w'} \right\rangle} \,\forall w,w'\in W_J,\, q(gw) = q(w)}\right\} .\end{align*} This recovers:

\(C\) \(H_J^1\) \({\mathfrak{g}}_J\)
\({\mathbb{Q}}\) \(\operatorname{C}_3\) \({\operatorname{F}}_4\)
\(K\) \(\operatorname{A}_5\) \({\mathbf{E}}_6\)
\(B\) \({ \mathsf{D} }_6\) \({\mathbf{E}}_7\)
\(\Theta\) \(E_7\) \({\mathbf{E}}_8\)

Define \begin{align*} {\mathfrak{g}}_J ={\mathfrak{sl}}_2[0] + {\mathfrak{h}}_J^0[0] + (V_2\otimes W_J)[1] ,\end{align*} where \({\mathfrak{h}}_J^0 = \mathsf{Lie}(H_J^1)\), and define \begin{align*} G_J \mathrel{\vcenter{:}}=\mathop{\mathrm{Aut}}^0({\mathfrak{g}}_J) .\end{align*}

If \(n_C: C\otimes{\mathbb{R}}\to {\mathbb{R}}\) is positive definite, we say \({\mathsf{G}}_J\) is a quaternionic exceptional group.

If \(K_J \subseteq {\mathsf{G}}_J({\mathbb{R}})\) is a maximal compact, then \begin{align*} K_J = {{\operatorname{SU}}_2 \times L_J' \over \mu_2} \end{align*} where \(L_J^1\) is a compact form of \(H_J^1\).

There is a Cartan involution \(\theta: {\mathfrak{g}}_J\to {\mathfrak{g}}_J\), which over \({\mathbb{C}}\) yields \begin{align*} {\mathfrak{g}}_J^{\theta = \operatorname{id}} &= k_0\otimes{\mathbb{C}}\cong {\mathfrak{sl}}_2 + {\mathfrak{h}}_J^0 \\ {\mathfrak{g}}_J^{\theta = -\operatorname{id}} &= p_0\otimes{\mathbb{C}}\cong V_2 + W_J .\end{align*}

There is an action \(K_J\curvearrowright V_\ell = \operatorname{Sym}^{2\ell}({\mathbb{C}}^2)\otimes\one\)

A modular form on \({\mathsf{G}}_3\) of weight \(\ell\) is an automorphic form \begin{align*} \phi: \dcosetl{G_J({\mathbb{R}})}{G_J({\mathbb{A}})} \to V_\ell \end{align*} such that

  1. \(\phi(gk) = k^{-1}\phi(g)\) for all \(k\in K_J\)
  2. \(D_\ell \phi \equiv 0\)

Here \(D_\ell\) is defined as in the \({\mathsf{G}}_2\) case, replacing \(\operatorname{Sym}^3(V_2) = W\) with \(W_J\).

There is a Heisenberg parabolic \(P = MN \leq G_J\) with \(M = H_J\), and \(N \supseteq Z\) a two-step filtration with \(Z\) 1-dimensional and \(N/Z \cong W_J\) abelian.

Modular forms on \(G_J\) of weight \(\ell\)

  1. Have Fourier coefficients and expansions along \(N/Z\): \begin{align*} \phi_Z(g) = \phi_N(g) + \sum_{w\in W_J({\mathbb{Q}}), w\geq 0} a_\phi(w) W_w(g) ,\end{align*} where \(a_\phi(w)\in {\mathbb{C}}\) are the Fourier coefficients of \(\phi\) and \(W_w\) are completely explicit.

  2. Under appropriate embeddings \begin{align*} {\mathsf{G}}_2 \hookrightarrow{\operatorname{F}}_4 \hookrightarrow{\mathbf{E}}_{6,4}\hookrightarrow E_{7,4} \hookrightarrow E_{8,4} ,\end{align*} for a modular form \(\phi\) of weight \(\ell\) of one groups, the pullbacks \(i^* \phi\) to a smaller group are again modular forms of weight \(\ell\) whose Fourier coefficients are sums of Fourier coefficients of \(\phi\).

  1. There exists a nonzero weight 4 modular form \(\theta_{\min}\) on \({\mathbf{E}}_{8, 4}\) with rational Fourier coefficients.
  2. There exists a nonzero weight 8 modular form \(\tilde\theta_{\min}\) on \({\mathbf{E}}_{8, 4}\) with rational Fourier coefficients.

Say a modular form \(\phi\) on \({\mathsf{G}}_3\) is distinguished iff

Suppose \(K/{\mathbb{Q}}\) is quadratic imaginary, then there exists a distinguished modular form of weight 4 \(\Theta_K\) on \(G_{J_K} = {\mathbf{E}}_{6,4}\).

Set \(\Theta_K i^*(\Theta_{\min})\), pullback to \({\mathbf{E}}_{6,4}\). By arithmetic invariant theory, one shows it is distinguished.

15 Zhiwei Yun: Rigidity method for automorphic forms over function fields (Lecture 1)

Goal: construct automorphic data over function fields and working out the Langlands correspondence for such examples. Setup:

The goal of this theory is to study \({\mathcal{A}}_K\) as an object of \(\mathsf{ {\mathcal{H}}_K}{\hbox{-}}\mathsf{Mod}\).

