1 Introduction
For V a finite dimensional
${\mathbb C}$
vector space,
$G \leq \mathrm {GL}(V)$
a Zariski closed subgroup whose identity component
$G^\circ $
is semisimple, and
$(a,b)$
a pair of non-negative integers, the
$(a,b)$
-moment of G acting on V, denoted

is defined to be the dimension of the space
$(V^{\otimes a}\otimes (V^\ast )^{\otimes b})^G$
of G-invariants.
By Larsen’s Alternative [Reference KatzKa3, 1.1.6], one knows that if
$M_{2,2}(G,V)=2$
, then either G is finite or
$G^\circ =\mathrm {SL}(V)$
. If V is endowed with an orthogonal autoduality and
$G \leq O(V)$
, and if
$M_{2,2}=3$
, then either G is finite or
$G^\circ =\mathrm {SO}(V)$
.Footnote
1
If V is endowed with a symplectic autoduality,
$\dim (V) \ge 4$
, and
$G \leq \mathrm {Sp}(V)$
, then
$M_{2,2}=3$
implies that either G is finite or
$G=\mathrm {Sp}(V)$
.
The cases of Larsen’s Alternative in which G is finite and
$\dim (V) \ge 5$
are completely determined in [Reference Guralnick and TiepGT2, Theorem 1.5]. Two natural questions then occur. Which of these finite groups can be obtained as the geometric monodromy group of a hypergeometric sheaf on
${\mathbb G}_m$
in characteristic
$p> 0$
? Which of these finite groups can be obtained as the geometric monodomy group of a family of one-variable exponential sums?
The kinds of families of one-variable exponential sums in a given characteristic
$p>0$
we have in mind are these. We fix a prime
$\ell \neq p$
and a nontrivial additive character
$\psi : {\mathbb F}_p \rightarrow \mu _p(\overline {{\mathbb Q}_\ell })$
. [In down to earth terms, we embed
${\mathbb Q}(\zeta _p)$
into
$\overline {{\mathbb Q}_\ell }$
, which amounts to choosing a place of
${\mathbb Q}(\zeta _p)$
over
$\ell $
. The expressions we will write down will lie in
${\mathbb Q}(\zeta _p)$
, but we need to view them as lying in
$\overline {{\mathbb Q}_\ell }$
in order to apply
$\ell $
-adic cohomology.]
We are given a finite extension
$k/{\mathbb F}_p$
, a polynomial
$f(x)\in k[x]$
, say

of degree
$d \ge 1$
which is Artin-Schreier reduced (meaning that
$A_i=0$
whenever
$p|i$
). Let

be prime-to-p integers. Suppose that either
$\deg (f)> b$
or that
$\deg (f)<b$
. In the case when
$\deg (f)<b$
, we require that f is not a constant multiple of
$x^a$
. Another way of expressing this last condition is that the polynomial
$sx^a +tx^b +f(x)$
contains monomials of least
$3$
different degrees, a condition which is automatic if
$\deg (f)>b$
.
Let
$\chi $
be a multiplicative character of
$k^\times $
. When
$\deg (f)> b$
, consider the local system

on
${\mathbb A}^2/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$s,t \in L$
,

with the usual convention that
$\chi _L(0)=0$
for
, but
$\chi _L(0)=1$
for
. When
$\deg (f)<b$
, we consider the same local system, but on
${\mathbb A}^1 \times {\mathbb G}_m$
, since we need t to be invertible in this
$\deg (f)<b$
case. These families are pure of weight zero, and lisse of rank
$\max (\deg (f),b)-1$
when
and of rank
$\max (\deg (f),b)$
when
. They are geometrically irreducible precisely when

which we will assume in what follows.
As we will see in Theorems 2.3 and 2.6, the
$M_{2,2}$
for the
$G_{\mathrm {geom}}$
of this local system is given by the answer to what should be an easy question about intersections of Fermat surfaces in
${\mathbb P}^3$
, with homogeneous coordinates
$x,y,z,w$
, or equivalently about intersections of their affine cones in
${\mathbb A}^4$
. Although this question is motivated by its applications to the determination of monodromy groups, it turns out that its answer comes from the a priori determination of certain monodromy groups.
Let us first elaborate the Fermat surfaces intersection question, and then make explicit the local systems to whose monodromy this question is related. For an integer
$n \ge 1$
, denote by
$\Sigma _{n,\mathrm {proj}} \subset {\mathbb P}^3$
the locus

and denote by
$\Sigma _{n} :=\Sigma _{n,\mathrm {aff}} \subset {\mathbb A}^4$
the locus

In what follows, when no confusion is possible, we also denote the polynomial
$x^n+y^n-z^n-w^n$
by
$\Sigma _n$
. The intersection
$\Sigma _{a} \cap \Sigma _{b}$
has dimension two. [Equivalently,
$\Sigma _{a,\mathrm {proj}} \cap \Sigma _{b,\mathrm {proj}}$
has dimension one. Here is one argument. Because each of
$a,b$
is prime to p, each of
$\Sigma _{a,\mathrm {proj}}$
and
$\Sigma _{b,\mathrm {proj}}$
is a smooth, geometrically connected surface. The intersection, viewed as lying in
$\Sigma _{b,\mathrm {proj}}$
, is either one-dimensional or it is all of
$\Sigma _{b,\mathrm {proj}}$
. The second case could only occur if the polynomial
$\Sigma _a$
is divisible by
$\Sigma _b$
, which cannot happen, because
$a < b$
. A second argument is this. If the intersection had dimension
$2$
, it would be equal to both
$\Sigma _{a,\mathrm {proj}}$
and to
$\Sigma _{b,\mathrm {proj}}$
, and we would get the conclusion that
$\Sigma _{a,\mathrm {proj}}=\Sigma _{b,\mathrm {proj}}$
, impossible because their
${\mathbb Q}_\ell $
Euler characteristics differ.]
Denote by
$\Sigma (f,a,b)$
the intersection in
${\mathbb A}^4/\overline {{\mathbb F}_p}$
of the following affine Fermat threefolds:

We define
$\Sigma _{\mathrm {proj}}(f,a,b)$
to be the intersection in
${\mathbb P}^3$
of the projective Fermat surfaces
$\Sigma _{a,\mathrm {proj}},\Sigma _{b,\mathrm {proj}}$
and every
$\Sigma _{i,\mathrm {proj}}$
with
$A_i \neq 0$
. Then
$M_{2,2}$
is the number of reduced irreducible components of dimension one of
$\Sigma _{\mathrm {proj}}(f,a,b)$
or, equivalently, the number of reduced irreducible components of dimension two of
$\Sigma (f,a,b)$
.
The loci
$\Sigma (f,a,b)$
and
$\Sigma _{\mathrm {proj}}(f,a,b)$
depend only on the set S of degrees of the Fermat surfaces being intersected. Given a set S of prime-to-p positive integers with
$\#S \ge 2$
, let us denote

Recall that by assumption, the set S of degrees occurring in
$\Sigma _{\mathrm {proj}}(f,a,b)$
satisfies
$\gcd (S)=1$
: this is equivalent to the geometric irreducibility of the family.
Every Fermat surface
$\Sigma _{a,\mathrm {proj}}$
contains the two lines
$(x=z,y=w)$
and
$(x=w,y=z)$
. If a is odd,
$\Sigma _{a,\mathrm {proj}}$
contains the third line
$(x=-y,z=-w)$
. One knows that in any odd characteristic, the intersection
$\Sigma _{1,\mathrm {proj}} \cap \Sigma _{2,\mathrm {proj}}$
consists precisely of the two lines
$(x=z,y=w)$
and
$(x=w,y=z)$
; cf. [Reference KatzKa3, p. 117]. And one knows that in any characteristic
$p \neq 3$
, the intersection
$\Sigma _{1,\mathrm {proj}}\cap \Sigma _{3,\mathrm {proj}}$
consists precisely of the three lines
$(x=z,y=w)$
,
$(x=w,y=z)$
, and
$(x=-y,z=-w)$
; cf. [Reference KatzKa4, 3.11.3].
Thus, the question breaks into two natural parts: First, for which sets S with
$\gcd (S)=1$
consisting only of odd degrees will
$\Sigma _{\mathrm {proj}}(S)$
have precisely three reduced irreducible components of dimension one (which would necessarily be the three known lines). There may also be zero-dimensional reduced irreducible components (i.e., finitely many closed points) outside these lines; these do not affect
$M_{2,2}$
. Second, for which sets S with
$\gcd (S)=1$
of degrees, at least one of which is even, will
$\Sigma _{\mathrm {proj}}(S)$
have precisely two reduced irreducible components of dimension one (which would necessarily be the two known lines). Again, having finitely many points outside the two known lines does not affect
$M_{2,2}$
.
Our original idea was to attack directly this algebro-geometric question about intersections of Fermat surfaces, and then use its solution to help determine monodromy groups. But in fact, we turn this idea on its head: after a good deal of work, we determine the monodromy groups and then use their determination to give a complete solution to the algebro-geometric question; cf. Theorem 9.2. As one can see from the solution, the ‘generic’ value of
$M_{2,2}$
– equivalently, of the number of one-dimensional irreducible components of the corresponding intersection of Fermat surfaces – is
$2$
or
$3$
. Nonetheless, there are still infinitely many cases with a different value of
$M_{2,2}$
: each of these cases comes out of specific representations of specific finite groups of Lie type and would have looked mysterious (perhaps even hidden) had one tried to approach the problem only by algebro-geometric means.
Let us explain how this works. Given a set S of prime to p integers with
$\#S =r+1 \ge 3$
and
$\gcd (S)=1$
, enumerate the elements of S, say

and consider the corresponding universal family of monic one-variable polynomials whose allowed degrees are precisely S:

We obtain a local system
${\mathcal F}(S)$
on
${\mathbb A}^r/{\mathbb F}_p$
whose trace function is given as follows. For
$k/{\mathbb F}_p$
a finite extension, and
$(t_1,\ldots ,t_r) \in {\mathbb A}^r(k)$
,

Given a multiplicative character
$\chi $
of some finite extension
$k_0/{\mathbb F}_p$
, we also have the local system
${\mathcal F}(S,\chi )$
on
${\mathbb A}^r/k_0$
whose trace function is given as follows. For
$k/k_0$
a finite extension, and
$(t_1,\ldots ,t_r) \in {\mathbb A}^r(k)$
,

In this notation, the above
${\mathcal F}(S)$
is just
The local system
${\mathcal F}(S,\chi )$
is geometrically irreducible, lisse of rank

and pure of weight
$0$
. Its geometric monodromy group
$G_{\mathrm {geom},{\mathcal F}(S,\chi )}$
is a Zariski closed subgroup of
$\mathrm {GL}_{D}/\overline {{\mathbb Q}_\ell }$
whose identity component is semisimple. On the one hand,
$M_{2,2}$
of the local system
${\mathcal F}(S,\chi )$
is the number of reduced irreducible
$2$
-dimensional components of
$\Sigma (S)$
over
$\overline {{\mathbb F}_p}$
, on which
${\mathcal L}_{\chi (xy)}\otimes {\mathcal L}_{\overline {\chi }(zw)}$
is geometrically trivial; cf. Theorem 2.4. On the other hand,
$M_{2,2}$
is the
$M_{2,2}$
for the given D dimensional representation
$V:={\mathcal F}(S,\chi )_{\overline {\eta }}$
of
$G:=G_{\mathrm {geom},{\mathcal F}(S,\chi )}$
.
The key point is that we can explicitly determine the group
$G_{\mathrm {geom},{\mathcal F}(S,\chi )}$
. This task, in the case the group is finite, was done in [Reference Katz and TiepKT6, Theorem 11.2.3]. One of the main results of this paper, Theorem 7.8, completes the task in the infinite case. In turn, this allows us to determine
$M_{2,2}$
for
$G_{\mathrm {geom},{\mathcal F}(s,\chi )}$
, and thus solve the aforementioned algebro-geometric question about intersections of Fermat hypersurfaces, in Theorem 9.2.
Once we have these results in hand, a new question arises. Suppose given an S as in (1.0.2),
$A>B_1>\ldots >B_r $
, with
$r \ge 3$
. Pick two indices in
$\{B_1,\ldots ,B_r\}$
, say
$a:=B_i < b:=B_j <A$
, and denote by

with
${\mathcal C}$
enumerated as

Suppose further given a finite extension
$k/k_0$
and elements
$c_i \in k^\times $
for
$i=1, \ldots , r-2$
. Consider the local system on
${\mathbb A}^2/k$
obtained from
${\mathcal F}(S,\chi )$
by specializing the coefficient of
$x^{C_i}$
to
$ c_i$
. Call it

This is the local system on
${\mathbb A}^2/k$
whose trace function is given as follows. For
$K/k$
a finite extension, and
$(s,t)\in {\mathbb A}^2(K)$
,

By Theorem 2.3 and Corollary 2.5, each such system
${\mathcal F}(f,B_i,B_j,\chi )$
has the same
$M_{2,2}$
as the system
${\mathcal F}(S,\chi )$
. Because
${\mathcal F}(f,B_i,B_j,\chi )$
is a pullback of
${\mathcal F}(S,\chi )$
, we have the a priori inclusion

In the case when
$G_{\mathrm {geom}, {\mathcal F}(S,\chi )}$
is a (known!) finite group, we wish to classify those of its subgroups which in the given D-dimensional representation have the same
$M_{2,2}$
. We succeed entirely when the known finite group is (the image of) one of
$\mathrm {Sp}_{2n}(q)$
,
$n \ge 1$
, or
$\mathrm {SU}_n(q)$
,
$n\ge 3$
, in a Weil representation, by showing that, with very few exceptions, the only subgroups with the same
$M_{2,2}$
are the whole group itself; see Theorems 8.2 and 8.4. This gives Theorem 11.9. A striking aspect of part (ii) of Theorem 11.9 is that it applies to the relevant
${\mathcal F}(f,a,b,\chi )$
for any f all of whose coefficients are nonzero and any
$(a,b)$
.
We also consider one-parameter specializations of such
${\mathcal F}(S,\chi )$
– that is, systems
${\mathcal F}(f,a,\chi )$
with trace function as follows: for
$K/k$
a finite extension and
$t\in K$
,

In Theorem 11.9, we prove that for given a, the local system
${\mathcal F}(f,a,\chi )$
will have the same
$G_{\mathrm {geom}}$
as
${\mathcal F}(S,\chi )$
for f in a dense open set of the affine space of allowed f’s.
In the cases when , and
is an extraspecial normalizer, we do not classify subgroups with the same
$M_{2,2}$
. Nonetheless, we prove that for given
$(a,b)$
, the local system
will have the same
$G_{\mathrm {geom}}$
as
for f in a dense open set of the affine space of allowed f’s; see Theorems 11.7 and 11.8. Again, in this case, we have the same ‘dense open set’ result for one-parameter specializations
, with the added wrinkle that the case
$a=1$
behaves quite differently in the extraspecial normalizer case. In each of the Theorems 11.7, 11.8 and 11.9, there are unknown dense open sets. It would be of some interest to determine them explicitly.
The main results of this paper include Theorems 7.8, 9.2, 10.1, 11.7, 11.8, 11.9.
2 Moments and point counting
We begin this section with the basic fact about approximating moments by large L limits.
Theorem 2.1. Let k be a finite field of characteristic p,
$\ell $
a prime
$\ell \neq p$
,
$X/k$
a smooth, geometrically connected scheme of dimension
$d \ge 1$
, and
${\mathcal F}$
a lisse
$\overline {{\mathbb Q}_\ell }$
sheaf on X which is
$\iota $
-pure of weight zero for a chosen field embedding
$\iota :\overline {{\mathbb Q}_\ell }\hookrightarrow {\mathbb C}$
. For integers
$a,b \ge 0$
, the moment
$M_{a,b}$
of
$G_{\mathrm {geom},{\mathcal F}}$
is

Proof. In terms of the auxiliary sheaf

which is
$\iota $
-pure of weight zero, and hence geometrically semisimple (cf. [Reference DeligneDe2, 3.4.1(iii)]), we have

Our asserted formula for this dimension is

By the Lefschetz trace formula, this is

By Deligne’s fundamental estimate [Reference DeligneDe2, 3.4],
$H^i_c$
is
$\iota $
-mixed of weight
$\le i$
, while
$H^{2d}_c$
is
$\iota $
-pure of weight
$2d$
. But
$\#X(L)=(\# L)^d + O((\# L)^{d-1/2})$
, and hence, the
$H^i_c$
summands with
$i < 2d$
contribute
$0$
to the
$\limsup $
. So we must prove that
$ \dim H^{2d}_c(X_{\overline {k}},{\mathcal G})$
is

If this
$H^{2d}_c$
vanishes, we are done.
If
$H^{2d}_c$
is nonzero, the eigenvalues of
$\mathrm {Frob}_k$
on this
$H^{2d}_c$
are each of the form
$(\#k)^d\alpha _i$
, for
$i=1, \ldots , \dim H^{2d}_c$
, and each of these
$\alpha _i$
has complex absolute value
$|\alpha _i|=1$
. Thus, for
$L/k$
a finite extension, we have

For any
$L/k$
, this last expression visibly has absolute value at most

As
$L/k$
grows, the tuple
$(\alpha _1^{\deg (L/k)},\ldots , \alpha _{\dim H^{2d}_c}^{\deg (L/k)})$
will, infinitely often, come arbitrarily close to
$(1,\ldots ,1)$
, while the ratio
$\#X(L)/(\# L)^d$
has limit
$1$
as L grows.
We next give a lemma on counting geometrically irreducible components.
Lemma 2.2. Let k be a finite field, and
$X/k$
a separated k-scheme of finite type, of dimension
$d \ge 0$
. Then

is the number of geometrically irreducible components of
$X_{\overline {k}}$
of dimension d.
Proof. Each geometrically irreducible component of
$X_{\overline {k}}$
is defined over some finite extension of k, so at the expense of replacing k by a finite extension of itself, we reduce to the case where each geometrically irreducible component Z is defined over k (i.e., is a geometrically irreducible k-scheme of dimension
$e_Z \le d$
). The result then follows from the Lang-Weil estimate, that for each such component Z,
$\#Z(L) =(\# L)^{e_Z} + O( (\# L)^{e_Z -1/2}).$
Theorem 2.3. Let k be a finite field of characteristic
$p>0$
, and
$f(x) \in k[x]$
, say
$f(x)=\sum _iA_ix^i$
, of degree
$d \ge 3$
which is Artin-Schreier reduced. Let
$1 \le a <b < \deg (f)$
be prime to p integers,
$\chi $
a multiplicative character of
$k^\times $
, and consider the local system
${\mathcal F}_\chi $
on
${\mathbb A}^2/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$s,t \in L$
,

with the convention that , but
$\chi _L(0)=0$
for
$\chi $
nontrivial. Consider the set
${\mathcal E}$
of exponents which occur in f:

and the affine locus
$\Sigma (S)$
as defined in (1.0.1) with
$S:=\{a,b\} \cup {\mathcal E}$
. Then

Moreover, if and all integers in S are odd, then

More precisely, is the number of geometrically irreducible components of dimension one in
$\Sigma _{\mathrm {proj}}(S)$
, while
$M_{2,2}({\mathcal F}_\chi )$
is the number of those components on which
$\chi (xy)\bar \chi (zw)$
is geometrically trivial.
Proof. Recall that
$\Sigma _d$
denotes the Fermat form
$x^d+y^d-z^d-w^d$
for any
$d \in {\mathbb Z}_{\geq 1}$
. By Theorem 2.1,
$M_{2,2}({\mathcal F}_\chi )$
is the limsup over L of the sums