Define \begin{align*} {\mathsf{Bun}}_G = \left\{{G{\hbox{-}}\text{bundles on } X}\right\} \cong \dcoset{G(F)}{G({\mathbb{A}})}{K^\natural}, \qquad K^\natural \cong \prod_{x\in {\left\lvert {X} \right\rvert}} G({\mathcal{O}}_x) .\end{align*}

For \(G = \operatorname{GL}_n\), passing from a vector bundle to its frame bundle yields a bijection \begin{align*} \left\{{\operatorname{GL}_n{\hbox{-}}\text{bundles}}\right\} &\rightleftharpoons\left\{{\text{Vector bundes of rank } n}\right\} \\ \operatorname{\mathcal{\mathop{\mathrm{Isom}}}}({\mathcal{O}}{ {}^{ \scriptscriptstyle\oplus^{n} } }, {\mathcal{E}})\curvearrowleft\operatorname{GL}_n &\mapsfrom {\mathcal{E}} ,\end{align*} where the Isom sheaf is regarded as principal \(G{\hbox{-}}\)bundles over \(X\). This generalizes the frame bundle construction.

Due to Weil: enrich to sets with automorphisms, i.e. groupoids. Then there is an equivalence of groupoids \begin{align*} \dcoset{\operatorname{GL}_n(F)}{\operatorname{GL}_n({\mathbb{A}})}{K^\natural} &{ { \, \xrightarrow{\sim}\, }}\mathsf{VectBun}_n(X) .\end{align*}

Let \((g_x) \in \operatorname{GL}_n({\mathbb{A}})\) and assume \(g_x = 1\) for all \(x\neq x_0\). Assign a lattice in the local field \begin{align*} \Lambda_{x_0} \mathrel{\vcenter{:}}= g_{x_0} {\mathcal{O}}_{x_0}{ {}^{ \scriptscriptstyle\oplus^{n} } } \subseteq F_{x_0}{ {}^{ \scriptscriptstyle\oplus^{n} } } ,\end{align*} which is an \({\mathcal{O}}_{x_0}\) submodule of rank \(n\). Now construct a bundle by gluing with the trivial bundle on \(X\) away from \(x_0\), so glue \(\Lambda_{x_0}\) with \({\mathcal{O}}_{X\setminus x_0}{ {}^{ \scriptscriptstyle\oplus^{n} } }\) in the following way: let \(j: X\setminus\left\{{x_0}\right\} \to X\) and form \(j_* {\mathcal{O}}_{X\setminus x_0}{ {}^{ \scriptscriptstyle\oplus^{n} } }\), which is no longer coherent and it quasicoherent, so looks like meromorphic functions but with no control on the poles. For \(U \subseteq X\) an affine open, take the functions regular away from \(x_0\) and constrain its behavior at \(x_0\) and take the sheaf associated to the following: \begin{align*} U\mapsto {{\Gamma}\qty{U\setminus\left\{{ x_0 }\right\}; {\mathcal{O}}_X{ {}^{ \scriptscriptstyle\oplus^{n} } } } }\cap\Lambda_{x_0} \subseteq F_{x_0}{ {}^{ \scriptscriptstyle\oplus^{n} } } .\end{align*}

For \(t_{x_0}\) a uniformizer, set \(g_{x_0} = \operatorname{diag}(t_{x_0}, 1, 1, \cdots, 1)\). The construction above yields the bundle \({\mathcal{O}}(-x_0) \oplus {\mathcal{O}}{ {}^{ \scriptscriptstyle\oplus^{(n-1)} } }\).

Conversely, starting a vector bundle, you can get a double coset in \(\dcoset{\operatorname{GL}_n(F)}{\operatorname{GL}_n({\mathbb{A}})}{K^\natural}\): for \(V\in \mathsf{VectBun}(X)\), there exists a \(U \subseteq X\) with \({ \left.{{V}} \right|_{{U}} } \cong {\mathcal{O}}_U{ {}^{ \scriptscriptstyle\oplus^{n} } }\). Take \(\Lambda_x = { \left.{{V}} \right|_{{\operatorname{Spec}{\mathcal{O}}_x}} } = g_x {\mathcal{O}}_x{ {}^{ \scriptscriptstyle\oplus^{n} } }\).

Check that this gives an equivalence of groupoids.

This equivalence holds for more general split \(G\). For \(G= {\mathsf{Sp}}_{2n}\), a \(G{\hbox{-}}\)bundle is the same as a pair \((V, \omega)\) where \(V\) is a vector space of rank \(2n\) and \(\omega V\otimes_{{\mathcal{O}}_x} V\to {\mathcal{O}}_x\) is symplectic.

So far, this is a pointwise story, so we’ll geometrize. It’s a fact that \({\mathsf{Bun}}_G\) is a moduli stack, and its \(k{\hbox{-}}\)points and \(R_{/ {k}}\) points are \begin{align*} {\mathsf{Bun}}_G(k) &= \left\{{G{\hbox{-}}\text{bundles on } X}\right\} \\ {\mathsf{Bun}}_G(R) &= \left\{{G{\hbox{-}}\text{bundles on } X\otimes_k R}\right\} .\end{align*} It’s a theorem that these moduli functors are representable by Artin stacks.

Take \(X = {\mathbb{P}}^1\), then \({\mathsf{Bun}}_G(k){_{\scriptstyle / \sim} }\) can be described in terms of group-theoretic data. \(G{\hbox{-}}\)bundles for \(G= \operatorname{GL}_n\) are classified by Grothendieck: \begin{align*} \mathsf{vectBun}({\mathbb{P}}^1) &\rightleftharpoons\left\{{d_1 \geq d_2 \geq \cdots \geq d_n {~\mathrel{\Big\vert}~}d_i\in {\mathbb{Z}}}\right\} \\ \bigoplus_i {\mathcal{O}}(d_i) &\mapsfrom \left\{{d_i}\right\} .\end{align*}

In general, fixing \(T\leq G\) a torus and \(W\) the Weyl group yields \begin{align*} {\mathsf{Bun}}_{G/{\mathbb{P}}^1}(k){_{\scriptstyle / \sim} }\rightleftharpoons X_*(T)/W ,\end{align*} i.e. bundles are parameterized by the cocharacter lattice, modulo the Weyl group action.

We can regard \({\mathcal{A}}_{K^\natural}\) as functions on \({\mathsf{Bun}}_G(k)\), so what is the \({\mathcal{H}}_K\) action?

Let \(G=\operatorname{GL}_n\), let \(t_x\) be a uniformizer at \(x\) and take \begin{align*} _x \mathrel{\vcenter{:}}=\chi_{S}, S = K_x \operatorname{diag}(t_x,1,1\cdots,1) K_x .\end{align*} For \(f: {\mathsf{Bun}}_G(k)\to {\mathbb{C}}\), we get the elementary upper modifier of \(f\): \begin{align*} f\ast h_x: {\mathsf{Bun}}_G(k) &\to {\mathbb{C}}\\ V &\mapsto \sum_{0\to V \hookrightarrow V' \twoheadrightarrow k_x\to 0} f(V') .\end{align*} where \(k_x\) is the skyscraper sheaf at \(x\). This is analogous to summing over elliptic curves that are \(p{\hbox{-}}\)isogenous to a given curve.

One could alternatively define a Hecke operator defined by \begin{align*} h_x = \chi_S, S\mathrel{\vcenter{:}}= K_-x \operatorname{diag}(t_x^{\lambda_1}, t_x^{\lambda_2}, \cdots, t_x^{\lambda_n}) ,\end{align*} where \(\mathbf{\lambda}\) is a collection of integers, and \begin{align*} (f\ast h_x)(V) = \sum_{\substack{ V\to V' \\ \lambda, x} } .\end{align*}

15.1 Level Structures

For interesting automorphic forms, we need to use more general things than \(K^\natural\) – many interesting examples come from parahoric subgroups of \(G(F_x)\). First we define the Iwahori as a total preimage of a Borel under a reduction:

Link to Diagram

For \(G=\operatorname{GL}_2\), one gets \begin{align*} I_x = \left\{{{ \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} } {~\mathrel{\Big\vert}~}a,b,c,d\in {\mathcal{O}}_X, c\in {\mathfrak{m}}_x}\right\} .\end{align*} Now parahorics are groups that contain the Iwahori, so there is an analogy:

\(G/k\) \(G/F_x\)
Borel Iwahori
Parabolic Parahoric

For \(G= \operatorname{GL}_n\) and \(\Lambda_x \subseteq F_x{ {}^{ \scriptscriptstyle\oplus^{n} } }\), we can consider \begin{align*} {\operatorname{Stab}}_{G(F_x)} (\Lambda_x) = \left\{{g\in \operatorname{GL}_n(F_x){~\mathrel{\Big\vert}~}g\Lambda_x = \Lambda_x}\right\} .\end{align*} One can ask for simultaneous stabilizers to get parahorics. In fact, all parahorics occur as stabilizers of chains of lattices where each stage differs by dividing by a uniformizer.

To visualize these, one needs affine Dynkin diagrams – these are generally obtained by adding a new point connected only to the long root. In \({\mathsf{G}}_2\), the diagram is:

Here taking

16 Zhiwei Yun, Lecture 2

Today: what is rigidity?

16.1 Automorphic Data

Given the following:

Link to Diagram

The pair \((K_S, \chi_S)\) is automorphic data.

Note that \(\chi_S = 1\) recovers \(f\in {\mathcal{A}}_K\) where \(K = K_S \times \prod_{x\not\in S} G({\mathcal{O}}_x)\).

A map \begin{align*} f\in C^0(\dcoset{G(F)} {G({\mathbb{A}})} {\prod_{x\not\in S} G({\mathcal{O}}_x) } \to {\mathbb{C}}) \end{align*} is \((K_S, \chi_X){\hbox{-}}\)typical iff \begin{align*} f(gk_x) = \chi_x(k_x) f(g)\qquad \forall x\in S, k_x\in K_x, g\in G({\mathbb{A}}) .\end{align*}

We want to make \(\dim {\mathcal{A}}_c(K_S,\chi_S) = 1\). In this case, the Hecke algebra \({\mathcal{H}}_{K_y} \curvearrowright f\in {\mathcal{A}}_c(K_S, \chi_S)\) by a character, making \(f\) a Hecke eigenform.