The key observation is that the affine variety

in
${\mathbb A}^4$
is homogeneous, the affine cone over the projective variety
$\Sigma _{a,b,\mathrm {proj}} \subset {\mathbb P}^3$
defined by these same equations. We may omit the origin
$(0,0,0,0) \in {\mathbb A}^4$
without changing the large L limit. Then we choose, for each point in
$\Sigma _{a,b,\mathrm {proj}}(L)$
, a representative
$(x_0,y_0,z_0,w_0) \in \Sigma _{a,b}(L)$
. Then every point
$(x,y,z,w)$
in
$\Sigma _{a,b}(L) \smallsetminus \{ 0\}$
is uniquely of the form
$(rx_0,ry_0,rz_0,rw_0)$
with
$r \in L^\times $
and
$(x_0,y_0,z_0,w_0) \in \Sigma _{a,b}(L)$
a chosen representative. Moreover,

Thus, we are looking at the limsup over L of the sums

The innermost sum is
$O(\# L)^{1/2}$
so long as the polynomial

in r is not Artin-Schreier trivial. The number of L-valued points on
$\Sigma _{a,b,\mathrm {proj}}$
is
$O(\# L)$
, so the Artin-Schreier nontrivial cases contribute
$O((\# L)^{3/2})/(\# L)^2$
to the sum, and hence contribute
$0$
to the large L limit.
Because
$f(x)$
is Artin-Schreier reduced, the only way the polynomial

in r can be Artin-Schreier trivial is for every
$\Sigma _i(x_0,y_0,z_0,w_0)$
with
$i\in {\mathcal E}$
to vanish, in which case the inner sum is
$\# L-1$
. Thus, our large L limiting sum is

We break the domain of summation into finitely many closed points and the one-dimensional geometrically irreducible components Z of the projective variety
$\Sigma _{\mathrm {proj}}(S)$
defined by

each of which is defined over some finite extension of k. At the expense of enlarging k, we may assume each Z is defined over k. Then

So our
$\limsup $
is the
$\limsup $
of the sum

When
$\chi (xy)\bar \chi (zw)$
is geometrically trivial on (the dense open set where
$xyzw\neq 0$
of) Z, this sum over Z contributes
$1$
to the
$\limsup $
, while if
$\chi (xy)\bar \chi (zw)$
is geometrically nontrivial on (the dense open set where
$xyzw\neq 0$
of) Z, it contributes
$0$
to the limsup. Thus, we have

So is the number of geometrically irreducible components of dimension one in
$\Sigma _{\mathrm {proj}}(S)$
, while
$M_{2,2}({\mathcal F}_\chi )$
is the number of those components on which
$\chi (xy)\bar \chi (zw)$
is geometrically trivial.
Now assume that all integers in S are odd. Then
$\Sigma _{\mathrm {proj}}(S)$
contains the line
$x+y=0=z+w$
. For any character
$\chi $
of
$k^\times $
, the sum of
$\chi _L(xy)\bar \chi _L(zw)$
over this line is
$\# L-1$
if
and
$0$
otherwise. Thus, if
, this line is an irreducible component on which
$\chi (xy)\bar \chi (zw)$
is geometrically nontrivial, and hence the asserted inequality

if .
For ease of later reference, we given a slight generalization of this last result.
Theorem 2.4. Let k be a finite field of characteristic
$p>0$
, and
$f(x) \in k[x]$
, say
$f(x)=\sum _iA_ix^i$
, of degree
$d \ge 3$
which is Artin-Schreier reduced (meaning that
$A_i=0$
if
$p|i$
). Let
$n \ge 2$
, and let

be prime to p integers,
$\chi $
a multiplicative character of
$k^\times $
, and consider the local system
${\mathcal F}_\chi $
on
${\mathbb A}^n/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$(t_1,\ldots ,t_n) \in L^n$
,

with the convention that , but
$\chi _L(0)=0$
for
$\chi $
nontrivial. Consider the set
${\mathcal E}$
of exponents which occur in f:

and the affine locus
$\Sigma (S)$
as defined in (1.0.1) with
$S:=\{b_1,\ldots ,b_n\} \cup {\mathcal E}$
. Then

Moreover, if and all integers in S are odd, then

More precisely, is the number of geometrically irreducible components of dimension one in
$\Sigma _{\mathrm {proj}}(S)$
, while
$M_{2,2}({\mathcal F}_\chi )$
is the number of those components on which
$\chi (xy)\bar \chi (zw)$
is geometrically trivial.
Proof. The proof is essentially identical to that of the previous Theorem 2.3, which is the case
$n=2$
. Let us denote

The role of
$\Sigma _{a,b}$
there is played by
$\Sigma _B:=\cap _i\Sigma _{b_i}$
here. The affine variety
$\Sigma _B$
is homogeneous, the affine cone over the projective variety
$\Sigma _{B,\mathrm {proj}}$
defined by the same equations. Because
$n \ge 2$
, the projective variety
$\Sigma _{B,\mathrm {proj}}$
has dimension at most one (i.e., all its geometrically irreducible components have dimension
$\le 1$
), so over any finite extension,
$L/k$
has
$O(\# L) L$
-valued points. From here on, the proof is identical.
Corollary 2.5. In the setting of Theorem 2.3, with
$S:={\mathcal E} \cup \{a,b\}$
, write S as

with
$r \ge 2$
. Consider the local system
${\mathcal F}(S,\chi )$
on
${\mathbb A}^r$
, whose trace function is given as follows: For
$k/k_0$
a finite extension, and
$(t_1,\ldots ,t_r) \in {\mathbb A}^r(k)$
,

It is lisse of rank and pure of weight zero. [It is geometrically irreducible if and only if
$\gcd (S)=1$
, but that is irrelevant here.] Then
${\mathcal F}(f,a,b,\chi )$
has the same
$M_{2,2}$
as
${\mathcal F}(S,\chi )$
.
Proof. That
${\mathcal F}(S,\chi )$
has its
$M_{2,2}$
given by the same recipe, purely in terms of the data
$(S,\chi )$
, as did
${\mathcal F}(f,a,b,\chi )$
, is the special case
$f(x)=x^A$
,
$n=r$
, and
$b_i=B_{r+1-i}$
, of Theorem 2.4.
Theorem 2.6. Let k be a finite field of characteristic
$p>0$
, and
$f(x) \in k[x]$
, say
$f(x)=\sum _iA_ix^i$
, of degree
$d \ge 1$
which is Artin-Schreier reduced. Let
$1 \le a <b $
be prime to p integers, and suppose
$\deg (f) < b$
. For
$\chi $
a character of
$k^\times $
, consider the local system
${\mathcal F}_\chi $
on
$({\mathbb A}^1\times {\mathbb G}_m)/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$s \in L$
,
$t \in L^\times $
,

with the convention that , but
$\chi _L(0)=0$
for
$\chi $
nontrivial. Consider the set
${\mathcal E}$
of exponents which occur in f:

and the affine locus
$\Sigma (S)$
as defined in (1.0.1) with
$S:=\{a,b\} \cup {\mathcal E}$
.
-
(i) Suppose that
$f(x)$ is not of the form
$\mathrm {(nonzero\ constant)}x^a$ . Then
and all integers in S are odd, then
is the number of geometrically irreducible components of dimension one in
$\Sigma _{\mathrm {proj}}(S)$ , while
$M_{2,2}({\mathcal F}_\chi )$ is the number of those components on which
$\chi (xy)\bar \chi (zw)$ is geometrically trivial.
-
(ii) Suppose that
$f(x)=\mathrm {(nonzero\ constant)}x^a$ . If
, then
, we have
$$ \begin{align*}M_{2,2}({\mathcal F}_\chi) \le \limsup_{\# L \to \infty}\frac{\#\Sigma(S)(L)}{(\# L)^2}.\end{align*} $$
$a,b$ are both odd, and
, we have
$$ \begin{align*}M_{2,2}({\mathcal F}_\chi) < \limsup_{\# L \to \infty}\frac{\#\Sigma(S)(L)}{(\# L)^2}.\end{align*} $$
Proof. By Theorem 2.1 and the argument of Theorem 2.3,
$M_{2,2}$
for
${\mathcal F}_\chi $
is the
$\limsup $
over L of
$\frac {1}{(\# L)(\# L-1)}$
times

If the summation were over all
$(s,t)\in L^2$
, this would be

and just as in the proof of Theorem 2.3, we would get

However, the summation is only over
$(s,t)\in L\times L^\times $
. So we must subtract, for each
$L/k$
, the expression

So long as
$f(x)$
contains monomials of degree
$e_i \neq a$
, the ray calculation used in the proof of Theorem 2.3 shows that this limit (not just
$\limsup $
) vanishes. The assertion about
is proven exactly as in Theorem 2.3.
Suppose now that
$f(x)$
is a constant multiple of
$x^a$
and
. Then the term we are subtracting is equal to

which tends to
$1$
as L grows, simply because
$\Sigma _{a}$
is the affine cone over the smooth surface
$\Sigma _{a,\mathrm {proj}}$
.
Suppose finally that
$f(x)$
is a constant multiple of
$x^a$
and
. Then the sum we are subtracting is

This sum will be
$O(1/\sqrt {\# L})$
, and thus have large L limit zero, if the Kummer sheaf
${\mathcal L}_{\chi (xy/zw)}$
is geometrically nontrivial on the dense open set U of
$\Sigma _{a,\mathrm {proj}}$
where
$xyzw$
is invertible. Thus, U is the open set in the affine surface
$x^a+y^a=z^a+1$
where
$xyz$
is invertible, and our sheaf is
${\mathcal L}_{\chi (xy/z)}$
on U. We will show that this sheaf has a geometrically nontrivial pullback.
Choose an element
$\alpha \in {\mathbb F}_{p^2} \smallsetminus {\mathbb F}_p$
, and
$\beta $
with
$\beta ^a=\alpha $
. It suffices to show the pullback of
${\mathcal L}_{\chi (xy/z)}$
to the closed subscheme
$y=\beta $
of U is geometrically nontrivial. This pullback is
${\mathcal L}_{\chi (\beta x/z)}$
, on the open set of the curve

where
$xz$
is invertible. But the function
$\beta x/z$
on
${\mathcal C}$
has a simple zero at each point
$(0,\gamma )$
with
$\gamma $
one of the a distinct roots of the polynomial
$T^a=\alpha -1$
. Hence,
${\mathcal L}_{\chi (\beta x/z)}$
is geometrically nontrivial on
${\mathcal C}$
.
So in this case when
$f(x)$
is a constant multiple of
$x^a$
and
, we have

[Of course, in this case, the set
$S =\{a,b\}$
.] The argument in the proof of Theorem 2.3 shows that if
$a,b$
are both odd, but
, then

The following result explains the moment drop for some local systems.
Theorem 2.7. Let k be a finite field of odd characteristic
$p>0$
, and
$f(x) \in k[x]$
, say
$f(x)=\sum _iA_ix^i$
, of degree
$d \ge 1$
which is Artin-Schreier reduced. Let
$1 \le a <b $
be prime to p integers, and suppose
$\deg (f) \neq b$
. Consider the local system
${\mathcal F}$
on
$({\mathbb A}^1\times {\mathbb G}_m)/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$s \in L$
,
$t \in L^\times $
,

Suppose further that f is odd (i.e., that
$f(-x)=-f(x)$
), that f is not a constant multiple of
$x^a$
, and that both
$a,b$
are odd. Let
$g(x) \in k[x]$
, say
$g(x)=\sum _iB_ix^i$
, of degree
$e \ge 1$
which is Artin-Schreier reduced. Consider the local system
${\mathcal G}$
on
$({\mathbb A}^1\times {\mathbb G}_m)/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$s \in L$
,
$t \in L^\times $
,

Then
$M_{2,2}({\mathcal G}) \le M_{2,2}({\mathcal F})-1$
.
Proof. Consider the set
${\mathcal E}_f$
of exponents which occur in f:

and
$S_f:=\{a,b\} \cup {\mathcal E}_f$
.
Consider also the set
${\mathcal E}_{f,+}$
of exponents with occur in
$f(x)+g(x^2)$
:

and
$S_{f,+}:=\{a,b\} \cup {\mathcal E}_{f,+}$
. Then from Theorems 2.3 and 2.6, we know that


As
$S_f \subset S_{f,+}$
, we trivially have
$M_{2,2}({\mathcal G}) \le M_{2,2}({\mathcal F})$
. Because
$S_f$
consists entirely of odd integers, among the two-dimensional geometrically irreducible components of
$\Sigma (S_f)$
is the locus
$x+y=0=z+w$
.
It suffices to show that this locus
$x+y=0=z+w$
does not lie in
$\Sigma (S_{f,+})$
. Indeed,
$S_{f,+}$
contains some nonzero even integer
$2j $
, and hence,
$\Sigma (S_{f,+})$
lies inside the hypersurface of equation
$x^{2j}+y^{2j}=z^{2j}+w^{2j}$
. So it suffices to show that the locus
$x+y=0=z+w$
does not lie in this hypersurface. The intersection of this hypersurface with the locus
$x+y=0=z+w$
is the locus in
$(x,z)$
space defined by
$x^{2j} +(-x)^{2j}=z^{2j} +(-z)^{2j}$
. As we are in odd characteristic, this intersection is the locus
$x^{2j}=z^{2j}$
, which is the union of
$2j$
lines.
3
$M_{2,2}$
and resultants
We will need the following property of resultants, which is well known:
Lemma 3.1. Let R, S be commutative rings,
$f,g \in R[x]$
, and let
$\varphi : R \to S$
be a ring homomorphism.
-
(i) If
$\varphi (\mathsf {Res}(f,g)) \neq 0$ , then
$\mathsf {Res}(\varphi (f),\varphi (g))$ (computed as the resultant of two polynomials in
$S[x]$ ) is also nonzero.
-
(ii) If
$\varphi $ preserves the degree of each of f and g, then
$\mathsf {Res}(\varphi (f),\varphi (g)) \neq 0$ implies
$\varphi (\mathsf {Res}(f,g)) \neq 0$ .
-
(iii) If S is an integral domain and
$\varphi $ preserves the degree of at least one of f and g, then
$\mathsf {Res}(\varphi (f),\varphi (g)) \neq 0$ if and only if
$\varphi (\mathsf {Res}(f,g)) \neq 0$ .
Proof. (i) Assume that
$\varphi (\mathsf {Res}(f,g)) \neq 0$
. Let
$f(x)$
be of degree d and with leading term
$ax^d$
, and let
$g(x)$
be of degree e and with leading term
$bx^e$
. Suppose that
$\varphi (a)=\varphi (b)=0$
, so that
$\varphi (f) \in S[x]$
has degree
$<d$
and
$\varphi (g) \in S[x]$
has degree
$<e$
. In this case,
$\varphi (\mathsf {Res}(f,g))=0$
, a contradiction. So we may assume that
$\varphi (a) \neq 0$
, so that
$\varphi (f) \in S[x]$
has degree d. Now, if
$\varphi (g)$
has degree
$e' \leq e$
, then

and hence,
$\mathsf {Res}(\varphi (f),\varphi (g)) \neq 0$
.
(ii) follows from (3.1.1) (with
$e'=e$
).
(iii) follows from (i), (3.1.1), and the assumption that S is an integral domain.
Fix a prime p. First we look at any set
${\mathcal Q}:= \{q_1 < \ldots < q_n\}$
of
$n \geq 1$
positive powers of p, and consider

In the special case of characteristic
$p=2$
, we have
$(-1)^p=1$
, and so

The following observation is helpful in computing
$\mu _{total}({\mathcal Q})$
.
Lemma 3.2. Let
$n \geq 2$
, p any prime,
$q=p^f$
,
$q_i = q^{m_i}$
for
$1 \leq i \leq n$
, and
$m_1 < \ldots < m_n$
. Also let
$e:=\gcd (m_1, \ldots ,m_n)$
. Then

Proof. The statement is obvious when
$p = 2$
, so we will assume
$p> 2$
. Replacing q by
$q^e$
, we may assume that
$\gcd (m_1, \ldots ,m_n)=e=1$
. Suppose
$2|m_i$
,
$2 \nmid m_j$
, and
$\zeta \in \mu _{total}({\mathcal Q})$
. Since
$\zeta ^{q^{m_j}-1}=-1$
and
$m_j$
is odd, we see that the
$2$
-part
$2^f$
of the order of
$\zeta $
is
$2(q^{m_j}-1)_2 = 2(q-1)_2$
, twice the
$2$
-part of
$q-1$
. As
$p> 2$
,
$2^f$
divides
$(q^2-1)_2$
, which in turn divides
$q^{m_i}-1$
because
$2|m_i$
, and this contradicts the equality
$\zeta ^{q^{m_i}-1} = -1$
.
Assume now that
$2 \nmid m_i$
for all i, so that
$2 \nmid (q^{m_i}-1)/(q-1)$
, and choose a primitive
${(2q-2)}^{\mathrm {th}}$
root of unity
$\theta \in \overline {{\mathbb F}_p}$
. Then
$-1=\theta ^{q-1}=\theta ^{q^{m_i}-1}$
, and hence,
$\zeta \in \mu _{total}({\mathcal Q})$
if and only if
$(\zeta \theta )^{q^{m_i}-1}=1$
for all i. There are exactly

possibilities for such
$\zeta \theta $
.
For any
$a \in {\mathbb Z}_{\ge 2}$
, let

Note that
${\mathcal M}_p(a)$
is finite (by looking at the condition at
$j=a$
) if
$p \nmid a$
. In fact,
${\mathcal M}_p(2) = \varnothing $
if
$p> 2$
,
${\mathcal M}_p(3) = \varnothing $
if
$p> 3$
; more generally,
${\mathcal M}_p(a)=\varnothing $
if
$2 \leq a < p$
or if
$p \nmid a(a-1)$
. As we will see in the proof of Proposition 3.3 (see (3.3.6)), for
$q=p^f$
, we have