16.2 Examples of naive ridigidy

Let

Then choosing characters \(\chi_x: k^{\times}\to {\mathbb{C}}^{\times}\) generically will imply \(\dim {\mathcal{A}}_c(K_S, \chi_S) = 1\). Here generic means that \(\prod \chi_i^{\pm 1}\neq 1\). By global Langlands for \({\operatorname{SL}}_2\), any \(f\in {\mathcal{A}}_c(K_S, \chi_S)\) will yield a 2-dimensional local system on \({\mathbb{P}}^1\setminus\left\{{ 0,1,\infty }\right\}\) ramified at the 3 punctures. These will be solutions to hypergeometric differential equations.

For \(G=\operatorname{PGL}_2\) (where the example works similarly), for \(\chi_0, \chi_1 = 1\) and \(\chi_\infty\) quadratic, there is a cover

Moreover \(\left\{{H^1(E_t)}\right\}\) will be a rank 2 local system on this base.

Let

There is a map \(K_\infty \to k\) where \({ \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} } \to b+{c\over \tau}\operatorname{mod}\tau\). Any character \(k \xrightarrow{\psi} {\mathbb{C}}^{\times}\) can be extended to \(\chi_\infty: K_\infty \to k \xrightarrow{\psi} {\mathbb{C}}^{\times}\), and \(\dim {\mathcal{A}}_c(K_S, \chi_S)\). This yields a Kloosterman local system on \({\mathbb{P}}^1\setminus\left\{{ 0,\infty }\right\}\), where \begin{align*} \operatorname{Kl}(a) = \sum_{x\in k^{\times}} \psi\qty{x + a\over x} \end{align*} recovers the classical Kloosterman sum by taking trace of Frobenius.

16.3 Naive Rigidity

\begin{align*} (K_S, \chi_S) \text{ is rigid }\iff \dim {\mathcal{A}}_c(K_S, \chi_S) = 1 .\end{align*}

If \(\pi_1 G\neq 1\), then \(\pi_0 {\mathsf{Bun}}_G\geq 2\), yielding multiple components. It’s also not clear if this type of dimension bound will hold after a base change \(k\to k'\).

16.4 Base Change

For \(k'/k\) finite, write \(X'\mathrel{\vcenter{:}}= X\otimes_k k'\) for the base change. Let \(S\to S'\) be the preimage of \(S\) in \(S'\), and consider \(k'_x \mathrel{\vcenter{:}}= K_x\otimes_k k'\) How can we base change a character? We need a norm map to fill in the following diagram:

Link to Diagram

Link to Diagram

Here \(I_x(k') = \left\{{a,b,c,d\in {\mathcal{O}}_x\widehat{\otimes} k' \cong k'{\left[\left[ t \right]\right] }}\right\}\).

Link to Diagram

We now geometrize this process to send characters to character sheaves, i.e. rank one local systems on \(K_x\). We have a way of taking \((K_S, \chi_S)\) to \((K'_S, \chi_S')\) for extensions \(K\to K'\), so we can form \({\mathcal{A}}(k'; K'_S, \chi'_S)\).

Automorphic data \((K_S, \chi_S)\) is weakly rigid iff \(\dim {\mathcal{A}}_c(k'; K'_S, \chi_S')\) is uniformly bounded for all extensions \(k\to k'\).

16.5 Relevant Points

Recall that there is a bijection \begin{align*} \dcoset{G(F)}{G({\mathbb{A}})}{K} \rightleftharpoons{\mathsf{Bun}}_G(K)(k) ,\end{align*} so functions \(f\in {\mathcal{A}}_c(K_S, \chi_S)\) are functions on \({\mathsf{Bun}}_G(K_S^+)(k)\) where for \(x\in S\), \(K^+_x {~\trianglelefteq~}K_x\) with \({ \left.{{\chi_x}} \right|_{{K_x^+}} } = 1\) and \(K_x/K_x^+\) are the \(k{\hbox{-}}\)points of a finite dimensional group \(L_x\).

\(I_x^+{~\trianglelefteq~}I_x = K_x \to {\mathbb{G}}_m(k)\).

\(I_\infty^{++} = K_\infty^+ {~\trianglelefteq~}I_\infty^+ = K_\infty \to k{ {}^{ \scriptscriptstyle\oplus^{2} } }\) where \({ \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} } \mapsto \qty{b, {c\over \tau}} \operatorname{mod}\tau\).