We also set

Keep the notation

Recall that for a positive integer n, its prime to p part
$n_{p'}$
is the prime to p integer such that
$n=p^en_{p'}$
for some
$e \ge 0$
.
Proposition 3.3. Let
$2 \leq b < c$
be integers coprime to p. For finite extensions
$L/{\mathbb F}_p$
, the following statements hold for the set
$\Sigma (L)$
of L-points of the surface

of
${\mathbb A}^4(x,y,z,w)$
.
-
(i)
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2 \geq 2+\#\big ( {\mathcal M}_p(b) \cap {\mathcal M}_p(c) \big )$ .
-
(ii) If the resultant
$R(A):= \mathsf {Res}_v(F_b(A,v),F_c(A,v))$ of the polynomials
$F_b(A,v)$ and
$F_c(A,v)$ in the variable v is not identically zero as a function of A, then
$$ \begin{align*}\lim_{\# L \to \infty}\#\Sigma(L)/(\# L)^2 = 2+\#\big( {\mathcal M}_p(b) \cap {\mathcal M}_p(c) \big).\end{align*} $$
-
(iii) If
$b=2 < p$ , then
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2 = 2$ .
-
(iv) If
$b=3 < p$ , then
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2$ equals
$2$ when
$2|c$ and
$3$ when
$2 \nmid c$ .
-
(v) The equality in (ii) holds if
$b=p^f+1$ .
-
(vi) Suppose
$\gcd \big ((b-1)_{p'},(c-1)_{p'}\big )=1$ . Then
$R(A) \not \equiv 0$ , and hence, the equality in (ii) holds.
-
(vii) If
$\gcd (b-1,c-1)=1$ , then
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2 = 2$ .
-
(viii) If
$\gcd \big ((b-1)_{p'},(c-1)_{p'}\big )=1$ ,
$p> 2$ , and
$(b-1)_p = p^{fm}$ ,
$(c-1)_p = p^{fn}$ with
$f,m,n \in {\mathbb Z}_{\geq 1}$ ,
$\gcd (m,n)=1$ , and
$2|mn$ , then we also have
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2 = 2$ .
Proof. For (i), consider any point
$P=(x,y,z,w) \in \Sigma (L)$
. Then
$x+y=z+w$
. Certainly,
$\Sigma $
contains the two planes

which contribute
$2(\# L)^2-\# L$
points to
$\Sigma (L)$
. So we have to count the points
$P \in \Sigma (L)$
for which
$z \neq x,y$
. For these points, we can use the parametrization

for P in terms of
$A,u,y$
, where
$u:=z-y \neq 0$
and
$A := (x-z)/(z-y) \neq 0$
. The condition
$\Sigma _b(P)=0$
now reads

First, we look at such points P with
$y=0$
. Since
$u \neq 0$
, (3.3.2) implies
$(A+1)^b-A^b-1=0$
. The leading term of this polynomial equation in A is
$bA^{b-1}$
. Since
$p \nmid b$
, there are at most
$b-1$
such A’s, which contributes at most
$(b-1)\# L$
points to
$\Sigma (L)$
. This dies in the large L-limit.
So we may now assume
$y \neq 0$
, and replace
$(A,y,u)$
by
$(A,y,v)$
, where
$v:= u/y \neq 0$
. Since
$y \neq 0$
, now (3.3.2) becomes

Note that the coefficient for
$v^j$
in the left-hand side of (3.3.3) is

when
$2 \leq j \leq b$
, and
$0$
if
$j = 0,1$
. So the condition
$P \in \Sigma (L)$
now reads

Furthermore, if
$A \in {\mathcal M}_p(b)$
, then (3.3.3) is vacuously true. Hence, if

then (3.3.4) is vacuously true, and each A contributes
$(\# L-1)^2$
points to
$\Sigma (L)$
with
$y,v \neq 0$
, which do not belong to the two planes
$(x=z,~y=w)$
and
$(x=w,~y=z)$
. This yields the lower bound in (i).
Now we look at
$A \notin {\mathcal M}_p(b) \cap {\mathcal M}_p(c)$
, and assume that
$R(A) \not \equiv 0$
as a function of A. Applying Lemma 3.1 to the specialization homomorphism
$A \mapsto \gamma $
at any point
$\gamma $
where
$R(\gamma ) \neq 0$
, we see that (3.3.4) has no solution v when
$A = \gamma $
. Thus, (3.3.4) can have solutions in v only at
$A=\gamma $
with
$R(\gamma )=0$
. This implies that the number of A for which (3.3.4) has a common solution in v is bounded independently of L (in fact by
$2bc$
, an upper bound for the degree of
$R(A)$
). If
$A \notin {\mathcal M}_p(b)$
for instance, then
$F_b(A,v)$
is a nonzero polynomial in v, and hence has at most b zeros once A is fixed. Thus, each such A contributes at most
$\max (b,c)(\# L-1)$
points to
$\Sigma (L)$
(with y running), and again this dies in the large L-limit. This proves the equality in (ii).
Suppose
$b=2 < p$
. Then
$F_2(A,v)=2$
, and hence, (3.3.4) has no solutions. Furthermore,
${{\mathcal M}_p(2)=\varnothing }$
, proving (iii).
Suppose
$b=3 < p$
. Then
$F_3(A,v)=3((A+1)v+2)$
. Hence, (3.3.4) is equivalent to
$(A+1)v=-2$
and
$(-1)^c-(-v-1)^c-(v+1)^c+1=0$
. If
$2|c$
, this shows that
$(v+1)^c=1$
. Thus, there are at most c pairs
$(A,v)$
that satisfy (3.3.4), contributing at most
$c(\# L-1)$
points to
$\Sigma (L)$
, and this dies in the large L-limit. Suppose
$2 \nmid c$
. This argument then shows that there are exactly
$\# L-2$
pairs
$(A,v)$
that satisfy (3.3.4) and
$A,v \neq 0$
(namely, one for each
$v \neq 0,-2$
). This gives
$(\# L-1)(\# L-2)$
more points to
$\Sigma (L)$
, proving (iv).
Next, suppose that
$b= q+1$
with
$q := p^f \geq p$
. Then (3.3.3) becomes

which shows that

that is,
$A \in \mu _{total}(\{q\})$
. Now, if
$A \notin {\mathcal M}_p(b)$
, then (3.3.5) has no solution since
$v \neq 0$
, and hence, (3.3.3), respectively (3.3.4), has no solution. If
$A \in {\mathcal M}_p(b) \smallsetminus {\mathcal M}_p(c)$
, then we have at most
$b-2=q-1$
possibilities for A, for each of which
$F_c(A,v)=0$
yields at most c possibilities for v. This contributes at most
$(b-2)c(\# L-1)$
points to
$\Sigma (L)$
, and this dies in the large L-limit. Hence, we have to count only the A’s in
${\mathcal M}_p(b) \cap {\mathcal M}_p(c)$
, and hence, (v) holds.
For (vi), note that the coefficient for
$v^{j-2}$
in
$F_b(A,v)$
is

when
$2 \leq j \leq b$
. Hence,

Thus, the only roots of
$F_b(0,v)$
are the elements of
$\mu _{(b-1)_{p'}} \smallsetminus \{1\}$
(subtracted by
$1$
). Similarly, the set of roots of
$F_c(0,v)$
is
$\mu _{(c-1)_{p'}} \smallsetminus \{1\}$
(translated by
$-1$
). So the assumption
$\gcd \big ( (b-1)_{p'},(c-1)_{p'}\big ) = 1$
implies that
$F_b(0,v)$
and
$F_c(0,v)$
have no common root. Furthermore, the specialization
$A \mapsto 0$
preserves the degree
$b-2$
of
$F_b(A,v)$
(as
$p \nmid b$
). It follows from Lemma 3.1 that
$R(0) \neq 0$
, and so
$R(A) \not \equiv 0$
.
Note that (vi) implies (iii) and (v), since
$(b-1)_{p'} = 1$
when
$b=p^f+1$
with
$f \geq 0$
.
Assume now that
$\gcd (b-1,c-1)=1$
. If
$p \nmid b(b-1)$
, then the
$j=2$
condition in the definition of
${\mathcal M}_p(b)$
is the vanishing of

and hence,
${\mathcal M}_p(b) = \varnothing $
, implying
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2 = 2$
by (vi). If
$p \nmid c(c-1)$
, the same
$j=2$
condition shows that
${\mathcal M}_p(c) = \varnothing $
, with the same conclusion that
$\lim _{\# L \to \infty }\#\Sigma (L)/(\# L)^2 = 2$
by (vi).
For (viii), note that
$(b-1)_p=p^{fm}$
implies that
$p \nmid \binom {b}{j}$
for
$j = p^{fm}+1$
. Now

where
$q:= p^f$
. Thus,
${\mathcal M}_p(b)$
is contained in
$\{A \mid A^{q^m-1}=-1\}$
. Similarly,
${\mathcal M}_p(c)$
is contained in
$\{A \mid A^{q^n-1}=-1\}$
. By Lemma 3.2, the set
$\{A \mid A^{q^m-1}=A^{q^n-1}=-1\}$
is empty, and so we are done by (vi).
We will need the following well-known observation:
Lemma 3.4. Let p be a prime and
$n = \sum _{i \geq 0}n_ip^i$
and
$m = \sum _{i \geq 0}m_ip^i$
be the base p expansions of integers
$n, m \geq 1$
. Suppose that
$m_i \leq n_i$
for all i. Then
$p \nmid \binom {n}{m}$
.
Proof. The hypothesis implies that
$n-m = \sum _{i \geq 0}(n_i-m_i)p^i$
is the base p expansion of
$n-m$
. Now for any
$j \geq 0$
, we have

Since
$\sum _{j \geq 0}\lfloor \frac {n}{p^j} \rfloor $
is the exponent of the highest power of p that divides
$n!$
, and similarly for
$m!$
and
$(n-m)!$
, the above equalities imply the claim.
Proposition 3.5. Fix a prime p, integers
$n,r \geq 1$
, and consider prime to p integers

For finite extensions
$L/{\mathbb F}_p$
, consider the set
$\Sigma (L)$
of L-points of the surface

of
${\mathbb A}^4(x,y,z,w)$
. Then exactly one of the following statements holds for

-
(a)
$p> 2$ and
$M=2$ .
-
(b)
$p=2$ and
$M = 3$ .
-
(c)
$p> 2$ and
$M= p^e+1 \geq 4$ , where
$b_i = p^{m_i}+1$ for
$1 \leq i \leq r$ , and the integers
$n/e$ ,
$m_1/e, \ldots , m_r/e$ are all odd for
$e:= \gcd (n,m_1, \ldots ,m_r)$ .
-
(d)
$p = 2$ and
$M= 2^e+1 \geq 5$ , where
$b_i = 2^{m_i}+1$ for
$1 \leq i \leq r$ , and
$e:= \gcd (n,m_1, \ldots ,m_r) \geq 2$ .
Proof. We will follow the proof of Proposition 3.3 and count the points

that lie outside of the two planes
$(x=z,y=w)$
and
$(x=w,y=z)$
, for which we can use the parametrization (3.3.1). For these points, the condition
$P \in \Sigma $
reads

cf. (3.3.4). Since
$\Sigma _a(P)=0$
and
$a=p^n+1$
, we have
$A^{p^n-1}=-1$
; see (3.3.5). Now the proof of Proposition 3.3(v) can be repeated verbatim to show that

where
${\mathcal M}_p(a)$
and
${\mathcal M}_p({b_i})$
are defined in (3.2.1).
We will assume that
$M> 2$
if
$p> 2$
,
$M> 3$
if
$p=2$
, and aim to show that we are in (c) with
$M=p^e+1 \geq 4$
or in (d) with
$M=2^e+1 \geq 5$
. Note that when
$p=2$
,
$1 \in {\mathcal M}_p(b)$
for any integer
$b \geq 3$
. Hence, our assumption implies that

Consider the base p expansion

of
$c:=b_1$
. We already noted that
${\mathcal M}_p(c) = \varnothing $
if
$2 < p \nmid c(c-1)$
, contrary to (3.5.1). However, if
$p=2$
, then
$2 \nmid c$
and so
$p|(c-1)$
. Henceforth, we may assume that
$p|(c-1)$
, whence
$c_0=1$
.
Consider any digit
$c_i \geq 1$
of c, with
$i \geq 1$
. By Lemma 3.4,
$p \nmid \binom {c}{p^i+1}$
. Taking
$j:=p^i+1$
in the definition (3.2.1) of
${\mathcal M}_p(c)$
, we get

for
$A \in {\mathcal M}_p(c)$
. As
$A \neq 0$
, we get

in particular,

Assume in addition that
$c_i \geq 2$
(and so
$p>2$
as
$c_i \leq p-1$
.) Then by Lemma 3.4, we have
$p \nmid \binom {c}{2p^i+1}$
. Taking
$j:=2p^i+1$
in the definition (3.2.1) of
${\mathcal M}_p(c)$
, we get

and so
$A=-1$
. But this is impossible by (3.5.2) (since
$p^i \geq 3$
is odd in the case under consideration), and so
${\mathcal M}_p(c)=\varnothing $
, again contradicting (3.5.1).
We have shown that any positive digit
$c_i$
of c must be equal to
$1$
. Suppose now that
$c_i=1=c_j$
for some
$i> j \geq 1$
. Then (3.5.2) holds for both
$A^{p^i}$
and
$A^{p^j}$
, and so

Furthermore, by Lemma 3.4, we have
$p \nmid \binom {c}{p^i+p^j+1}$
. Taking
$j:=p^i+p^j+1$
in the definition (3.2.1) of
${\mathcal M}_p(c)$
, we now get

and so
$A=-1$
. If
$p> 2$
, then this is again impossible by (3.5.2), and so
${\mathcal M}_p(c)=\varnothing $
, contrary to (3.5.1). If
$p=2$
, then
${\mathcal M}_p(c) \subseteq \{1\}$
, contradicting (3.5.1).
We have shown that
$b_1=c$
has only two positive digits,
$c_0$
and
$c_{m_1}$
, and both are equal to
$1$
. Thus,
$b_1=p^{m_1}+1$
. Applying the same argument to any
$b_i$
, we see that
$b_i=p^{m_i}+1$
. Hence,

Let
$e:=\gcd (n,m_1, \ldots ,m_r)$
. If
$p> 2$
, then it follows from Lemma 3.2 that

equals
$p^e+1$
if all
$n/e$
and
$m_i/e$
are odd, and
$0$
otherwise, and thus, we arrive at (c). Similarly, if
$p=2$
, then using Lemma 3.2, we arrive at (d).
Corollary 3.6. Fix a prime p, a power
$q=p^f$
, an integer
$r \geq 1$
, and consider
$q_i := q^{m_i}$
with
$1 \leq m_1 < \ldots < m_r$
and
$\gcd (m_1, \ldots ,m_r)=1$
. If
$p>2$
, assume in addition that
$2 \nmid m_1m_2 \ldots m_r$
. For finite extensions
$L/{\mathbb F}_p$
, consider the set
$\Sigma (L)$
of L-points of the surface

of
${\mathbb A}^4(x,y,z,w)$
. Then

Proof. Arguing as in the proof of Proposition 3.5, we have

According to (3.3.6),
$\cap ^r_{i=1} {\mathcal M}_p(q_i+1)$
is precisely
$\mu _{total}({\mathcal Q})$
for
${\mathcal Q}:= \{q_1, \ldots ,q_r\}$
. The statement now follows from Lemma 3.2.
In hindsight, Corollary 3.6 is a reflection of [Reference Katz and TiepKT6, Theorem 16.7(i-bis), (ii)] and the fact that
$\mathrm {SU}_N(q)$
acting on the natural module
${\mathbb F}_{q^2}^N$
, respectively
$\Omega ^{-}_{2N}(q)$
acting on the natural module
${\mathbb F}_q^{2N}$
when
$p=2$
, has at least
$q+1$
orbits. (Also see Theorem 1.5 and Lemma 5.1 of [Reference Guralnick and TiepGT2].)
Theorem 3.7. Let k be a finite field of characteristic
$p>0$
, and
$f(x) \in k[x]$
, say
$f(x)=\sum _iA_ix^i$
, of degree
$d \ge 1$
which is Artin-Schreier reduced. Let
$1 \le a <b $
be prime to p integers. Suppose that we are in one of the following two situations.
-
(a) We have
$1 \le a <b < \deg (f)$ . We consider the local system
${\mathcal F}$ on
${\mathbb A}^2/k$ whose trace function is given as follows: for
$L/k$ a finite extension, and
$s,t \in L$ ,
$$ \begin{align*}\mathrm{Trace}(\mathrm{Frob}_{(s,t),L}|{\mathcal F})=\frac{-1}{\sqrt{\# L}}\sum_{x\in L}\psi_L(sx^a +tx^b +f(x)).\end{align*} $$
-
(b) We have
$1 \le a <b$ ,
$\deg (f) < b$ ,
$f(x)$ is not of the form
$\mathrm {(nonzero\ constant)}x^a$ . We consider the local system
${\mathcal F}$ on
$({\mathbb A}^1\times {\mathbb G}_m)/k$ whose trace function is given as follows: for
$L/k$ a finite extension, and
$s,t \in L\times L^\times $ ,
$$ \begin{align*}\mathrm{Trace}(\mathrm{Frob}_{(s,t),L}|{\mathcal F})=\frac{-1}{\sqrt{\# L}}\sum_{x\in L}\psi_L(f(x) +sx^a +tx^b).\end{align*} $$
Consider the set
${\mathcal E}$
of exponents which occur in f:

and denote by S the set

Choose two elements
$A\neq B$
in S, and enumerate the set S:

Suppose that

Then we have the following results.
-
(i) Suppose
$B=2A$ . Then
$M_{2,2}({\mathcal F})=2$ .
-
(ii) Suppose
$B=3A$ . Then
$M_{2,2}({\mathcal F})=3$ if every
$s \in S$ is odd, and
$M_{2,2}({\mathcal F})=2$ if some
$s \in S$ is even.
Proof. The idea is to make use of the limsup formulas of Theorems 2.3 and 2.6 to compute
$M_{2,2}$
.
Consider first the case when
$B=2A$
. Then the two equations

which we view as the equations
$\Sigma _1=\Sigma _2=0$
applied to the variables
$x^A,y^A,z^A, w^A$
, show that we have an equality of sets

If any of
$x,y,z,w$
vanishes, this equality of sets has
$O(\# L)$
solutions, so we may assume that each of
$x,y,z,w$
is nonzero. Then we are in one of
$2A^2$
cases, as follows. For each ordered pair
$\zeta ,\eta $
of
$A^{\mathrm {th}}$
roots of unity in
$\mu _A(\overline {k})$
, either

In this first case of
$[z,w] =[\zeta x,\eta y]$
, we use the various
$\Sigma _{C_i}$
equations, that
$x^{C_i}+y^{C_i}=z^{C_i}+w^{C_i}$
, to get

This equation has
$O(\# L)$
solutions unless we have
$\zeta ^{C_i}=\eta ^{C_i}=1$
. But
$\gcd (\mathrm {the\ }C_i,A,2A)=1$
, and hence,
$\gcd (\mathrm {the\ }C_i,A)=1$
. So in order to have more than
$O(\# L)$
solutions, we must have

As both
$\zeta ,\eta $
are
$A^{\mathrm {th}}$
roots of unity, and
$\gcd (\mathrm {the\ }C_i,A)=1$
, these equalities force
$\zeta =1=\eta $
, Thus, in this first case, we have the solution
$[z,w] =[x,y]$
, with its
$(\# L)^2$
points, and
$A^2-1$
other solutions, each with
$\# L$
points. The treatment of the second case,
$[w,z] =[\zeta x,\eta y]$
, is identical.
Consider now the case when
$B=3A$
. Then the two equations

which we view as the equations
$\Sigma _1=\Sigma _3=0$
applied to the variables
$x^A,y^A,z^A, w^A$
, show that either we have an equality of sets

or we have the relations

Exactly as in the
$B=2A$
discussion above, we use the fact that

to show that from the equality of sets
$\{x^A,y^A\}=\{z^A,w^A\}$
we get, up to
$O(\# L)$
, the
$(\# L)^2$
solutions
$\{x,y\}=\{z,w\}.$
It remains to deal with with the equation
$x^A+y^A=0=z^A+w^A.$
Fix an
$A^{\mathrm {th}}$
root
$\tau $
of
$-1$
. Then this breaks into the
$A^2$
cases
$ y =\tau \zeta x, z=\tau \eta w$
, for each pair
$\zeta ,\eta $
of
$A^{\mathrm {th}}$
roots of unity. We then use the
$\Sigma _{C_i}$
equations to obtain the relations

In order to get more than
$O(\# L)$
solutions, we must have

Suppose first that A is odd. Then we take
$\tau :=-1$
, and our equations become

If all
$C_i$
are odd, these are the equations

In order to get more than
$O(\# L)$
solutions, we must have
$\zeta =1=\eta $
.
Suppose next that A is odd but some
$C_i$
is even, say
$C_1$
is even. (This can only happen if we are in odd characteristic, as f is Artin-Schreier reduced, and
$p \nmid ab$
.) Then we have the equation

But
$\zeta $
and
$\eta $
are roots of unity of odd order, so no powers of either can be
$-1$
. So in this case, we have only
$x=y=z=w=0$
.
Finally, consider the case when A is even. Then
$\gcd (\mathrm {the\ }C_i,A)=1$
, so there is some odd
$C_i$
, say
$C_1$
is odd. Then the two equations

rewritten as

and raised to the A power, give

But
$\zeta ^A=\eta ^A=1$
, so we get
$\tau ^{AC_1}=1=\eta ^{AC_1}$
. But
$\tau ^A=-1$
and
$C_1$
is odd, so we get
$-1=1$
, which is nonsense. Thus, in this case as well, the only solution is
$x=y=z=w=0$
.
Theorem 3.8. Let k be a finite field of characteristic
$p>0$
, and
$f(x) \in k[x]$
, say
$f(x)=\sum _iA_ix^i$
, of degree
$d \ge 3$
which is Artin-Schreier reduced (meaning that
$A_i=0$
if
$p|i$
). Let
$1 \le a <b < \deg (f)$
be prime to p integers, and consider the local system
${\mathcal F}$
on
${\mathbb A}^2/k$
whose trace function is given as follows: for
$L/k$
a finite extension, and
$s,t \in L$
,

Consider the set
${\mathcal E}$
of exponents which occur in f:

Suppose that the set
$\{a,b\}\cup {\mathcal E}$
contains
$1$
, c, and d, where
$1 < c < d$
and either of the following conditions is satisfied.
-
(i)
$\gcd (c-1,d-1)=1$ .
-
(ii)
$p> 2$ ,
$\gcd \big ((c-1)_{p'},(d-1)_{p'}\big )=1$ , and
$(c-1)_p = p^{fm}$ ,
$(d-1)_p = p^{fn}$ with
$f,m,n \in {\mathbb Z}_{\geq 1}$ ,
$\gcd (m,n)=1$ , and
$2|mn$ .
Then
$M_{2,2}({\mathcal F})=2$
.
Proof. The local system
${\mathcal F}$
is pure of weight zero, so geometrically semisimple, and of rank

so has
$M_{2,2}({\mathcal F}) \ge 2$
. Thus, it suffices to show that
$M_{2,2}({\mathcal F}) \le 2$
under the stated hypotheses. Now apply Theorem 2.3 and Proposition 3.3(vii), (viii).
4 p-finite and strongly p-finite data
In this and the next section, we consider local systems
${\mathcal F}$
on
${\mathbb A}^r/{\mathbb F}_p$
defined as follows. We are given a list of integers

For
$L/{\mathbb F}_p$
a finite extension, and
$(t_1,\ldots , t_r) \in L^r$
,

Here, we make a choice of
$\sqrt {p} \in \overline {{\mathbb Q}_\ell }$
, and define
$\sqrt {\# L}:=\sqrt {p}^{\deg (L/{\mathbb F}_p)}$
. We will name this
${\mathcal F}$
as

when confusion about ‘which
${\mathcal F}$
?’ is possible. Recall from [Reference Katz and TiepKT4, 2.5, 2.6] that such an
${\mathcal F}$
is geometrically irreducible.
When
$r=1$
, these local systems were the main subject of study in Chapter 10 of [Reference Katz and TiepKT6]. In general, the local systems
with finite
$G_{\mathrm {geom}}$
(and their
$G_{\mathrm {geom}}$
) have been classified in Chapter 11 of [Reference Katz and TiepKT6], some of whose results can be stated using the following notion.
Definition 4.1. Data
$(A,B_1,\ldots ,B_r)$
with
$r \geq 1$
subject to (4.0.1) is said to be p-finite if one of the following conditions holds.
-
(i)
$p> 2$ ,
$q=p^f$ ,
$A = (q^n+1)/2$ , and
$B_i = (q^{m_i}+1)/2$ for
$1 \leq i \leq r$ and
$n> m_1 > \ldots > m_r \geq 0$ are integers such that
$2|nm_1 \ldots m_r$ and
$\gcd (n,m_1, \ldots ,m_r)=1$ .
-
(ii)
$q=p^f$ and
$A = q^n+1$ . Furthermore, either
$(r,B_1,n) = (1,1,1)$ , or
$r \ge 2$ and
$B_i = q^{m_i}+1$ ,
$1 \leq i \leq r-1$ , where
$n> m_1 > \ldots > m_{r-1} \geq 0$ are integers with
$\gcd (n,m_1, \ldots ,m_{r-1})=1$ , and
$B_r=1$ .
-
(iii)
$p = 2$ ,
$q=2^f$ ,
$A = q^n+1$ ,
$B_i = q^{m_i}+1$ ,
$1 \leq i \leq r$ , where the integers
$n> m_1 > \ldots > m_r \geq 1$ are such that
$2|nm_1 \ldots m_r$ and
$\gcd (n,m_1, \ldots ,m_r)=1$ .
-
(iv)
$q=p^f$ ,
$A=(q^n+1)/(q+1)$ ,
$B_i = (q^{m_i}+1)/(q+1)$ ,
$1 \leq i \leq r$ , where
$n> m_1 > \ldots > m_r \geq 1$ are odd integers with
$\gcd (n,m_1, \ldots ,m_r)=1$ .
-
(v)
$p=2$ ,
$(A,B_1,\ldots , B_r) = (13,3)$ or
$(13,3,1)$ .
-
(vi)
$p=3$ ,
$1 \leq r \leq 3$ ,
$A=7$ ,
$\{B_1, \ldots ,B_r\} \subseteq \{4,2,1\}$ .
-
(vii)
$p=3$ ,
$1 \leq r \leq 3$ ,
$A=5$ ,
$\{B_1, \ldots ,B_r\} \subseteq \{4,2,1\}$ .
-
(viii)
$p=5$ ,
$1 \leq r \leq 2$ ,
$A=3$ ,
$\{B_1, \ldots ,B_r\} \subseteq \{2,1\}$ .
-
(ix)
$p=5$ ,
$r=1$ ,
$A=7$ ,
$B_1=1$ .
-
(x)
$p=7$ ,
$r=1$ ,
$A=5$ ,
$B_1=2$ .
Definition 4.2. Data
$(A,B_1,\ldots ,B_r)$
with
$r \geq 1$
subject to (4.0.1) is said to be strongly p-finite, if it satisfies 4.1(i) with
$q \in \{3,5\}$
, 4.1(ii) with
$r \geq 2$
and either
$q=2$
, or
$2 \nmid q$
but
$2|nm_1 \ldots m_{r-1}$
, 4.1(iii) with
$q=2$
, 4.1(iv) with
$q=2$
, or one of (v)–(x) of 4.1.
Theorem 4.3. A local system in characteristic p subject to (4.0.1) has finite
$G_{\mathrm {geom}}$
if and only if
$(A,B_1,\ldots ,B_r)$
is p-finite. If the data is strongly p-finite, then
$M_{2,2}({\mathcal F})$
equals
$2$
if
$2|AB_1 \ldots B_r$
, and
$3$
otherwise.
Proof. The first statement summarizes Theorems 10.2.6, 10.3.13 and 11.2.3 of [Reference Katz and TiepKT6]. The second statement follows from the explicit determination of
$G_{\mathrm {geom}}$
and [Reference Guralnick and TiepGT2, Theorem 1.5], if we assume in addition that
$A> 9$
in the cases of 4.1(ii), (iii) with
$q=2$
. Assume we are in the cases of 4.1(ii), (iii) with
$q=2$
and
$A =2^n+1 \leq 9$
. Now if
$B_r =1$
(so we are in 4.1(ii) with
$r \geq 2$
), then
$M_{2,2}=3$
by Corollary 3.6. Thus, we are left with the cases where
$p=2$
and, moreover,
$(A,B_1, \ldots ,B_r)= (5,3)$
,
$(9,5)$
,
$(9,5,3)$
. The third case has
$M_{2,2}=3$
by Theorem 3.7. The two remaining local systems of rank
$8$
and
$4$
, with
$r=1$
and
$(A,B_1)=(9,5)$
,
$(5,3)$
, are dealt with in the next result, which also resolves some open cases left in [Reference Katz and TiepKT6, Chapter 8].
Theorem 4.4. Suppose
$p=2$
. Then the following statements hold.
-
(i) Each of the following local systems
,
, and
${\mathcal H}_{53}:={\mathcal H} yp(\mathsf {Char}^\times _5,\mathsf {Char}^\times _3)$ has geometric monodromy group
$2^{1+4}_- \cdot \mathsf {A}_5$ , which is also the arithmetic monodromy group over any finite extensions of
${\mathbb F}_4$ . For each of them, the arithmetic monodromy group over
${\mathbb F}_2$ is
$2^{1+4}_- \cdot \mathsf {S}_5$ .
-
(ii) Each of the local systems
,
,
,
, and
${\mathcal H}_{95}:={\mathcal H} yp(\mathsf {Char}^\times _9,\mathsf {Char}^\times _5)$ has geometric monodromy group
$2^{1+6}_- \cdot \Omega ^-_6(2)$ , which is also the arithmetic monodromy group over any finite extensions of
${\mathbb F}_4$ . For each of them, the arithmetic monodromy group over
${\mathbb F}_2$ is
$2^{1+6}_- \cdot \mathrm {O}^-_6(2)$ .
-
(iii) The local system
has geometric monodromy group
$2^{1+6}_- \cdot \mathrm {SU}_3(2)$ , which is also the arithmetic monodromy group over any finite extensions of
${\mathbb F}_4$ . Over
${\mathbb F}_2$ , the arithmetic monodromy group is
$2^{1+6}_- \cdot \mathrm {SU}_3(2) \cdot 2$ .
Furthermore, all the local systems considered in this theorem have
$M_{2,2}=3$
.
Proof. (a) First, we note that both
${\mathcal H}_{95}$
and
${\mathcal H}_{53}$
satisfy
$\mathrm {(\mathbf {S+})}$
by [Reference Katz and TiepKT3, Theorem 3.13]. Furthermore, each of
${\mathcal F}_{531}$
,
${\mathcal F}_{9531}$
,
${\mathcal F}_{951}$
,
${\mathcal F}_{931}$
has
$M_{2,2}= 3$
by Corollary 3.6. We also use the facts that if
$\varphi $
denotes the character of the underlying representation for the arithmetic monodromy group
$G_{\mathrm {arith},{\mathbb F}_2}$
of any of the listed sheaves over
${\mathbb F}_2$
, then
$\varphi $
is irreducible of symplectic type; in particular,
${\mathbf Z}(G_{\mathrm {arith},{\mathbb F}_2}) \leq C_2$
. (Indeed,
$\varphi $
is visibly real-valued, and its restriction to
$G_{\mathrm {geom}}$
of
${\mathcal H}_{95}$
, respectively
${\mathcal H}_{53}$
, is symplectically self-dual by [Reference KatzKa2, 8.8.1-2].) Furthermore, the restriction of
$\varphi $
to the arithmetic monodromy group
$G_{\mathrm {arith},{\mathbb F}_4}$
of any of the listed sheaves over
${\mathbb F}_4$
is rational-valued by [Reference Katz and TiepKT6, Theorem 7.1.2].
(b) Let
$\tilde {\mathcal F}$
denote any of the systems
${\mathcal F}_{9531}$
,
${\mathcal F}_{951}$
,
${\mathcal F}_{953}$
, and let
$\tilde {G}$
denote its geometric monodromy group. By the above,
$\tilde {G}$
is a finite irreducible subgroup of
$\mathrm {Sp}_8({\mathbb C})$
with
$M_{2,2}=3$
. Now we can apply [Reference Guralnick and TiepGT2, Theorem 1.5] to
$\tilde {G}$
, and note that case (B) cannot occur because the dimension
$D=8$
, whereas case (D) cannot occur because
$\varphi $
is of symplectic type. It follows that we are in case (C) of [Reference Guralnick and TiepGT2, Theorem 1.5]:

and
$\tilde {G}/E \leq \mathrm {O}^-_6(2)$
acts transitively on
$27$
(nonzero) singular vectors and on
$36$
nonsingular vectors of the natural module
${\mathbb F}_2^6$
for
$\mathrm {O}^-_6(2)$
. In particular,
$27$
divides
$|\tilde {G}/E|$
. In fact, the observations in (a) imply that (4.4.1) also holds for
$G_{\mathrm {arith},{\mathbb F}_2,\tilde {\mathcal F}}$
, the arithmetic monodromy group of
$\tilde {\mathcal F}$
over
${\mathbb F}_2$
.
Next, observe that a pullback of
$\tilde {\mathcal F}$
yields
${\mathcal F}_{95}$
, which is a Kummer pullback of
${\mathcal H}_{95}$
. In particular, if G denotes the geometric monodromy group of
${\mathcal F}_{95}$
and H denotes that of
${\mathcal H}_{95}$
, then
$G \lhd H$
,
$H/G \hookrightarrow C_9$
, and
$G \hookrightarrow \tilde {G}$
. Clearly,
$5$
divides
$|H|$
, so it also divides
$|\tilde {G}|$
and
$|\tilde {G}/E|$
. Thus,
$27 \cdot 5$
divides
$|\tilde {G}/E|$
. Using the list of maximal subgroups of
$\mathrm {O}^-_6(2)$
[Reference Conway, Curtis, Norton, Parker and WilsonAtlas], we deduce that
$\tilde {G}/E$
is either
$\Omega ^-_6(2)$
or
$\mathrm {O}^-_6(2)$
. In the latter case, [Reference Katz and TiepKT6, Proposition 8.2.4] implies, however, that
$|\varphi (g)| = \sqrt {2}$
for some
$g \in \tilde {G}$
, which is impossible by (a). Hence, we conclude that
$\tilde {G} = E \cdot \Omega ^-_6(2)$
, and the same holds for
$G_{\mathrm {arith},{\mathbb F}_4,\tilde {\mathcal F}}$
. However, the Frobenius at
$(1,0,\ldots ,0)$
over
${\mathbb F}_2$
(where
$1$
is the coefficient for
$x^5$
) has trace
$-2/\sqrt {2}$
and hence does not belong to
$G_{\mathrm {arith},{\mathbb F}_4,\tilde {\mathcal F}}$
. Together with (4.4.1), this implies that
$G_{\mathrm {arith},{\mathbb F}_2,\tilde {\mathcal F}}=E \cdot \mathrm {O}^-_6(2)$
.
(c) To identify H, the
$G_{\mathrm {geom}}$
for
${\mathcal H}_{95}$
, we recall that H satisfies
$\mathrm {(\mathbf {S+})}$
by (a). First, suppose that H is an extraspecial normalizer. Together with (a), this implies that

We already mentioned that each of
$C_9$
and
$C_5$
injects in H, hence also in
$H/E_1 \leq \mathrm {O}^-_6(2)$
. Again using the list of maximal subgroups of
$\Omega ^-_6(2)$
[Reference Conway, Curtis, Norton, Parker and WilsonAtlas], we deduce that
$H/E_1$
is either
$\Omega ^-_6(2)$
or
$\mathrm {O}^-_6(2)$
. In the latter case, [Reference Katz and TiepKT6, Proposition 8.2.4] implies, however, that
$|\varphi (g)| = \sqrt {2}$
for some
$g \in H$
, which is impossible by (a). Hence, we conclude that
$H = E_1 \cdot \Omega ^-_6(2)$
(in fact, the same holds for
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal H}_{95}}$
because it normalizes
${\mathbf O}_2(H)=E_1$
and hence also satisfies (4.4.2)). In particular, H is perfect. Since
$H/G \hookrightarrow C_9$
, we also have
$G=H$
. Knowing now that

we conclude that
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{95}}=G$
. Next, again using the Frobenius at
$s=1$
of
${\mathcal F}_{95}$
with trace
$-\sqrt {2}$
, we see that this Frobenius is in
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{951}}$
but not in its subgroup G of index
$2$
. This shows that
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{95}} = G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{951}}=E \cdot \mathrm {O}^-_6(2)$
. As
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{95}}$
is a subgroup of
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal H}_{95}}$
, which normalizes
${\mathbf O}_2(H)=E_1$
and hence satisfies (4.4.2), we deduce that
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal H}_{95}}=2^{1+6}_- \cdot \mathrm {O}^-_6(2)$
.
Assume now that H is almost quasisimple, with R the unique non-abelian composition factor. Then
$G^{(\infty )} = H^{(\infty )}$
is a cover of R with center

cf. (4.4.1), (4.4.2). However,
$E \cap G^{(\infty )}$
is a normal
$2$
-subgroup of
$G^{(\infty )}$
, so

We also know from
$G \leq \tilde {G} = E\cdot S$
that

It follows that R is a simple subquotient of
$\mathrm {SU}_4(2)$
. Using [Reference Conway, Curtis, Norton, Parker and WilsonAtlas], we readily see that
$R=\mathsf {A}_5$
,
$\mathsf {A}_6$
, or
$\mathrm {SU}_4(2)$
; in particular,
$\mathrm {Out}(R)$
is a
$2$
-group. Recalling that

we have that
$C_9 \hookrightarrow R$
. This rules out the possibilities
$\mathsf {A}_5$
and
$\mathsf {A}_6$
, and so
$R = \mathrm {SU}_4(2)$
. But H acts irreducibly on
${\mathcal H}_{95}$
of dimension
$8$
, so we must have that
$H \cong \mathrm {Sp}_4(3) \cdot 2$
. This is, however, impossible because
$H={\mathbf O}^2(H)$
.
(d) In dimension
$8$
, it remains to determine
$G_1$
, the
$G_{\mathrm {geom}}$
for
${\mathcal F}_{931}$
, which also has
$M_{2,2}=3$
. As in the case of
$\tilde {G}$
, this equality implies by [Reference Guralnick and TiepGT2, Theorem 1.5] that