There is a right action \begin{align*} C^0({\mathsf{Bun}}_G(K_S^+)(k) \to {\mathbb{C}}?) \curvearrowleft\prod_{x\in S} L_x(k) ,\end{align*} and the eigenfunctions with eigenvalues \((\chi_x)_{x\in S}\) are in \({\mathcal{A}}_c(K_S, \chi_S)\). As a set, \({\mathsf{Bun}}_G(K_S^+)\) has commuting left and right actions, where quotienting by the right action yields a principal homogeneous space. The left action is by \(\mathop{\mathrm{Aut}}({\mathcal{E}})\) for \({\mathcal{E}}\in {\mathsf{Bun}}_G(K_S)\), and permutes points in the fiber in \(\tilde {\mathcal{E}}\in {\mathsf{Bun}}_G(K_S^+)\). So there is an evaluation map which is well-defined up to conjugacy \begin{align*} \operatorname{ev}_{\mathcal{E}}: \mathop{\mathrm{Aut}}({\mathcal{E}}) \to \prod_{x\in S} L_x(k) .\end{align*}

A \(k{\hbox{-}}\)point \({\mathcal{E}}\in {\mathsf{Bun}}_G(K_s)(k)\) is \((K_S, \chi_S){\hbox{-}}\)relevant iff \begin{align*} { \left.{{ \operatorname{ev}_{\mathcal{E}}^*\qty{ \prod_{x\in S} \chi_x } }} \right|_{{\mathop{\mathrm{Aut}}({\mathcal{E}})^0(k) }} } = 1 .\end{align*} Similarly one can define relevant \(k'{\hbox{-}}\)points for \(k'/k\) a finite extension.

\begin{align*} \dim {\mathcal{A}}_c(k'; K_S, \chi_S) \leq {\sharp}\operatorname{Rel}(K_S), \quad \operatorname{Rel}(K_S) \mathrel{\vcenter{:}}=\left\{{(K_S', \chi_S'){\hbox{-}}\text{relevant }k' \text{ points of } {\mathsf{Bun}}_G(K_S) }\right\} .\end{align*} Note that taking connected components in the definition is needed to make this stable under base change.

\((K_S, \chi_S)\) is weakly rigid \(\iff {\sharp}\operatorname{Rel}(K_S) < \infty\).

Let

Note that \begin{align*} {\mathsf{Bun}}_G(K_S)(k) = \left\{{ V\in \mathsf{VectBundle}^{\operatorname{rank}= 2}, \iota: \bigwedge\nolimits^2 V\cong {\mathcal{O}}_X, \left\{{\ell_x \subseteq V_x}\right\}_{x\in S} }\right\} ,\end{align*} where the \(\ell\) are lines. So these are bundles with extra structure at fixed places, and are parameterized by 5-tuples \({\mathcal{E}}= (V, \iota, \ell_0, \ell_1, \ell_\infty)\). For all \(x\in S\) we have \begin{align*} I_x &\to {\mathbb{G}}_m = L_x \\ { \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} } &\mapsto a\operatorname{mod}\tau ,\end{align*} and the evaluation map is \(\operatorname{ev}: \mathop{\mathrm{Aut}}({\mathcal{E}}) \to \prod_{x\in S} {\mathbb{G}}_m\). Each \(\gamma\in \mathop{\mathrm{Aut}}({\mathcal{E}})\) is a map \(V\to V\) where \(\gamma_0 \curvearrowright V_0\) preserving \(\ell_0\). Is it the case that \begin{align*} { \left.{{\prod_{x\in S} \chi_x}} \right|_{{\mathop{\mathrm{Aut}}({\mathcal{E}})^0}} } \underset{?}{=} 1 .\end{align*} For \(V = {\mathcal{O}}^2\) and \(\ell_x \subseteq k^2\) in generic position, \(\mathop{\mathrm{Aut}}({\mathcal{E}}) = \left\{{\pm 1}\right\}\) so they are relevant. Other points are irrelevant: if \(V = L \oplus L'\) with \(\ell_x\in L_x\) or \(L_x'\), \(\mathop{\mathrm{Aut}}({\mathcal{E}})\) will contain a copy of \({\mathbb{G}}_m\) that acts by scaling each \(L\) which will map nontrivially to \(\prod_{x\in S} {\mathbb{G}}_m\). Since \(\prod \chi_i^{\pm 1}\neq 1\), we get \({ \left.{{\operatorname{ev}_{\mathcal{E}}^* \prod \chi_x}} \right|_{{{\mathbb{G}}_m}} } \neq 1\).

17 Zhiwei Yun, Lecture 3

The Langlands correspondence:

Automorphic Galois
\(G\) \(G {}^{ \vee }\)
Eigenforms \(f\in {\mathcal{A}}_c(K_S, \chi_S)\) Local systems, \(\pi_1(X\setminus S) \to G {}^{ \vee }(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu)\)
Rigid automorphic data Rigid local systems

Today we’ll discuss going from the automorphic side to the Galois side by designing rigid automorphic data.

17.1 Numerical Rigidity

Automorphic functions will be functions on the algebraic stack \({\mathsf{Bun}}_G(K_S)\), so we want to consider its \(k{\hbox{-}}\)points. This stack should have (possibly negative) dimension at most zero, what does this tell us about the level groups? For a curve \(C\), there is a formula: \begin{align*} \dim {\mathsf{Bun}}_G(K_S) = 0 \iff \sum_{x\in S} [G({\mathcal{O}}_x): K_x] = (1-g(C))\dim G ,\end{align*} where the brackets indicate relative dimension, which is always non-negative. Recall that \(I_x\) is a preimage of a Borel under reduction, and for \(K_x = I_x\) we have \begin{align*} [G({\mathcal{O}}_x): I_x] = \dim G({\mathcal{O}}_x)/I_x = \dim G/B = {\sharp}\Phi^+ .\end{align*} If \(K_x\) is not contained in \(G({\mathcal{O}}_x)\), then \begin{align*} [G({\mathcal{O}}_x): K_x] = \dim G({\mathcal{O}}_x)/G({\mathcal{O}}_x) \cap K_x - \dim K_x/G({\mathcal{O}}_x) \cap K_x .\end{align*}

The RHS in the formula is non-negative only when \(g=0, 1\), so we expect most rigid data to come from \({\mathbb{P}}^1\). Genus 1 is a very special case, we get \(K_x\sim G({\mathcal{O}}_x)\).