Moreover,
$G_1/E_2 \leq \mathrm {O}^-_6(2)$
still acts acts transitively on
$27$
(nonzero) singular vectors and on
$36$
non-singular vectors of the natural module
$W={\mathbb F}_2^6$
for
$\mathrm {O}^-_6(2)$
; in particular,
$27$
divides
$|G_1/E_2|$
. Using the list of maximal subgroups of
$\mathrm {O}^-_6(2)$
[Reference Conway, Curtis, Norton, Parker and WilsonAtlas], we deduce that
$G_1/E_2$
is either
$\mathrm {O}^-_6(2)$
,
$\Omega ^-_6(2)$
, a subgroup of
$M:=\mathrm {O}^-_2(2) \wr \mathsf {S}_3$
, or a subgroup of
$N:=\mathrm {GU}_3(2) \cdot 2 \cong 3^{1+2}_+ \rtimes 2\mathsf {S}_4$
. The first case is impossible since
$|G_1| \leq |\tilde {G}| = |E_2| \cdot |\Omega ^-_6(2)|$
. To rule out the second possibility, we make use of [Reference Katz and TiepKT6, Corollary 7.1.5], which shows that

for any odd-order element
$x \in G_1$
. Indeed, in this case, we have
$G_1 = \tilde {G}$
since
$|G_1|=|\tilde {G}|$
and
$G_1 \leq \tilde {G}$
; in particular,
$G_1$
contains an element
$g_1$
of order
$5$
which has rational trace. The latter condition implies that
$\varphi (g_1) \in \{-2,3\}$
, violating (4.4.4). In the third possibility, we can realize M as the stabilizer of the decomposition

where the quadratic form Q on W takes value

at the vector
$\sum ^6_{i=1}x_ie_i$
. But then the vectors
$u:=e_1+e_2$
and
$v:=\sum ^6_{i=1}e_i$
have
$Q(u)=Q(v)=1$
and belong to different M-orbits, showing that M is not transitive on the non-singular vectors of W. This leaves only the fourth possibility:
$G_1/E_2 \leq N$
. In particular,
$G_1$
is solvable.
Now we use the embedding
$G_1 \hookrightarrow \tilde {G}=G_{\mathrm {geom},{\mathcal F}_{9531}} = E \cdot \Omega ^-_6(2)$
. Then

showing that
$M_{2,2}(EG_1)=3$
. Thus,
$EG_1/E$
is a solvable subgroup of
$\Omega ^-_6(2)$
which acts transitively on
$27$
singular vectors and on
$36$
non-singular vectors of
${\mathbb F}_2^6$
. Using the list of maximal subgroups of
$\Omega ^-_6(2)$
as in the preceding paragraph, we see that
$EG_1/E$
is contained in
$N_1 \cong \mathrm {GU}_3(2)$
. Recalling E is a
$2$
-group and
$G_1={\mathbf O}^{2'}(G_1)$
(as the
$G_{\mathrm {geom}}$
for a local system on
${\mathbb A}^2/\overline {{\mathbb F}_2}$
), we then have

Moreover,
$27$
and
$36$
both divide
$|G_1/E_2| = |G_1|/|E|$
, so in fact, we have

Suppose that
$EG_1/E = 3^{1+2}_+ \rtimes C_4$
in (4.4.5). Note that we can turn the quadratic space
$W = {\mathbb F}_2^6$
into the Hermitian space
$W_1:={\mathbb F}_4^3$
for
$\mathrm {SU}_3(2)$
in such a way that the set
$N(W)$
of
$36$
non-singular vectors of W is exactly the set
$N(W_1)$
of
$36$
non-singular vectors of
$W_1$
. Since
$EG_1/E$
acts transitively on
$N(W)=N(W_1)$
, the stabilizer of any
$w \in N(W_1)$
has order
$3$
, which implies that a fixed involution j in
$EG_1/E$
does not fix any
$w \in N(W_1)$
. The Sylow
$2$
-subgroups of
$\mathrm {SU}_3(2)$
are isomorphic to
$Q_8$
, so any involution in
$\mathrm {SU}_3(2)$
is conjugate to j and hence does not fix any
$w \in N(W_1)$
. But this is a contradiction, since the stabilizer of any
$w \in N(W_1)$
in
$\mathrm {SU}_3(2)$
is
$\mathrm {SU}_2(2) \cong \mathsf {S}_3$
, which clearly contains an involution. We have therefore shown that

Recall that a pullback of
${\mathcal F}_{931}$
contains the Pink-Sawin system
which has
$2^{1+6}_-$
as its
$G_{\mathrm {geom}}$
by [Reference Katz and TiepKT6, Theorem 7.3.8]. This implies that
$G_1$
contains
${\mathbf Z}(E) \cong C_2$
. As
$[E,E] = {\mathbf Z}(E)$
, in the conjugation action on
$E/{\mathbf Z}(E) \cong {\mathbb F}_2^6$
the subgroup E acts trivially, whereas
$\mathrm {SU}_3(2)$
acts irreducibly (indeed, no proper parabolic subgroup of
$\mathrm {GL}_6(2)$
can contain
$3^{1+2}_+ = {\mathbf O}_3(\mathrm {SU}_3(2))$
as a subgroup). So (4.4.6) shows that
$G_1$
acts irreducibly on
$E/{\mathbf Z}(E)$
. It follows that
$E \cap G_1 = {\mathbf Z}(E)$
or E. In the former case, (4.4.6) implies that
$|G_1| = 2|\mathrm {SU}_3(2)|=2^4\cdot 3^3$
, which is impossible since
$G_1$
contains
$E_2$
of order
$2^7$
. We conclude that
$G_1 \lhd E$
, and

by (4.4.6).
To identify the group
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{931}}$
, we note that
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{931}} = \langle G_1,g_1 \rangle $
, where
$g_1 = \mathrm {Frob}_{(0,1),{\mathbb F}_2}$
. The pullback
$s=0$
of
${\mathcal F}_{931}$
is the Pink-Sawin system
, so by [Reference Katz and TiepKT6, Theorem 7.3.8],
$g_1^2$
is contained in its
$G_{\mathrm {geom}}$
, which is contained in
$G_1$
. Moreover,
$g_1$
has trace
$-\sqrt {2}$
, showing
$g_1 \notin G_1$
. Thus
$G_1$
has
$2$
in
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{931}}$
, whence we also have
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{931}}=G_1$
.
(e) Now we work in dimension
$4$
. Let
$\tilde {G}$
denote the geometric monodromy group of
${\mathcal F}_{531}$
. Since it has
$M_{2,2}=3$
and is of symplectic type, the restriction of
$\mathrm {Sym}^2(\varphi )$
to
$\tilde {G}$
is irreducible, whence
$\tilde {G}$
satisfies
$\mathrm {(\mathbf {S+})}$
by [Reference Guralnick and TiepGT3, Lemma 2.1].
First, we consider the case where
$\tilde {G}$
is almost quasisimple. Then
$\tilde {G}^{(\infty )}$
is a quasisimple irreducible subgroup of
$\mathrm {Sp}_4({\mathbb C})$
. Using [Reference Hiss and MalleHM], we then deduce that
$\tilde {G}^{(\infty )}$
is
$2\cdot \mathsf {A}_5$
or
$2 \cdot \mathsf {A}_6$
. Potentially
$\tilde {G}$
could still have index
$2$
over
$\tilde {G}^{(\infty )}$
. But using the rationality of the restriction of
$\varphi $
to
$\tilde {G}$
, we get
$\tilde {G}=\tilde {G}^{(\infty )} \leq 2 \cdot \mathsf {A}_6$
. However, a pullback of
${\mathcal F}_{531}$
is the Pink-Sawin system
which has
$2^{1+4}_-$
as its
$G_{\mathrm {geom}}$
by [Reference Katz and TiepKT6, Theorem 7.3.8]. This yields a contradiction, since
$2^{1+4}_-$
cannot embed in
$2 \cdot \mathsf {A}_6$
.
We have therefore shown that
$\tilde {G}$
is an extraspecial normalizer, and so

note that
$\mathrm {O}^-_4(2) \cong \mathsf {S}_5$
. Now let H denote the
$G_{\mathrm {geom}}$
for
${\mathcal H}_{53}$
and let G denote the
$G_{\mathrm {geom}}$
for
${\mathcal F}_{53}$
, so that
$H/G \hookrightarrow C_5$
. Recall from (a) that H satisfies
$\mathrm {(\mathbf {S+})}$
. Assume in addition that H is almost quasisimple. Then
$G^{(\infty )} = H^{(\infty )}$
is a cover of a non-abelian simple group R. But
$G \hookrightarrow \tilde {G}$
, so (4.4.7) implies that R is a simple subquotient of
$\mathsf {S}_5$
. It follows that
$R = \mathsf {A}_5$
. We also know that
$H \leq \mathrm {Sp}_4({\mathbb C})$
is almost quasisimple with rational traces. Hence,
$H = \mathrm {SL}_2(5)$
in a faithful irreducible representation of degree
$4$
; in particular, any element of order
$3$
in H has trace
$1$
[Reference Conway, Curtis, Norton, Parker and WilsonAtlas]. Thus, any element t of order
$3$
in
$I(\infty )$
has trace
$1$
in
$\varphi $
, and trace
$-1$
on the tame part of
${\mathcal H}_{53}$
. So t has trace
$2$
on the wild part
$\mathsf {Wild}$
of
${\mathcal H}_{53}$
, which means that t acts trivially on
$\mathsf {Wild}$
, a contradiction.
We have now shown that H is also an extraspecial normalizer, and so (4.4.7) also holds for H. Note that both
$C_5$
and
$C_3$
inject in H, so
$15$
divides the order of
$H/E \leq \mathsf {S}_5$
. Inspecting the list of maximal subgroups of
$\mathrm {Sym}_5$
[Reference Conway, Curtis, Norton, Parker and WilsonAtlas], we see that
$H/E = \mathsf {S}_5$
or
$\mathsf {A}_5$
. But
$H={\mathbf O}^2(H)$
, so
$H= E \cdot \mathsf {A}_5$
; in particular, H is perfect. Since
$H/G \hookrightarrow C_5$
, we also have that
$G=H$
. Now
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{53}}$
normalizes G and
${\mathbf O}_2(G)=E$
, so (4.4.7) holds for
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{53}}$
, which already contains the subgroup
$G = E \cdot \mathsf {A}_5$
of index
$2$
in
$E \cdot \mathsf {S}_5$
. By [Reference Katz and TiepKT6, Proposition 8.2.4],
$E \cdot \mathsf {S}_5$
contains an element x with
$|\varphi (x)| = \sqrt {2}$
. Since
$\varphi $
is rational on
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{53}}$
, we conclude that
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{53}} = G$
. Noting that (4.4.7) holds for
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal H}_{53}}$
which has only rational traces and contains
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{53}}$
, we deduce that
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal H}_{53}} = G$
.
Next,
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{53}}$
normalizes G and
${\mathbf O}_2(G)=E$
, so (4.4.7) holds for
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{53}}$
. But now
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{53}}$
contains the Frobenius at
$s=1$
with trace
$-\sqrt {2}$
that does not belong to G. Using (4.4.7), we conclude that
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{53}} = E \cdot \mathsf {S}_5$
. As
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{53}}$
embeds in
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal H}_{53}}$
which also satisfies (4.4.7) (as it normalizes
${\mathbf O}_2(H)=E$
), we must have that
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal H}_{53}}=E \cdot \mathsf {S}_5$
.
Now,
$E \cdot \mathsf {A}_5 = G = G_{\mathrm {geom},{\mathcal F}_{53}} \leq G_{\mathrm {geom},{\mathcal F}_{531}} = \tilde {G} \leq E \cdot \mathsf {S}_5$
and
$\varphi |_{\tilde {G}}$
is rational-valued, so
$\tilde {G}=G$
. Repeating the same inclusions for
$G_{\mathrm {arith},{\mathbb F}_4}$
, we get
$G_{\mathrm {arith},{\mathbb F}_4,{\mathcal F}_{531}}=G$
. Finally, as
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{531}}$
normalizes
${\mathbf O}_2(\tilde {G})=E$
, we have

whence
$G_{\mathrm {arith},{\mathbb F}_2,{\mathcal F}_{531}} = E \cdot \mathsf {S}_5$
.
(f) As mentioned in the proof of Theorem 4.3, we already know
$M_{2,2}=3$
unless
${\mathcal F}={\mathcal F}_{53}$
or
${\mathcal F}_{95}$
. But
$G_{\mathrm {geom},{\mathcal F}_{53}} = G_{\mathrm {geom},{\mathcal F}_{531}}$
according to (i) and
$G_{\mathrm {geom},{\mathcal F}_{95}} = G_{\mathrm {geom},{\mathcal F}_{9531}}$
, so
${\mathcal F}_{53}$
and
${\mathcal F}_{95}$
both have
$M_{2,2}=3$
as well.
5 Multiparameter local systems: Balanced pairs and Infmono
$(A,B)$
We will now develop some framework to study the case in which
$r \geq 2$
and
has infinite
$G_{\mathrm {geom}}$
. First, we attach to the data
$(A,B_1,\ldots ,B_r)$
a balanced pair
$(A, B=\mathrm {some\ }B_i)$
as follows. We must distinguish three cases.
-
(i) If A and all
$B_i$ are odd, we choose any of the
$B_i$ .
-
(ii) If A is even, then some
$B_i$ is odd, and we choose any odd
$B_i$ .
-
(iii) If A is odd and some
$B_i$ is even, then we choose some even
$B_i$ .
Notice that, in all cases, at least one of
$A,B$
is odd, and hence,
$\gcd (A,B)$
is odd.
We now formulate the following hypothesis infmono
$(A,B)$
for a pair
$(A,B)$
of integers
$A>B \ge 1$
with
$p \nmid AB$
. For
$C:=\gcd (A,B)$
,
$A=CA_0$
,
$B=CB_0$
, we have the direct sum decomposition

where, in general,
${\mathcal F}(A,B,\chi )$
is the local system on
${\mathbb A}^1/{\mathbb F}_p(\chi )$
whose trace function is given as follows: for
$L/{\mathbb F}_p(\chi )$
a finite extension and
$ t \in L$
,

The hypothesis infmono
$(A,B)$
is the following:

Lemma 5.1. For
$C:=\gcd (A,B), A=CA_0,\ B=CB_0$
, suppose C is odd and
$C \ge 3$
. Then the following statements hold.
-
(i) If
has infinite
$G_{\mathrm {geom}}$ , then infmono
$(A,B)$ holds.
-
(ii) Suppose
has finite
$G_{\mathrm {geom}}$ but some summand
${\mathcal F}(A_0,B_0,\chi )$ of
has infinite
$G_{\mathrm {geom}}$ . Then there is a divisor
$C_0$ of C with
$C_0 < C$ such that a summand
${\mathcal F}(A_0,B_0,\varphi )$ of
has infinite
$G_{\mathrm {geom}}$ precisely when
, i.e.
$\varphi \in \mathsf {Char}(C) \smallsetminus \mathsf {Char}(C_0)$ .
Proof. (i) follows from [Reference Katz and TiepKT6, 10.2.6 and 10.3.13]. Indeed, for
$\chi $
a nontrivial character of odd order, if
${\mathcal F}(A_0,B_0,\chi )$
has finite
$G_{\mathrm {geom}}$
, then
${\mathcal F}(A_0,B_0,\chi )$
is in case (ii) of either [Reference Katz and TiepKT6, 10.2.6(ii) or 10.3.13(ii)] (the
$\mathrm {SU}$
cases). In these cases,
also has finite
$G_{\mathrm {geom}}$
.
For (ii), let S denote the set of
$\chi \in \mathsf {Char}(C)$
for which
${\mathcal F}(A_0,B_0,\chi )$
has infinite
$G_{\mathrm {geom}}$
; in particular,
$\chi \in S$
but
. If
$\mathsf {Char}(C) \smallsetminus S$
is the singleton
, then the assertion holds trivially with
$C_0=1$
. If not, choose a nontrivial

Then
${\mathcal F}(A_0,B_0,\rho )$
has finite
$G_{\mathrm {geom}}$
, so by [Reference Katz and TiepKT6, 10.2.6 and 10.3.13], we have

and
$\mathsf {o}(\rho )|\gcd (q+1,C)$
for some power q of p and some odd integers
$n> m$
with
$\gcd (n,m)=1$
. Then
${\mathcal F}(A_0,B_0,\sigma )$
has finite
$G_{\mathrm {geom}}$
precisely for
$\sigma \in \mathsf {Char}(q+1)$
. Thus,
${\mathcal F}(A_0,B_0,\Lambda )$
with
$\Lambda \in \mathsf {Char}(C)$
has finite
$G_{\mathrm {geom}}$
precisely when
$\Lambda $
has order dividing
$C_0:=\gcd (C,q+1)$
. It remains to observe that
$C_0 < C$
, for otherwise
$C|(q+1)$
and every
$\rho \in \mathsf {Char}(C)$
would have
${\mathcal F}(A_0,B_0,\rho )$
with finite
$G_{\mathrm {geom}}$
, contradicting the hypothesis that some summand
${\mathcal F}(A_0,B_0,\chi )$
has infinite
$G_{\mathrm {geom}}$
.
The following statement is a consequence of Lemma 5.1(ii), but we will offer an independent proof.
Lemma 5.2. Suppose that
$\chi $
and
$\rho $
are nontrivial characters of odd order C which are Galois conjugate under
$\mathrm {Gal}({\mathbb Q}(\zeta _C)/{\mathbb Q})$
. Then
${\mathcal F}(A_0,B_0,\chi )$
has finite
$G_{\mathrm {geom}}$
if and only if
${\mathcal F}(A_0,B_0,\rho )$
has finite
$G_{\mathrm {geom}}$
.
Proof. The question is geometric, so we may work over extensions of
${\mathbb F}_{p^2}(\chi ,\rho )$
. Over a finite extension
$k/{\mathbb F}_{p^2}(\chi ,\rho )$
, all traces of
${\mathcal F}(A_0,B_0,\chi )$
and of
${\mathcal F}(A_0,B_0,\rho )$
lie in
${\mathbb Q}(\zeta _p,\zeta _C)$
, and point by point their traces are conjugate under the action of
$\mathrm {Gal}({\mathbb Q}(\zeta _C,\zeta _p)/{\mathbb Q}(\zeta _p))$
. In both cases, finiteness of
$G_{\mathrm {geom}}$
is equivalent to all traces being algebraic integers, a condition which is invariant under Galois conjugation.
Because
${\mathcal F}$
is geometrically irreducible and starts life over
${\mathbb F}_p$
, if
$G_{\mathrm {geom},{\mathcal F}}$
is infinite, then its identity component
$G^\circ $
is semisimple, by Grothendieck’s local monodromy theorem [Reference DeligneDe2, 1.3.9]. Next, we determine
$G_{\mathrm {geom},{\mathcal F}}$
in some ‘easy’ cases.
Theorem 5.3. Consider the local system subject to (4.0.1). Suppose that
has infinite geometric monodromy group H for some balanced pair
$(A,B=B_i)$
with
$\gcd (A,B)=1$
. Then we have the following results.
-
(i) If
$2 \nmid AB_i$ , then
$G_{\mathrm {geom},{\mathcal F}}=\mathrm {Sp}_{A-1}$ .
-
(ii) Otherwise,
$\mathrm {SL}_{A-1} \leq G_{\mathrm {geom},{\mathcal F}} \leq \{g \in \mathrm {GL}_{A-1}|\det (g)^p=1\}.$
Proof. Suppose first that
$2 \nmid AB_i$
. Then by [Reference Katz and TiepKT6, Theorems 10.2.4(iii) and 10.3.21(iii)],
$H=\mathrm {Sp}_{A-1}$
. By the definition of a balanced pair, the fact that
$2 \nmid AB_i$
implies that A and all
$B_j$
are odd. As
is a pullback of
${\mathcal F}$
, we have
$H \leq G_{\mathrm {geom},{\mathcal F}}$
. But we also have an a priori inclusion
$G_{\mathrm {geom},{\mathcal F}} \leq \mathrm {Sp}_{A-1}$
. Hence,
$G_{\mathrm {geom},{\mathcal F}}=\mathrm {Sp}_{A-1}$
in this case.
Suppose next that
$2|AB_i$
. Then by [Reference Katz and TiepKT6, Theorems 10.2.4(i) and 10.3.21(i)], we have

As is a pullback of
${\mathcal F}$
, we again have
$H \leq G_{\mathrm {geom},{\mathcal F}}$
, and so

is irreducible. By [Reference Katz and TiepKT6, 2.3.1], we have
$G_{\mathrm {geom},{\mathcal F}} \leq \{g \in \mathrm {GL}_{A-1}|\det (g)^p=1\}$
. [To apply the cited result, use the fact that the question is geometric, and after pullback to
${\mathbb A}^r/{\mathbb F}_{p^2}$
, all Frobenius traces of
${\mathcal F}$
lie in
${\mathbb Q}(\zeta _p)$
.]
Theorem 5.4. Consider the local system with
$r \geq 2$
subject to (4.0.1). Suppose
$G_{\mathrm {geom},{\mathcal F}}$
is infinite. Then we have the following results.
-
(i) Suppose that
$A=2B_i$ for some i. Then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
-
(ii) Suppose that
$A=3B_i$ for some i. Then
$G_{\mathrm {geom},{\mathcal F}} =\mathrm {Sp}_{A-1}$ if
$2 \nmid AB_1 \ldots B_r$ , and
$G_{\mathrm {geom},{\mathcal F}}^\circ = \mathrm {SL}_{A-1}$ if and
$2|AB_1 \ldots B_r$ .
Proof. Both assertions result from Theorem 3.7. In (i), by Theorem 3.7, we have
$M_{2,2}({\mathcal F})=2$
. This in turn implies by [Reference Guralnick and TiepGT2, Theorem 1.5] that
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$
.
In (ii), by Theorem 3.7, we have
$M_{2,2}({\mathcal F})=3$
if
$2 \nmid AB_1 \ldots B_r$
, and
$M_{2,2}({\mathcal F})=2$
if
$2|AB_1 \ldots B_r$
. In the former case, we also have an a priori inclusion
$G_{\mathrm {geom},{\mathcal F}} \leq \mathrm {Sp}_{A-1}$
. This in turn implies by [Reference Guralnick and TiepGT2, Theorem 1.5] that
$G_{\mathrm {geom},{\mathcal F}} =\mathrm {Sp}_{A-1}$
. In the latter case, we have
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$
as in (i).
To work with pairs
$(A,B)$
with
$C = \gcd (A,B)> 1$
, we first observe the following:
Lemma 5.5. Let
$A_0> B_0$
be prime to p integers with
$\gcd (A_0,B_0)=1 $
, and
$\chi \neq \varphi $
two multiplicative characters. We have the following results.
-
(i) In all cases,
${\mathcal F}(A_0,B_0,\chi )$ is not geometrically isomorphic to
${\mathcal F}(A_0,B_0,\varphi )$ .
-
(ii) If
$A_0B_0$ is even, then
${\mathcal F}(A_0,B_0,\chi )$ is not geometrically isomorphic to
${\mathcal F}(A_0,B_0,\varphi )^\vee $ .
-
(iii) If
$A_0B_0$ is even, then
${\mathcal F}(A_0,B_0,\chi )$ is not geometrically isomorphic to
${\mathcal F}(A_0,B_0,\chi )^\vee $ .
-
(iv) If
$A_0B_0$ is odd, the dual of
${\mathcal F}(A_0,B_0,\chi )$ is
${\mathcal F}(A_0,B_0,\overline {\chi })$ .
-
(v) If
$A_0B_0$ is odd,
${\mathcal F}(A_0,B_0,\chi )$ is not geometrically isomorphic to
${\mathcal F}(A_0,B_0,\varphi )$ . It is isomorphic to the dual of
${\mathcal F}(A_0,B_0,\varphi )$ only for
$\varphi =\overline {\chi }$ .
Proof. We first prove that
${\mathcal F}(A_0,B_0,\chi )$
is not geometrically isomorphic to
${\mathcal F}(A_0,B_0,\varphi )$
(i.e., that
$H^2_c({\mathbb G}_m/\overline {{\mathbb F}_p},{\mathcal F}(A_0,B_0,\chi )\otimes {\mathcal F}(A_0,B_0,\varphi )^\vee )=0$
). The dimension of this
$H^2_c$
is the limsup, over finite extensions L of
${\mathbb F}_p(\chi ,\varphi )$
, of the sums

The inner sum for
$\zeta \neq 1$
has
$\zeta ^{A_0}\neq 1$
(because
$\gcd (A_0,B_0)=1$
), so this inner sum is bounded in absolute value by
$A_0\sqrt {\# L}$
(Weil bound). For
$\zeta =1$
, the inner sum is
$\sum _{x \in L^\times }\chi (x)\varphi (1/x)$
, which vanishes unless
, in which case the inner sum is
$\# L-1$
,
We next prove that if
$A_0B_0$
is even, then
${\mathcal F}(A_0,B_0,\chi )$
is not geometrically isomorphic to the dual
${\mathcal F}(A_0,B_0,\varphi )^\vee $
. This amounts to the vanishing of the space
$H^2_c({\mathbb G}_m/\overline {{\mathbb F}_p},{\mathcal F}(A_0,B_0,\chi )\otimes {\mathcal F}(A_0,B_0,\varphi ))$
. The dimension of this
$H^2_c$
is the limsup, over finite extensions L of
${\mathbb F}_p(\chi ,\varphi )$
, of the sums

Choose a root of unity
$\tau $
with
$\tau ^{B_0}=-1$
. Then this sum is

For every
$\zeta \in \mu _{B_0}$
, we claim that
$(\tau \zeta )^{A_0} \neq -1$
. Indeed, if
$(\tau \zeta )^{A_0}=-1$
, then
$(\tau \zeta )^{A_0B_0}=(-1)^{B_0}$
, but
$(\tau \zeta )^{A_0B_0}= (-1)^{A_0}(\zeta )^{A_0B_0}= (-1)^{A_0}$
, and hence,
$ (-1)^{A_0}=(-1)^{B_0}$
, impossible as
$A_0$
and
$B_0$
have opposite parities in the
$A_0B_0$
even case. Therefore, each inner sum is bounded in absolute value by
$A_0\sqrt {\# L}$
(Weil bound), and we are done in this
$A_0B_0$
even case.
The proof of (ii-bis) is identical: the particular,
$\chi ,\varphi $
play no role in the proof of (ii).
Assertion (iii) is obvious: the trace functions of
${\mathcal F}(A_0,B_0,\chi )$
and
${\mathcal F}(A_0,B_0,\overline {\chi })$
are complex conjugates of each other if
$A_0B_0$
is odd. Assertion (iv) then follows from (i) and (iii).
In view of assertion (iii) of Lemma 5.5, in the case when
$AB$
is odd, for
$C = \gcd (A,B)$
, we choose a set
$\mathrm {Rep}(C) \subset \mathsf {Char}(C)$
of
$(C-1)/2$
nontrivial characters such that for each nontrivial
$\chi \in \mathsf {Char}(C)$
, precisely one of
$\chi ,\overline {\chi }$
lies in
$\mathrm {Rep}(C)$
.
Theorem 5.6. Let
$A> B \ge 1$
be prime to p integers with
$2 \nmid \gcd (A,B)=C>1$
. Suppose that infmono
$(A,B)$
holds, and write
$(A,B)=(CA_0,CB_0)$
. Then we have the following results.
-
(i) Suppose that
$AB$ is even and
$A_0>2$ . Then
-
(ii) Suppose that
$AB$ is odd and
$A_0>3$ . Then precisely one of
$\chi ,\overline {\chi }$ lies in
$\mathrm {Rep}(C)$ , and
.
Proof. We begin with the direct sum decomposition

Recall from [Reference Katz and TiepKT4, 3.10 (i) and (ii)] that, up to the same multiplicative translation, the local systems
${\mathcal F}(A_0,B_0,\chi )$
are each geometrically isomorphic to Kummer
$[A_0]^\star $
pullbacks of hypergeometric sheaves. We have

and for and any choice of
$\rho _\chi $
with
$\rho _\chi ^{A_0}=\chi $
, we have

(i) Suppose first that
$AB$
is even, and
$A_0 \ge 3$
. By infmono
$(A,B)$
, each
$G_{\mathrm {geom},{\mathcal F}(A_0,B_0,\chi )}$
is infinite. Then by [Reference Katz and TiepKT6, 10.2.4 and 10.3.21], we have

and for each , we have

Now consider the direct sum of hypergeometric sheaves defined as

Up to multiplicative translation, we have a geometric isomorphism

As finite pullback does not change
$G_{\mathrm {geom}}^\circ $
, we have

and for each , we have

In this
$AB$
even case, it suffices to show that
is the asserted product

For this, we apply Goursat-Kolchin-Ribet [Reference KatzKa2, 1.8.2]. We must show that for any character
${\mathcal L}$
of
,
-
(a) there is no isomorphism between
${\mathcal L}\otimes {\mathcal H}_{small,A_0,B_0}$ and any
${\mathcal H}_{big,A_0,B_0,\rho _\chi }$ or its dual
${\mathcal H}_{big,A_0,B_0,\rho _\chi }^\vee $
-
(b) For
$\chi \neq \varphi $ both nontrivial, there is no isomorphism between
${\mathcal L}\otimes {\mathcal H}_{big,A_0,B_0,\rho _\chi }$ and either
$ {\mathcal H}_{big,A_0,B_0,\rho _\varphi }$ or its dual
$ {\mathcal H}_{big,A_0,B_0,\rho _\varphi }^\vee $ .
The first condition holds trivially, as the ranks are different,
$A_0-1$
versus
$A_0$
. It suffices to show the second condition with the stronger statement for
${\mathcal L}$
any character of
$\pi _1^{\mathrm {geom},\mathrm {tame\ at\ }0}({\mathbb G}_m/{\mathbb F}_p)$
. Such a character is a Kummer sheaf
${\mathcal L}_\sigma $
. Indeed, as
$A_0 \ge 3$
, either
$A_0-B_0>1$
, in which case all
$\infty $
-slopes are
$<1$
, and so
${\mathcal L}$
is tame at
$\infty $
, or
$A_0-B_0=1$
, in which case there is a single slope
$1$
at
$\infty $
, but
$A_0-1 \ge 2$
slopes
$0$
at
$\infty $
, so again
${\mathcal L}$
must be tame at
$\infty $
.
As the ‘upstairs’ characters of
$ {\mathcal H}_{big,A_0,B_0,\rho _\chi }$
and of both
$ {\mathcal H}_{big,A_0,B_0,\rho _\varphi }$
and its dual
$ {\mathcal H}_{big,A_0,B_0,\rho _\varphi }^\vee $
are
$\mathsf {Char}(A_0)$
, the set of all characters, the only possible
${\mathcal L}$
is a Kummer
${\mathcal L}_\chi $
for some
$\chi \in \mathsf {Char}(A_0)$
.
If there were such an isomorphism, it would persist after
$[A_0]^\star $
Kummer pullback, which makes the
${\mathcal L}$
disappear. So in this
$AB$
even case, we are reduced to showing that for
$\chi \neq \varphi $
both nontrivial,
${\mathcal F}(A_0,B_0,\chi )$
is not geometrically isomorphic to either
${\mathcal F}(A_0,B_0,\varphi )$
or its dual
${\mathcal F}(A_0,B_0,\varphi )^\vee $
. Applying Lemma 5.5, we complete the proof in the
$AB$
even case.
(ii) We now treat the case when
$AB$
is odd. Then
$A_0B_0$
is odd, and for each nontrivial
$\chi \in \mathsf {Char}(C)$
, the two local systems
${\mathcal F}(A_0,B_0,\chi )$
and
${\mathcal F}(A_0,B_0,\overline {\chi )}$
are dual. Therefore,
has the same
$G_{\mathrm {geom}}$
as the ‘reduced’ direct sum

Let us explain this last point. Our situation is that we have two local systems
${\mathcal A}$
and
${\mathcal B}$
of ranks M and N, respectively. We consider both the direct sum
${\mathcal A} \oplus {\mathcal B} $
and the direct sum
${\mathcal A}\oplus {\mathcal B} \oplus {\mathcal B}^\vee $
. For the latter, an element
$\gamma \in \pi _1({\mathbb A}_1/\overline {{\mathbb F}_p})$
maps to a ‘diagonal’ element
$\mathrm {diag}(Z,X,Y)$
in
$\mathrm {GL}({\mathcal A}\oplus {\mathcal B}\oplus {\mathcal B}^\vee )$
, This element satisfies the matrix equation
${}^t\! XY=\mathrm {Id}_{2N}$
. Hence, every element
$(Z,X,Y)$
in the Zariski closure also satisfies the matrix equation
${}^t\! XY=\mathrm {Id}_{2N}$
. Thus, the map

is injective: we recover Y as
${}^t\! X^{-1}$
. But this projection is surjective, so we get the asserted isomorphism

Analogously to the
$AB$
even case, we introduce the ‘reduced’ direct sum of hypergeometric sheaves

whose
$[A_0]^\star $
Kummer pullback is
. Then it suffices to prove that
is the asserted product

In view of Lemma 5.5, this is immediate from Goursat-Kolchin-Ribet [Reference KatzKa2, 1.8.2]. Indeed, with the hypothesis
$A_0>3$
, we can instead directly apply [Reference KatzKa2, 8.11.7.2] because the exclusion (1) of that result, concerning factors of rank
$2$
, is vacuous, as there are no such factors.
6 Multiparameter local systems with infinite monodromy. I
We continue to work with local systems defined in (4.0.1), for which the condition infmono
$(A,B)$
does not necessarily hold. First, we give a slight variant of Theorem 5.6.
Theorem 6.1. Given prime to p integers
$A> B \ge 1$
, suppose that
$C:=\gcd (A,B)$
is both odd and
$\ge 3$
. Write
$(A,B)=(CA_0,CB_0)$
. Let
$S_{\mathrm {inf}}$
, respectively
$S_{\mathrm {fin}}$
, be the set of those characters
$\chi \in \mathsf {Char}(C)$
for which
${\mathcal F}(A_0,B_0,\chi )$
has infinite, respectively finite,
$G_{\mathrm {geom}}$
. Suppose that
$\varnothing \neq S_{\mathrm {inf}} \neq \mathsf {Char}(C)$
. Then, by Lemma 5.1(ii),

for some divisor
$C_0 < C$
of C. Denote

Then the following statements hold for and
$H_{\mathrm {fin}}:= G_{\mathrm {geom},{\mathcal F}_{\mathrm {fin}}(A,B)}$
.
-
(i) Suppose
$A_0B_0$ is even. Then
$A_0-1 \neq B_0$ , then
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}=G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}^\circ $ and
$$ \begin{align*}G = H_{\mathrm{fin}} \times G_{\mathrm{geom}, {\mathcal F}_{\mathrm{inf}}(A,B)}. \end{align*} $$
$A_0-1=B_0$ but
$A_0> 2$ , then
$$ \begin{align*}[G,G] = [H_{\mathrm{fin}},H_{\mathrm{fin}}] \times G_{\mathrm{geom}, {\mathcal F}_{\mathrm{inf}}(A,B)}^\circ,\end{align*} $$
$G/[G,G]$ is a nontrivial finite elementary abelian p-group.
-
(ii) Suppose
$A_0B_0$ is odd. Choose a subset
$\mathrm {Rep}(S_{\mathrm {inf}}) \subset S_{\mathrm {inf}}$ of
$\#S_{\mathrm {inf}}/2$ nontrivial characters such that for each nontrivial
$\chi \in S_{\mathrm {inf}}$ , precisely one of
$\chi ,\overline {\chi }$ lies in
$\mathrm {Rep}(S_{\mathrm {inf}})$ . Then
$$ \begin{align*}G_{\mathrm{geom}, {\mathcal F}_{\mathrm{inf}}(A,B)} \cong \prod_{\chi \in \mathrm{Rep}(S_{\mathrm{inf}})} \mathrm{SL}_{A_0}, ~G = H_{\mathrm{fin}} \times G_{\mathrm{geom}, {\mathcal F}_{\mathrm{inf}}(A,B)}.\end{align*} $$
Proof. The proof of the identification of , via Goursat-Kolchin-Ribet, is a subset of the proof of Theorem 5.6 and is left to the reader.
Next, observe that
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}^\circ $
is perfect and has no nontrivial finite quotient; furthermore,

Assume in addition that either
$2 \nmid A_0B_0$
, or
$2\mid A_0B_0$
but
$A_0-1 \neq B_0$
. Then (5.3.1) and the arguments in the proof of Theorem 5.6 show that

Now, the action of G on the two summands in (6.1.1) projects G onto the finite group
$H_{\mathrm {fin}}$
and onto
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}$
. Since
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}$
has no finite quotient, an application of the classical Goursat lemma (cf. [Reference LangL, Exercise 5, p. 75]) shows that

Assume now that
$A_0-1=B_0$
but
$A_0> 2$
. By [Reference Katz and TiepKT6, Theorem 10.3.13], we have
$C_0=1$
, and precisely two cases:

The action of G on any summand
${\mathcal F}(A_0,B_0,\chi )$
of
projects G onto
$H_{\mathrm {fin}}$
when
and onto an intermediate group between
$\mathrm {SL}_{A_0}$
and
$\mathrm {SL}_{A_0} \cdot p$
when
$\chi \in S_{\mathrm {inf}}$
. Hence, it projects
$[G,G]$
onto the quasisimple group
$[H_{\mathrm {fin}},H_{\mathrm {fin}}]$
when
, and onto
$\mathrm {SL}_{A_0}$
when
$\chi \in S_{\mathrm {inf}}$
. Again using the classical Goursat lemma, we conclude that

Now, the above action projects
$G/[G,G]$
onto
$C_p$
on every finite summand, and onto
$1$
or
$C_p$
on
$\mathrm {SL}$
-type components, whence
$G/[G,G]$
is a nontrivial finite elementary abelian p-group.
This last result allows a partial strengthening of Lemma 5.5(i).
Corollary 6.2. Hypotheses and notations as in Theorem 6.1, the local systems
${\mathcal F}(A_0,B_0,\chi )$
with
$\chi \in S_{\mathrm {inf}}$
are pairwise non-isomorphic as representations of
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}^\circ $
.
Proof. In the case when
$A_0B_0$
is even, the group
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}^\circ $
is a product of nontrivial groups, and the constituents
${\mathcal F}(A_0,B_0,\chi )$
indexed by
$\chi \in S_{\mathrm {inf}}$
are nontrivial irreducible representations of the various nontrivial factor groups. In the case when
$A_0B_0$
is odd, the group
$G_{\mathrm {geom}, {\mathcal F}_{\mathrm {inf}}(A,B)}$
is a product of copies of
$\mathrm {SL}_{A_0}$
,
$A_0 \ge 3$
, and the constituents are either the natural module for one of the
$\mathrm {SL}_{A_0}$
factors or the dual of the natural module for that factor. In this case, it remains to observe that, because
$A_0>2$
, the natural module for a given
$\mathrm {SL}_{A_0}$
factor is not self-dual.
Theorem 6.3. Consider the local system with
$r \ge 2$
, subject to (4.0.1). Suppose that, for some (not necessarily balanced) pair
$(A,B:=B_i)$
with
$2 \nmid C=\gcd (A,B)$
,
has infinite geometric monodromy group. Then
${\mathcal F}$
is Lie-irreducible (i.e.,
$G_{\mathrm {geom},{\mathcal F}}^\circ $
acts irreducibly).
Proof. Write
$(A,B)=(CA_0,CB_0)$
. When
$C=1$
, or when
$A_0=2$
, or when
$A_0=3$
and B is odd, we have already established the statement in Theorem 5.3 (and its proof), and in Theorem 5.4.
It remains to treat the case when
$C\geq 3$
and either
$A_0>2$
or both
$A_0B_0$
is odd and
$A_0>3$
. In these cases, we have the direct sum decomposition

into pairwise non-isomorphic geometrically irreducible constituents. Precisely one of these constituents has rank
$A_0-1$
(namely,
), and the other
$C-1$
constituents each have rank
$A_0$
.
(a) Let
$S_{\mathrm {inf}}$
, respectively
$S_{\mathrm {fin}}$
, be the set of those characters
$\chi \in \mathsf {Char}(C)$
for which
${\mathcal F}(A_0,B_0,\chi )$
has infinite, respectively finite,
$G_{\mathrm {geom}}$
, and write