Consider

Then \begin{align*} \dim G = \sum_{x=0,1,\infty} [G({\mathcal{O}}_x): K_x] .\end{align*} If \(K_x\) corresponds to a subdiagram of a Dynkin diagram, we can read off the reductive quotient \(L_x\) to get a surjective quotient map \(K_x\twoheadrightarrow L_x\). In this case, \begin{align*} [G({\mathcal{O}}_x): K_x] = {1\over 2}\qty{\dim G - \dim L_x} .\end{align*} The condition then becomes \begin{align*} \dim G = \sum_{x=0,1,\infty}\dim L_x .\end{align*}

Consider the same setup for \(G={\mathsf{G}}_2\).

For \(G = {\mathbf{E}}_8\), take \(L_1\) to be the Iwahori and:

Idea: delete a node to try to get a group of roughly half-dimension. Cook up an order 2 character on the reductive quotient for \(x=0\):

Link to Diagram

Setting \(\chi_\infty = \chi_1 = 1\), this yields automorphic datum which turns out to be rigid.

17.2 Matching with local monodromy

Given a local system, restrict to a formal neighborhood of a puncture to get a representation of the local Galois group, which we can restrict to inertia:

Link to Diagram

If \(K_x = I_x\) and one forms the character \(K_x\to T(k)\to \mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu^{\times}\), the representation \(\rho_x\) is tamely ramified. There is a canonical map to the residue field \({\mathrm{In}}_x \to k_x^{\times}\), and \(\chi\) can be turned into a morphism \(k_x^{\times}\to T {}^{ \vee }(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu^{\times})\) to the dual torus. The composite dual character \({\mathrm{In}}_x \to k_x^{\times}\to T {}^{ \vee }(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu^{\times})\) has finite order and yields the semisimplification \(({ \left.{{\rho_x}} \right|_{{{\mathrm{In}}_x}} })^{{\mathrm{ss}}}\). This yields unipotent monodromy, usually “maximally” nontrivial.

For \(K_x = I_x^+ = \left\{{{ \begin{bmatrix} {a} & {b} \\ {c} & {d} \end{bmatrix} } {~\mathrel{\Big\vert}~}a-1,d-1,c\equiv 0 \operatorname{mod}t}\right\}\), one gets a character:

Link to Diagram

One gets a wildly ramified representation: \begin{align*} \rho_x: { \mathsf{Gal}} (\mkern 1.5mu\overline{\mkern-1.5muF\mkern-1.5mu}\mkern 1.5mu_x/F_x)\to \operatorname{GL}_2(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu) .\end{align*} The Swan conductor is \({\mathrm{Sw}}(\rho_x) = 1 = {1\over 2} + {1\over 2}\), where the first factor comes from the shape of the level group \(I_x^+\). There is a Moy-Prasad filtration on \(I_x\) indexed by \({1\over h}{\mathbb{Z}}\) for \(h\) the Coxeter number of \(G\), which for \(G={\operatorname{SL}}_2\) yields \(h=2\). The filtration is

Link to Diagram

If \(K_x \subseteq P_x(r)\) for \(r\in {\mathbb{Q}}\), then all slopes of \(\rho_x\) are at most \(r\), bounding the ramification. In this case depth matches up with slopes.

Ansatz for finding correct automorphic data: if \((K_S, \chi_S)\) is rigid and \(\rho: \pi_1\to G {}^{ \vee }\), there should be an equality involving \(a\) the Artin conductor: \begin{align*} [G({\mathcal{O}}_x): K_x] = {1\over 2} a({ \operatorname{Ad} }\rho_x) .\end{align*}

Let

Compose to get a character \(P_\infty^+ \xrightarrow{?} k \xrightarrow{\psi} {\mathbb{C}}^{\times}\), where the missing morphism is the interesting bit. For \(G = {\mathsf{Sp}}_{2n} = {\mathsf{Sp}}(V)\), a Siegel parabolic is the stabilizer of a Lagrangian subspace in \(V\) and has the following shape:

Write this as \(P_{\mathrm{Sieg}}\) preserving a Lagrangian \(L\) and take its transpose to get \(P_{\mathrm{Sieg}}^{\operatorname{op}}\) which preserves a complementary Lagrangian subspace \(L^c \cong L {}^{ \vee }\). Let \(P_0 \subseteq G({\mathcal{O}}_0), P_\infty \subseteq G({\mathcal{O}}_\infty)\) be the associated parahorics. Define a map \begin{align*} P_\infty^+ = \left\{{{ \begin{bmatrix} {A} & {B} \\ {C} & {D} \end{bmatrix} } \in G({\mathcal{O}}_\infty) {~\mathrel{\Big\vert}~}A-I, D-I, C\equiv 0 \operatorname{mod}\tau}\right\} &\to W = \operatorname{Sym}^2(L) \oplus \operatorname{Sym}^2(L {}^{ \vee }) \\ { \begin{bmatrix} {A} & {B} \\ {C} & {D} \end{bmatrix} } &\mapsto \qty{B\operatorname{mod}\tau, {C\over \tau} \operatorname{mod}\tau} .\end{align*}