Recall from Lemma 5.1 that either
$C_0=0$
, or
$C_0$
is a proper divisor of
$C \geq 3$
, and hence,

Because is a pullback of
${\mathcal F}$
, we have
$H \leq G$
for

and hence,
$H^\circ \leq G^\circ $
. Now we can apply Theorem 5.6 and Theorem 6.1 to see that

By Lemma 5.5(i) and Corollary 6.2, each of
$C-C_0 \geq 2$
constituents
${\mathcal F}(A_0,B_0,\chi )$
,
$\chi \in S_{\mathrm {inf}}$
, is irreducible under
$H^\circ $
, and they are pairwise non-isomorphic as representations of
$H^\circ $
.
(b) We argue by contradiction. We know [Reference Katz and TiepKT5, 2.6] that
${\mathcal F}$
is geometrically irreducible (i.e., that G is an irreducible subgroup of
$\mathrm {GL}_{A-1}=\mathrm {GL}(V)$
with
$V:={\mathcal F}_{\overline {\eta }}$
). Suppose that
$G^\circ $
is reducible on V. Because
$G^\circ \lhd G$
, the action of
$G^\circ $
on V is completely reducible. Let

be the decomposition of V into isotypical components under the action of
$G^\circ $
. Then G transitively permutes these m isotypical components, and it must also transitively permute the isomorphism classes of the
$W_i$
. Therefore, the multiplicities
$n_i$
have a common value n, and
$V =n(\oplus ^m_{i=1} W_i)$
under
$G^\circ $
. Now if
$n> 1$
, then some simple summand of
$V|_{H^\circ }$
has multiplicity
$\geq n$
, contradicting the conclusion of (a). Hence,

the latter because we assume
$V|_{G^\circ }$
is reducible.
Now the summands
$W_i$
are transitively permuted by G, so all have the same dimension as each other, say common dimension M. Under the subgroup
$H^\circ $
, each
$W_i$
is a partial direct sum of the
$H^\circ $
-components of V. In the case
$C_0=0$
, exactly one of these
-components has dimension
$A_0-1 \geq 2$
, and all the others have dimension
$A_0$
. So exactly one of the
$W_i$
has dimension which is
$-1 \ \ (\mathrm {mod} \ {A_0})$
, and any other
$W_i$
has dimension divisible by
$A_0$
. This contradicts the fact that
$\dim (W_i) = M$
for all i.
Assume now that
$C_0 \geq 1$
. By Theorem 6.1, H contains the subgroup

where
$H^{\prime }_{\mathrm {fin}}=H_{\mathrm {fin}}$
, unless
$A_0-1=B_0$
, in which case we take
$H^{\prime }_{\mathrm {fin}} = [H_{\mathrm {fin}},H_{\mathrm {fin}}]$
. Observe that in either case,
splits into a direct sum of
$C_0$
simple modules under
$H^{\prime }_{\mathrm {fin}}$
, one of dimension
$A_0-1$
, and the other
$C_0-1$
of dimension
$A_0$
. On all of these summands,
$H^\circ $
acts trivially; see (6.3.1).
However, the remaining
$C-C_0$
subsheaves
${\mathcal F}(A_0,B_0,\chi )$
,
$\chi \in S_{\mathrm {inf}}$
, give simple, pairwise non-isomorphic
$H^\circ $
-submodules, as mentioned in (a). Thus, each of these simple modules of multiplicity
$1$
must occur in some, and exactly one,
$W_i$
upon restriction to
$H^\circ $
. Call
$W_i$
big if
$W_i|_{H^\circ }$
is nontrivial, equivalently, contains some
${\mathcal F}(A_0,B_0,\chi )$
with
$\chi \in S_{\mathrm {inf}}$
, and small otherwise. As before, we have

and so

Suppose
$W_i$
is big, so its restriction to
$H^\circ $
contains some
${\mathcal F}(A_0,B_0,\chi _i)$
with
$\chi _i \in S_{\mathrm {inf}}$
, and consider any
$h \in H^{\prime }_{\mathrm {fin}}$
. Recall that h, as any other element in G, sends
$W_i$
to some
$W_{i'}$
. Since h centralizes
$H^\circ $
(see (6.3.3)), the
$H^\circ $
-modules
$h(W_i)$
and
$W_i$
are isomorphic and hence have the same
$H^\circ $
simple summands. But
${\mathcal F}(A_0,B_0,\chi _i)$
occurs with multiplicity
$1$
in
$V|_{H^\circ }$
; hence,
$h(W_i)=W_{i'}=W_i$
. Thus,
$W_i$
is stabilized by
$H^{\prime }_{\mathrm {fin}}$
, and hence, it is an
$H^{\prime }_{\mathrm {fin}} \times H^\circ $
-submodule. Recall that all but one simple summand of the
$H^{\prime }_{\mathrm {fin}} \times H^\circ $
-module V has dimension
$A_0$
, and the remaining one,
, has dimension
$A_0-1$
. As
$A_0 \geq 3$
, condition (6.3.4) now implies that
$W_i$
must contain
, which uniquely determines
$W_i$
. We have shown that among the
$W_j$
’s, there is exactly one big summand, and all others are small.
Relabeling the
$W_j$
’s, we may assume
$W_1$
is big and
$W_2, \ldots ,W_m$
are all small. As
$m \geq 2$
by (6.3.2), we have

However, each small
$W_j$
is trivial on
$H^\circ $
(by definition), and so must be contained in
, and does not contain
(which already occurs in the big
$W_1$
). It follows that

As
$C_0 < C/2$
, this contradicts (6.3.5).
In tandem with Theorem 6.3, we prove the following:
Proposition 6.4. Consider the local system with
$r \geq 2$
subject to (4.0.1). Suppose that
$A\ge 5$
, that for some j, A and
$B_j$
have different parity, and that
$G^\circ :=G_{\mathrm {geom},{\mathcal F}}^\circ $
acts irreducibly on
${\mathcal F}$
. Then
${\mathcal F}$
cannot be self-dual for the action of
$G^\circ $
.
Proof. We argue by contradiction. Assume that the underlying
$(A-1)$
-dimensional representation space

for
$G:=G_{\mathrm {geom},{\mathcal F}}$
is self-dual over
$G^\circ $
. Then
$\mathrm {Hom}_{G^\circ }(V^\vee ,V)$
is a one-dimensional representation of
$G/G^\circ $
; call it
${\mathcal L}$
. This means precisely that
$V\cong {\mathcal L} \otimes V^\vee $
as a representation of G. By pullback, we get a geometric isomorphism

for
${\mathcal L}_0$
the restriction of
${\mathcal L}$
to the subgroup
.
Now define
$C:=\gcd (A,B_j)$
, and write
$(A,B_j)=(CA_0,CB_0)$
. Then C is odd, and precisely one of
$A_0,B_0$
is even. In the decomposition

into a direct sum of local systems which are pairwise not geometrically isomorphic, the summand is the unique one of lowest rank
$A_0-1$
. Therefore, the isomorphism above,

gives a geometric isomorphism

for
${\mathcal L}_{0,0}$
the restriction of
${\mathcal L}_0$
to the image in
of
.
We now consider the local system . Up to a multiplicative translation, it is the
$[A_0]^\star $
pullback

Thus, is lisse at
$0$
, and (as
${\mathcal H}_{small,A_0,B_0}$
is hypergeometric of type
$(A_0-1,B_0-1)$
) its
$I(\infty )$
-representation is the direct sum
$\mathsf {Tame} \oplus \mathsf {Wild}$
with
$\mathsf {Tame}$
of rank
$B_0-1$
and
$\mathsf {Wild}$
of rank
$A_0-B_0$
, with all slopes
$A_0/(A_0-B_0)$
. The same statements about local monodromy hold for its dual
.
Our
${\mathcal L}_{0,0}$
is a constituent of
, so is lisse on
${\mathbb A}^1$
.
We first treat the case when
$A_0$
is odd and
$B_0$
is even. Suppose first that
$A_0-B_0 \neq 1$
. Because
$\gcd (A_0,A_0-B_0)=1$
, the slope
$A_0/(A_0-B_0)> 1$
is not an integer. But the
$\infty $
-slope of
${\mathcal L}_{0,0}$
– namely,
$\mathrm {Swan}_\infty ({\mathcal L}_{0,0})$
– is an integer. So if
$\mathrm {Swan}_\infty ({\mathcal L}_{0,0})> 0$
, then
will be totally wild at
$\infty $
, so cannot be geometrically isomorphic to
, which at
$\infty $
has a tame part of dimension
$B_0-1 \ge 1$
(
$\ge 1$
because
$B_0$
is odd). Therefore,
${\mathcal L}_{0,0}$
must be tame at
$\infty $
; hence, is geometrically trivial. But then we have a geometric isomorphism of
with its dual, contradicting (ii-bis) of Lemma 5.5.
Suppose next that
$A_0-B_0 =1$
but that
$A_0-B_0 \neq B_0-1>1$
. Then
at
$\infty $
has a wild part of dimension
$1$
with slope
$A_0$
and a tame part of dimension
$B_0-1>1$
. So if
$\mathrm {Swan}_\infty ({\mathcal L}_{0,0})> 0$
, then
will have a wild part of dimension
$\ge 2$
, and so cannot be geometrically isomorphic to
. Again,
${\mathcal L}_{0,0}$
must be tame at
$\infty $
, hence geometrically trivial, and again a contradiction of (ii-bis) of Lemma 5.5.
Finally, we have the case when
$A_0-B_0=1$
and
$B_0-1=1$
(i.e., the case
$(A_0,B_0) =(3,2)$
). Here,
$(A,B)=(3C,2C)$
. As
$A \ge 5$
, we have
$C>1$
. So in the decomposition

there are
$C-1>1$
distinct irreducible components
${\mathcal F}(A_0,B_0,\chi )$
of rank
$A_0$
. The geometric isomorphism

then gives a geometric isomorphism

Matching irreducible constituents, we see that for some pair
$\chi , \varphi $
of (not necessarily distint) nontrivial characters in
$\mathsf {Char}(C)$
, we have a geometric isomorphism

for some lisse
${\mathcal L}_{\chi ,\varphi }$
of rank one. Again in this situation, both
${\mathcal F}(A_0,B_0,\chi )$
and
${\mathcal F}(A_0,B_0,\varphi )$
are lisse on
${\mathbb A}^1$
. Each is the
$[A_0]^\star $
pullback of a hypergeometric
${\mathcal H}_{big,A_0,B_0,\rho _\chi }$
, respectively
${\mathcal H}_{big,A_0,B_0,\rho _\varphi }$
. Thus, both of
${\mathcal F}(A_0,B_0,\chi )$
and
${\mathcal F}(A_0,B_0,\varphi )$
at
$\infty $
have a wild part of dimension
$A_0-B_0=3-2=1$
and a tame part of dimension
$B_0=2$
. So if
$ {\mathcal L}_{\chi ,\varphi }$
were not tame at
$\infty $
,
$ {\mathcal L}_{\chi ,\varphi }\otimes {\mathcal F}(A_0,B_0,\varphi )^\vee $
would have a wild part of dimension
$\ge B_0=2$
, impossible as
${\mathcal F}(A_0,B_0,\chi )$
has a wild part of dimension
$1$
. Thus,
$ {\mathcal L}_{\chi ,\varphi }$
is geometrically trivial, and hence, we get

contradicting either (ii) or (ii-bis) of Lemma 5.5. This concludes the proof in the case that
$A_0$
is odd and
$B_0$
is even.
We now treat the case when
$A_0$
is even and
$B_0$
is odd. If
$A_0=2$
, then
$B_0=1$
, and hence,
$(A,B)=(2C,C)$
; in this case, we have
$M_{2,2}({\mathcal F})=2$
by Theorem 3.7. As
$G_{\mathrm {geom},{\mathcal F}}$
is infinite, we have
$G_{\mathrm {geom},{\mathcal F}}^0 =\mathrm {SL}_{A-1}$
, and we are done in this
$A_0=2$
case.
It remains to treat the case when
$A_0 \ge 4$
is even and
$B_0$
is odd. If
$\gcd (A,B)=1$
, then
$G_{\mathrm {geom},{\mathcal F}(A,B)}$
is infinite, and hence has
$G_{\mathrm {geom},{\mathcal F}}^0 =\mathrm {SL}_{A-1}$
, and we are done in this case.
Suppose now that
$C>1$
. Just as in the case
$(A_0,B_0) =(3,2)$
discussed above, we match irreducible constituents to see that for some pair
$\chi , \varphi $
of (not necessarily distint) nontrivial characters in
$\mathsf {Char}(C)$
, we have a geometric isomorphism

Both
${\mathcal F}(A_0,B_0,\chi )$
and
${\mathcal F}(A_0,B_0,\varphi )^\vee $
are lisse at
$0$
. As
$I(\infty )$
-representations, each is
$\mathsf {Tame} \oplus \mathsf {Wild}$
, with
$\mathsf {Tame}$
of rank
$B_0$
and
$\mathsf {Wild}$
of rank
$A_0-B_0$
with all slopes
$A_0/(A_0-B_0)$
.
If
$A_0-B_0 \neq 1$
, then the slope
$A_0/(A_0-B_0)$
is not an integer. If
$ {\mathcal L}_{\chi ,\varphi }$
had
$\mathrm {Swan}_\infty>0$
, then
$ {\mathcal L}_{\chi ,\varphi }\otimes {\mathcal F}(A_0,B_0,\varphi )^\vee $
would be totally wild at
$\infty $
, impossible because
${\mathcal F}(A_0,B_0,\chi )$
has a tame part of dimension
$B_0 \ge 1$
. Thus,
$ \ {\mathcal L}_{\chi ,\varphi }$
is geometrically trivial. Then

contradicting either (ii) or (ii-bis) of Lemma 5.5.
It remains to treat the case when
$A_0-B_0=1$
. Here,
$\mathsf {Tame}$
has dimension
$B_0=A_0-1 \ge 3$
, while
$\mathsf {Wild}$
has dimension
$1$
. If
$ {\mathcal L}_{\chi ,\varphi }$
had
$\mathrm {Swan}_\infty>0$
, then
$ {\mathcal L}_{\chi ,\varphi }\otimes {\mathcal F}(A_0,B_0,\varphi )^\vee $
would have a wild part of dimension at least
$B_0 \ge 3$
, impossible because
${\mathcal F}(A_0,B_0,\chi )$
has a wild part of dimension
$1$
. So again here, we get

contradicting either (ii) or (ii-bis) of Lemma 5.5.
Remark 6.5. We exclude the case
$A=3$
in Proposition 6.4 because
has rank two, and in any characteristic,
$p>5$
has infinite
$G_{\mathrm {geom}}$
, and hence
$G_{\mathrm {geom}}^\circ =\mathrm {SL}_2=\mathrm {Sp}_2$
in any characteristic
$p> 5$
.
Now we can determine
$G_{\mathrm {geom}}$
in the presence of infmono
$(A,B)$
:
Theorem 6.6. Consider the local system with
$r \geq 2$
subject to (4.0.1). Suppose that condition infmono
$(A,B)$
holds for some, not necessarily balanced, pair
$(A,B=B_i)$
with
$2 \nmid C=\gcd (A,B)$
. Then the following statements hold for
$G_{\mathrm {geom},{\mathcal F}}$
.
-
(i) Suppose that
$AB$ is even. Then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
-
(ii) Suppose that
$AB$ is odd. If
$2\nmid AB_1 \ldots B_r$ , then
$G_{\mathrm {geom},{\mathcal F}} =\mathrm {Sp}_{A-1}$ . If
$2|AB_1 \ldots B_r$ , then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
Proof. (a) First, we assume that the pair
$(A,B)$
is balanced, and write
$A_0=A/C$
and
$B_0=B/C$
. In view of Theorems 5.3 and 5.4, it suffices to treat the cases with

Recall the condition infmono
$(A,B)$
for the balanced pair
$(A,B)$
implies that
$G=G_{\mathrm {geom},{\mathcal F}}$
is infinite, so
$G^\circ $
is semisimple, say of rank

By Theorem 6.3,
$G^\circ $
acts irreducibly on the underlying representation V of dimension
$D=A-1$
.
We aim to show that
$G^\circ $
is a simple algebraic group. Assume the contrary:

where
$n> 1$
,
$G_i$
is a simple algebraic group of rank
$a_i$
for each
$1 \leq i \leq n$
, and
$a_1 \geq a_2 \geq \ldots \geq a_n \geq 1$
. Thus,

We will derive a contradiction when
$n \ge 2$
.
As
$G^\circ $
acts irreducibly and faithfully on V, the underlying representation of G, we can write

where
$V_i$
is an irreducible
$G_i$
-module of dimension
$d_i \geq 2$
. In fact, by [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we have

Since
$(x+1)(y+1) \geq x+y+1$
for all
$x,y \in {\mathbb Z}_{\geq 0}$
, we have

where

(a1) First, we consider the case
$2|AB$
. By Theorem 5.6(i),
$G^\circ $
contains a semisimple subgroup of rank

and so
$r \geq A-1-C$
. As
$C = A/A_0 \leq A/3$
by (6.6.1), we have

and so

As
$A_0 \geq 3$
and
$C \geq 3$
by (6.6.1), we also have

However,
$a, b \geq 1$
implies that
$(2a-1)(2b-1) \geq 1$
(i.e.,
$2ab \geq a+b$
). Hence, using (6.6.2), we now have

contrary to (6.6.3).
(a2) Now suppose that
$2 \nmid AB$
. By Theorem 5.6(ii),
$G^\circ $
contains a semisimple subgroup H of rank

and so
$r \geq (A-C)/2$
. As
$C = A/A_0 \leq A/5$
by (6.6.1), we have

and so

As
$A_0 \geq 5$
and
$C \geq 3$
by (6.6.1), we also have

Assume in addition that
$b \geq 2$
. Then either
$a,b \geq 3$
, or
$b=2$
but
$a \geq 4$
, and so
$(2a-3)(2b-3) \geq 5$
(i.e.,
$2ab \geq 3a+3b-2$
). Hence, using (6.6.2), we now have

contrary to (6.6.5).
It remains to consider the case
$a_n=b =1$
. Write
$G^\circ = X * Y$
, where
$X := G_1 * \ldots * G_{n-1}$
and
$Y:= G_n$
. Then

is isomorphic to
$\mathrm {SL}_2$
or
$\mathrm {PSL}_2$
. Recall that each of the simple factors of the subgroup H has rank
$\geq (A_0-1)/2 \geq 2$
, and hence, any homomorphism from H to
$\mathrm {SL}_2$
or
$\mathrm {PSL}_2$
is trivial. It follows that
$H \leq X$
. Note that X acts on V via a sum of
$d_n =b+1 \geq 2$
copies of
$V_1 \otimes \ldots \otimes V_{n-1}$
; we see that each simple summand of
$V|_H$
has multiplicity
$\geq 2$
. However, the simple summand
of
$V|_H$
has multiplicity
$1$
, again a contradiction.
(b) Continue with the assumption of (a). We have shown that
$G^\circ $
is a simple algebraic group of rank
$r \geq 5$
; see (6.6.4), (6.6.6). Furthermore, (6.6.3), respectively (6.6.5), still holds, so