We then have

We can further apply the trace pairing, fixing \(S\in \operatorname{Sym}^2(L)\) and \(T\in \operatorname{Sym}^2(L {}^{ \vee })\): \begin{align*} \operatorname{Trace}: W\to k \\ (X, Y) &\mapsto \operatorname{Trace}(XT) + \operatorname{Trace}(YS) .\end{align*} Choosing a pair \((S, T)\) yields a character: \begin{align*} P_\infty^+ \to W \xrightarrow{(S, T)} k \xrightarrow{\psi} {\mathbb{C}}^{\times} .\end{align*} Stable pairs \((S, T)\) will yield rigid data, where stable is the open condition that \(ST\in \mathop{\mathrm{End}}(L)\) has distinct nonzero eigenvalues in \(\mkern 1.5mu\overline{\mkern-1.5muk\mkern-1.5mu}\mkern 1.5mu\), so regular semisimple and invertible.

Epipelagic reps of \(G(F_\infty)\) due to Reeder-J-K. Yu: for \({\mathsf{Sp}}_{2n}\) this amounts to choosing a matrix of the following shape with equally sized blocks:

18 Zhiwei Yun, Lecture 4

18.1 Kloosterman Automorphic Systems

We’ve just been on the automorphic side: today we harvest on the Galois side! For \((K_S, \chi_S)\) automorphic data, there is a Hecke action \begin{align*} f\in {\mathcal{A}}_c(K_S,\chi_S) \curvearrowleft{\mathcal{H}}_{K_x},\quad x\in {\left\lvert {X} \right\rvert} \setminus S .\end{align*} The Satake isomorphism yields a correspondence \begin{align*} \left\{{\text{Functions } {\mathcal{H}}_{K_x} \to \mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu}\right\} \rightleftharpoons \left\{{\text{Semisimple conjugacy classes in } \widehat{G}(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu)}\right\} .\end{align*} So for all places \(x\not\in S\), one gets a Satake parameter \(\sigma_x\in \widehat{G}_{{\mathrm{ss}}}(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu)\). By Langlands, there exists a representation \(\rho: \pi_1(X\setminus S) \to \widehat{G}(\mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu)\) such that \(\rho(\operatorname{Frob}_x)^{{\mathrm{ss}}} \sim \sigma_x\). How do you construct \(\rho\) from \(f\)?

We’ll geometrize this along the lines of Drinfeld, Laumon, etc. Set \(G=\operatorname{GL}_n\) and let \(T_x\in {\mathcal{H}}_{K_x}\) be the characteristic function on \(K_x \operatorname{diag}(t_x, 1,\cdots, 1) K_x\). For \(f: {\mathsf{Bun}}_G(k)\to \mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu\), define an operator \begin{align*} (T_x f)({\mathcal{E}}) \mathrel{\vcenter{:}}=\sum_{{\mathcal{E}}' \hookrightarrow{\mathcal{E}}\text{ length 1 at } x} f({\mathcal{E}}') .\end{align*} Note that the index set is isomorphic to \({\mathbb{P}}({\mathcal{E}}_x)\). This translates functions to sheaves: summing corresponds to taking cohomology, characters become character sheaves. Let \({\mathsf{Hk} }^1 = \left\{{{\mathcal{E}}' \to {\mathcal{E}}\text{ of length 1}}\right\}\), then there is a span:

Link to Diagram

The operator \(T_x\) geometrizes in the following way: \begin{align*} T_1 {\mathcal{F}}\mathrel{\vcenter{:}}=(h_2){}_{!}(h_1{}^{*}{\mathcal{F}}) \in {\mathsf{Sh}}({\mathsf{Bun}}_G\times (X\setminus S)) ,\end{align*} and being an eigensheaf translates \(T_x f = \lambda_x f\) for \(\lambda_x\in \mkern 1.5mu\overline{\mkern-1.5mu{ {\mathbb{Q}}_\ell }\mkern-1.5mu}\mkern 1.5mu\) to the condition \begin{align*} T_1{\mathcal{F}}= {\mathcal{F}}\boxtimes E .\end{align*} The goal is to compute \(E\); this will yield \begin{align*} \operatorname{Trace}(\operatorname{Frob}_x; E) = \lambda_x \qquad\forall x\in X\setminus S .\end{align*}

Benedict Gross constructed \(\operatorname{Kl}\) automorphic datum, showed rigidity using a trace formula, and conjectured some properties of \(\rho\) related to a Kloosterman local system. Heinloth-Ngo-Y. constructed this \(\rho\), uncovering the story of rigidity here.

Let

In this case, \begin{align*} {\mathsf{Bun}}_G(K_0, K_\infty) = \left\{{ \begin{array}{l} V\in { {\mathsf{Bun}}\qty{\operatorname{GL}_r} }^{\operatorname{rank}= n}({\mathbb{P}}^1), \\ F^n \supseteq F^{n-1} \supseteq\cdots\supseteq F^1 \text{ a full flag on } V_0, \\ F_0 \subseteq F_1 \subseteq \cdots\subseteq F^n \text{a full flag on }V_\infty, \\ \left\{{e_i}\right\} \text{a basis of }{\mathsf{gr}\,}_i (F_\bullet) \end{array} }\right\}/{\operatorname{Pic}} .\end{align*} There is a unique relevant point on each component of \({\mathsf{Bun}}_G(K_0, K_\infty)\), where \(\deg V \operatorname{mod}n\) is well-defined. It’s given by \({\mathcal{E}}_0\) where \({\mathcal{O}}{ {}^{ \scriptscriptstyle\oplus^{n} } } = \bigoplus_{i\leq n} {\mathcal{O}}_i\), with a flag \(\left\{{e_n}\right\}, \left\{{e_n, e_{n-1}}\right\},\cdots\). One can show that \(\mathop{\mathrm{Aut}}({\mathcal{E}}_0) = 1\) making it automatically relevant.

A point \({\mathcal{E}}_1\) yields \(\bigoplus_{k\leq n-1} {\mathcal{O}}e_k \oplus {\mathcal{O}}(1)e_n\), with flags

There is a Hecke stack \({\mathsf{Hk} }\) containing \(\left\{{\phi: {\mathcal{E}}_0 \hookrightarrow{\mathcal{E}}_1}\right\}\), and a span:

Link to Diagram

Pull-push yields a local system. Similarly:

Link to Diagram

Defining \(E \mathrel{\vcenter{:}}={\mathbb{R}}^{n-1} \pi{}_{!}\sigma{}^{*}\mathrm{AS}_\psi \in \mathsf{LocSys}^{\operatorname{rank}= n}\) exactly recovers Deligne’s Kloosterman sheaf.

For more general \(G\), \({\mathcal{H}}_{K_x}\) has a Kazhdan basis \(C_\lambda\), where dominant weights \(\lambda \in X_*(T)\) correspond to irreducible reps of \(G {}^{ \vee }\). Taking \(T_x\) for \(\operatorname{GL}_n\) recovers the standard representation of \(G {}^{ \vee }= \operatorname{GL}_n\). The geometric incarnation of the Hecke operator is \(T_\lambda F\):

Link to Diagram

Here \({\mathcal{F}}\) is an eigensheaf, so \(T_\lambda {\mathcal{F}}= {\mathcal{F}}\boxtimes E_\lambda\). Note that the \(\operatorname{\mathcal{IC}}\) sheaf is not always constant.

\(\lambda \mapsto E_\lambda\) comes from a \(\widehat{G}{\hbox{-}}\)local system on \(X\setminus S\):

Link to Diagram

18.2 Applications

If \((K_S, \chi_S)\) is “tame,” where \(K_x\) is parahoric, this data will make sense over any base field \(k\) which \(\chi_S\) is replaced by a character sheaf. Note that this only works for multiplicative characters, since additive characters depend on characteristic. One can construct these Hecke eigensheaves and \(G {}^{ \vee }\) local systems for arbitrary fields, e.g. for \({\mathbb{P}}^1_{/ {{\mathbb{Q}}}} \setminus S\) where there may not even be a theory of automorphic forms. A first example constructs an \({\mathbf{E}}_8{\hbox{-}}\)local system on \({\mathbb{P}}^1_{/ {{\mathbb{Q}}}} \setminus\left\{{ 0,1,\infty }\right\}\), yielding a motive whose motivic Galois group is \({\mathbf{E}}_8\). One can then apply this to the inverse Galois problem, arguing that there exists a number field \(K\) such that \begin{align*} { \mathsf{Gal}} (K/{\mathbb{Q}}) \cong {\mathbf{E}}_8({\mathbb{F}}_\ell), \qquad \ell \gg 0 .\end{align*}

See the notes for relations to “rigidity methods” in inverse Galois theory.

18.3 Open Problems

Classification: say \(G=\operatorname{GL}_n\), can one classify all rigid automorphic data?

These should correspond under Langlands to rigid local systems, where there is an algorithmic classification due to Katz in the tame case and Arinkin in general. One can start with rank 1 local systems and apply one of three simple procedures to get local systems of higher rank. Note that hypergeometric local systems occur.

Is there an algorithmic way of producing automorphic data?

Is there a uniform way to check rigidity?

Checking rigidity requires knowing the specific geometry of \({\mathsf{Bun}}_G\) and some tricky linear algebra. There are some results that provide the uniform bound on dimensions \({\mathcal{A}}_c\) needed to prove weak rigidity.

Can \({\mathcal{A}}_c(K_S, \chi_S)\) be further decomposed into Hecke modules when the dimension is bigger than 1?

This dimension can grow exponentially.

19 Saturday, March 05

Goal: proving rigidity. Start with example 2.1.5 in the notes on Kloosterman automorphic data.

20 Bibliography


  1. This is already interesting in the case of \(X = {\mathbb{P}}^1\).↩︎