In particular,
$D \leq 14$
if
$r=6$
,
$D \leq 16$
if
$r = 7$
, and
$D \leq 19$
if
$r=8$
. Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.12], we see that
$G^\circ $
is not an exceptional algebraic group, and thus, it is a classical group. Since
$r \geq 5$
, we have that

and

when
$r=5$
. Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we conclude that
$V|_{G^\circ }$
(of dimension
$D=A-1$
) must be the natural module or its dual for the classical group
$G^\circ $
. In other words,
$G^\circ = \mathrm {SL}_D$
,
$\mathrm {Sp}_D$
, or
$\mathrm {SO}_D$
.
Suppose
$2|AB$
. Then (6.6.3) rules out the groups
$\mathrm {Sp}_D$
and
$\mathrm {SO}_D$
since they have
$D \geq 2r$
. Hence, we must have
$G^\circ = \mathrm {SL}_{A-1}$
in this case.
Suppose
$2\nmid AB$
. The choice of the balanced pair
$(A,B)$
implies that A and
$B_i$
are all odd, so V is symplectic, ruling out
$\mathrm {SL}_D$
and
$\mathrm {SO}_D$
. Hence, we must have
$\mathrm {Sp}_D=G^\circ \leq G \leq \mathrm {Sp}_D$
, and so
$G=\mathrm {Sp}_{A-1}$
.
(c) It remains to consider the case
$2|AB_1 \ldots B_r$
,
$2 \nmid AB$
, and infmono
$(A,B)$
holds for the (unbalanced) pair
$(A,B=B_i)$
. By Theorem 6.3, we still know that
$G^\circ $
is irreducible on V.
Suppose first that
$\gcd (A,B_i)=1$
. Then
is
$\mathrm {Sp}_{A-1}$
by [Reference Katz and TiepKT6, Theorems 10.2.4(iii) and 10.3.21(iii)]. As
${\mathcal F}(A,B)$
is a pullback of
${\mathcal F}$
, our
$G=G_{\mathrm {geom},{\mathcal F}}$
, and hence,
$G^\circ $
contains
$H = \mathrm {Sp}_{A-1}$
. Thus

Assume
$G^\circ $
is not simple. Now we can continue the analysis in (a2) to show that
$r=a+b$
with
$a \geq b \geq 1$
and
$D \geq (a+1)(b+1)$
. If
$b \geq 2$
, then
$(a-1)(b-1) \geq 1$
,
$ab \geq a+b$
, and so
$D \geq 2(a+b)+1 =2r+1$
, contradicting (6.6.7). If
$b=1$
, then as in (a2), we arrive at the contradiction that
$V|_H$
has simple summands with multiplicity
$\geq 2$
.
We have shown that
$G^\circ $
is simple of rank r. Recall that
$A> B_i$
are odd, so
$A \geq 3$
and
$D \geq 2$
. If
$r=1$
or
$A=3$
, then necessarily,
$D=2$
,
$G^\circ = \mathrm {SL}_2$
, and we are done. We may therefore assume

Hence, (6.6.7) implies
$D \leq (5r-2)/2$
. Assume in addition that
$r \geq 5$
. Then the same arguments as in (b) show that
$G^\circ \cong \mathrm {SL}_D$
,
$\mathrm {Sp}_D$
, or
$\mathrm {SO}_D$
. Applying Proposition 6.4, we conclude that
$G^\circ = \mathrm {SL}_D$
.
Suppose
$r= 4$
. Then
$G^\circ = \mathrm {SL}_5$
,
$\mathrm {SO}_9$
,
$\mathrm {Sp}_8$
,
$\mathrm {SO}_8$
or
$F_4$
, and
$D \leq 8$
by (6.6.7). Since V is irreducible and faithful over
$G^\circ $
, using [Reference LübeckLu], we see that
$(G^\circ ,D) = (\mathrm {SL}_5,5)$
,
$(\mathrm {Sp}_8,8)$
, or
$(\mathrm {SO}_8,8)$
. The latter two cases are impossible by Proposition 6.4, so
$G^\circ = \mathrm {SL}_D$
.
Suppose
$r= 3$
. Then
$G^\circ = \mathrm {SL}_4$
,
$\mathrm {SO}_7$
, or
$\mathrm {Sp}_6$
, and
$D \leq 6$
by (6.6.7). Since V is irreducible and faithful over
$G^\circ $
, using [Reference LübeckLu], we see that
$(G^\circ ,D) = (\mathrm {SL}_4,4)$
,
$(\mathrm {SO}_6,6)$
, or
$(\mathrm {Sp}_6,6)$
. The latter two cases are impossible by Proposition 6.4, so
$G^\circ = \mathrm {SL}_D$
.
Suppose
$r= 2$
. Then
$G^\circ = \mathrm {SL}_3$
,
$\mathrm {Sp}_4$
, or
$G_2$
, and
$D \leq 4$
by (6.6.7). Since V is irreducible and faithful over
$G^\circ $
, using [Reference LübeckLu], we see that
$(G^\circ ,D) = (\mathrm {SL}_3,3)$
, or
$(\mathrm {Sp}_4,6)$
. The latter case is ruled out by Proposition 6.4, and so
$G^\circ = \mathrm {SL}_D$
.
Now suppose that
$\gcd (A,B_i)=C>1$
for the unbalanced pair
$(A,B=B_i)$
with infmono
$(A,B)$
. Again write
$(A,B)=(CA_0,CB_0)$
. If
$A_0=3$
, then
$B_0=1$
, and we are done by Theorem 5.3(ii).
It now remains to treat the case
$A_0 \ge 5$
. Exactly as in the discussion of the case when
$2 \nmid AB$
in the balanced case, we prove that
$G_{\mathrm {geom},{\mathcal F}}^0$
is a simple algebraic group, and then that
$G^\circ $
is one of the classical groups
$\mathrm {SL}_{A-1}$
,
$\mathrm {Sp}_{A-1}$
if A is odd, or
$\mathrm {SO}_{A-1}$
, acting on its natural module or its dual. Proposition 6.4 then shows that
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$
.
Our next result visibly improves Theorem 6.6:
Theorem 6.7. Consider the local system with
$r \geq 2$
subject to (4.0.1). Suppose that, for some (not necessarily balanced) pair
$(A,B:=B_i)$
with
$2 \nmid C=\gcd (A,B)$
,
has infinite geometric monodromy group H. Then the following statements hold.
-
(i) Suppose that
$AB$ is even. Then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
-
(ii) Suppose that
$AB$ is odd. If
$2\nmid AB_1 \ldots B_r$ , then
$G_{\mathrm {geom},{\mathcal F}} =\mathrm {Sp}_{A-1}$ . If
$2|AB_1 \ldots B_r$ , then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
Proof. (a) Since is a pullback of
${\mathcal F}$
,
$H \leq G:=G_{\mathrm {geom},{\mathcal F}}$
, and G is infinite. Hence,
$G^\circ $
is semisimple, say of rank

We aim to show that
$G^\circ $
is a simple algebraic group. Assume the contrary:

where
$n> 1$
,
$G_i$
is a simple algebraic group of rank
$a_i$
for each
$1 \leq i \leq n$
, and
$a_1 \geq a_2 \geq \ldots \geq a_n \geq 1$
. Thus,

We will derive a contradiction when
$n \ge 2$
.
By Theorem 6.3,
$G^\circ $
acts irreducibly and faithfully on V, the underlying representation of G. So we can write

where
$V_i$
is an irreducible
$G_i$
-module of dimension
$d_i \geq 2$
. In fact, by [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we have

Since
$(x+1)(y+1) \geq x+y+1$
for all
$x,y \in {\mathbb Z}_{\geq 0}$
, we have

where

Write
$A_0=A/C$
and
$B_0=B/C$
, and let V denote the underlying representation for G. Also, let
$S_{\mathrm {inf}}$
denote the set of characters
$\chi \in \mathsf {Char}(C)$
such that
${\mathcal F}(A_0,B_0,\chi )$
has infinite
$G_{\mathrm {geom}}$
. By Theorem 6.6, we may assume

so that
$S_{\mathrm {inf}} = \mathsf {Char}(C) \smallsetminus \mathsf {Char}(C_0)$
for some proper divisor
$C_0$
of C by Lemma 5.1(ii). As C is odd, we have

Also, in view of Theorem 5.4, it suffices to treat the cases with

(a1) First, we consider the case
$2|AB$
. By Theorem 6.1,
$G^\circ $
contains a semisimple subgroup
$H^\circ $
of rank
$(A_0-1)(C-C_0) \geq 2C(A_0-1)/3$
. Namely,
$H^\circ $
is the product of the
$\mathrm {SL}_{A_0}$
factors, one for each
${\mathcal F}(A_0,B_0,\chi )$
with
$\chi \in S_{\mathrm {inf}}$
. By Corollary 6.2, the subsheaves
${\mathcal F}(A_0,B_0,\chi )$
with
$\chi \in S_{\mathrm {inf}}$
are simple summands with multiplicity
$1$
for the module
$V|_{H^\circ }$
. Thus, we have

In this case,
$C = (D+1)/A_0 \leq (D+1)/3$
, so
$3r \geq 2A_0C-2C \geq 4(D+1)/3$
, and thus,

Suppose that
$b < A_0-1$
. Then every homomorphism from
$H^\circ $
to
$G_n$
of rank b is trivial. It follows that
$H^\circ \leq G_1 * \ldots * G_{n-1}$
, and hence, the restriction of V to
$H^\circ $
is a sum of
$d_n=\dim (V_n) \geq 2$
copies of the same module. But this contradicts the above multiplicity-one assertion.
We have therefore shown that

Together with (6.7.3) and (6.7.4), this implies that

Hence, by (6.7.4), we have

This, however, contradicts (6.7.1).
(a2) Next, we consider the case
$2 \nmid AB$
. By Theorem 6.1,
$G^\circ $
contains a semisimple subgroup
$H^\circ $
of rank
$(A_0-1)(C-C_0)/2 \geq C(A_0-1)/3$
. Namely,
$H^\circ $
is the product of the
$\mathrm {SL}_{A_0}$
factors, one for each
${\mathcal F}(A_0,B_0,\chi )$
with
$\chi \in \mathrm {Rep}(S_{\mathrm {inf}})$
. By Corollary 6.2, the subsheaves
${\mathcal F}(A_0,B_0,\chi )$
with
$\chi \in S_{\mathrm {inf}}$
are simple summands with multiplicity
$1$
for the module
$V|_{H^\circ }$
. Thus, we have

In this case,
$C = (D+1)/A_0 \leq (D+1)/5$
, so
$3r \geq A_0C-C \geq 4(D+1)/5$
, and thus,

Arguing as in (a1), we see that (6.7.6) still holds. Together with (6.7.3) and (6.7.7), this implies that

Suppose
$C \geq 9$
. Then

since
$A_0 \geq 5$
. This, however, contradicts (6.7.1).
It remains to consider the case
$C \in \{3,5,7\}$
, in which case we have
$C_0 = 1$
and
$r \geq (A_0-1)(C-1)/2$
. Now we have

If
$C=7$
, then
$ab \geq 8(A_0-1)$
,
$a+b=r \geq 3(A_0-1)$
, and so

If
$C=5$
, then
$ab \geq 4(A_0-1)$
,
$a+b \geq 2(A_0-1)$
, and so

If
$C=3$
, then
$ab \geq (A_0-1)^2 \geq 4(A_0-1)$
,
$a+b =r \geq (A_0-1)$
, and so

In all cases, we arrive at a contradiction with (6.7.1).
(b) We have shown that
$G^\circ $
is a simple algebraic group of rank
$r \geq 4$
; see (6.7.4), (6.7.7). Furthermore, (6.7.5), respectively (6.7.8), still holds, so

In particular,
$D \leq 14$
if
$r=4$
,
$D \leq 21$
if
$r=6$
,
$D \leq 25$
if
$r = 7$
, and
$D \leq 29$
if
$r=8$
. Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.12], we see that
$G^\circ $
is not an exceptional algebraic group, and thus, it is a classical group.
Assume in addition that
$r \geq 7$
. Then

Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we conclude that
$V|_{G^\circ }$
(of dimension
$D=A-1$
) must be the natural module or its dual for the classical group
$G^\circ $
. In other words,
$G^\circ = \mathrm {SL}_D$
,
$\mathrm {Sp}_D$
, or
$\mathrm {SO}_D$
.
Suppose
$2|AB_1 \ldots B_r$
. As
$A=A_0C \geq 9$
, Proposition 6.4 rules out the groups
$\mathrm {Sp}_D$
and
$\mathrm {SO}_D$
. Hence, we must have
$G^\circ = \mathrm {SL}_{A-1}$
in this case.
Suppose
$2\nmid AB_1 \ldots B_r$
. Then V is symplectic, ruling out
$\mathrm {SL}_D$
and
$\mathrm {SO}_D$
. Hence, we must have
$\mathrm {Sp}_D=G^\circ \leq G \leq \mathrm {Sp}_D$
, and so
$G=\mathrm {Sp}_{A-1}$
.
(c) Now we return to the general case, where we know only that
$r \geq 4$
. If
$r \geq 7$
, then we are done by (b).
Suppose that
$A \geq 14$
if
$2|AB$
and
$A \geq 23$
if
$2 \nmid AB$
. In the former case,
$r> 6$
by (6.7.5). In the latter case,
$r> 6$
by (6.7.8). Thus, we have
$r \geq 7$
, and so are again done by (b).
The rest of the proof is to analyze the remaining cases, in which we may assume

Suppose
$A=21$
. Then
$2 \nmid B$
, and
$C \in \{3,7\}$
. By (6.7.3), we have
$A_0 \neq 3$
, so
$(A_0,C) =(7,3)$
and
$r \geq 6$
by (6.7.8). In view of (6.7.9), we now have
$r=6$
, but
$G^\circ \geq H^\circ = \mathrm {SL}_7$
. So in fact,
$G^\circ = H^\circ =H^\circ _{\mathrm {inf}}$
, and hence,
$G^\circ $
is reducible on V by (6.7.2), a contradiction.
Suppose
$A \in \{3,5,7,11,13,17,19\} \cup \{2,4,8\}$
. Since C is an odd proper divisor of A, in these cases, we must have
$C=1$
, violating (6.7.2).
Suppose
$A=15$
. Then
$G^\circ $
is a simple classical group of rank
$4 \leq r \leq 6$
acting irreducibly on
$V = {\mathbb C}^{14}$
. This is impossible by [Reference LübeckLu].
Suppose
$A=12$
. Then
$(A_0,C)=(4,3)$
and
$r \geq 6$
by (6.7.5), whence
$r=6$
. Now
$G^\circ $
is a simple classical group of rank
$6$
acting irreducibly on
$V = {\mathbb C}^{11}$
. This is impossible by [Reference LübeckLu].
If
$A=10$
, then
$C=5$
and
$A_0=2$
, violating (6.7.3).
Suppose
$A=9$
. Then
$(A_0,C)=(3,3)$
by (6.7.3), so
$2|AB$
. Now
$G^\circ $
is a simple classical group of rank
$4 \leq r \leq 6$
acting irreducibly on
$V = {\mathbb C}^{8}$
, whence
$G^\circ \cong \mathrm {Sp}_8$
or
$\mathrm {SO}_8$
by [Reference LübeckLu]. In either case, this contradicts Proposition 6.4.
Finally, if
$A=6$
, then
$C=3$
and
$A_0=2$
, violating (6.7.3).
Next, we prove the following extension of [Reference Katz and TiepKT6, Theorem 11.1.3]:
Theorem 6.8. Let
$V = {\mathbb C}^D$
with
$D \geq 6$
, and let
$G \leq \mathrm {GL}(V)$
be a Zariski closed, irreducible subgroup, with
$G^\circ \neq 1$
being semisimple. Suppose that G contains a subgroup
$G_1$
which is one of the following groups.
-
(a)
$G_1$ is the image of
$\mathrm {Sp}_{2n}(q)$ in a nontrivial subrepresentation of degree D of a total Weil representation of degree
$q^n$ for some odd prime power
$q=p^f$ and some
$n \geq 1$ . Furthermore, if
$D=q^n$ , assume that
$q^n \geq 13$ .
-
(b)
$G_1$ is the image of
$\mathrm {SU}_{n}(q)$ in a nontrivial subrepresentation of degree D of the total Weil representation of degree
$q^n$ for some prime power
$q=p^f$ and some odd
$n \geq 3$ with
$(n,q) \neq (3,2)$ . Furthermore, if
$G_1$ is reducible on V, assume that
$V|_{G_1}$ contains a simple summand of dimension
$(q^n-q)/(q+1)$ , and, in addition,
$(n,q) \neq (3,3)$ .
-
(c)
$G_1$ is the image of
$2 \cdot \mathrm {J}_2$ in an irreducible representation of degree
$D=6$ .
-
(d)
$G_1$ is the image of
$6_1 \cdot \mathrm {PSU}_4(3)$ in an irreducible representation of degree
$D=6$ .
-
(e)
$G_1$ is the image of
$2 \cdot G_2(4)$ in an irreducible representation of degree
$D=12$ .
Then
$G^\circ $
is a simple algebraic group acting irreducibly on V and
$G^\circ> G_1$
. Moreover, one of the following conclusions holds.
-
(i)
$G^\circ = \mathrm {SL}(V)$ ,
$\mathrm {Sp}(V)$ , or
$\mathrm {SO}(V)$ .
-
(ii)
$D=32$ ,
$G_1=\mathrm {SU}_5(2)$ , and
$G^\circ = \mathrm {Sp}_{10}$ .
-
(iii)
$(G^\circ ,D) = (G_2,7)$ , and
$G_1=\mathrm {PSL}_2(13)$ or
$\mathrm {SU}_3(3)$ .
-
(iv)
$(G^\circ ,D)=(E_6,27)$ and
$G_1=\mathrm {SL}_2(27)$ .
-
(v)
$(G^\circ ,D)=(E_7,56)$ and
$G_1=\mathrm {PSU}_3(8)$ .
Proof. (A) By assumption,
$q^n \geq D \geq 6$
in (a) and
$(n,q) \neq (3,2)$
in (b), so
$G_1$
is quasisimple. According to [Reference Kleidman and LiebeckKlL2, Table 5.2.A] and [Reference Conway, Curtis, Norton, Parker and WilsonAtlas], for the smallest index
$P(G_1)$
of proper subgroups of
$G_1$
we have
$P(G_1)\geq q^n+1> D$
in case (a), unless
$(n,q) = (1,11)$
, for which we have
$D \leq (q^n+1)/2$
by hypothesis, and
$P(G_1) \geq (q^n+3)/2> D$
. Similarly,
$P(G_1) \geq q^n+1> D$
in case (b), unless
$(n,q) \neq (3,5)$
, in which case
$P(G_1) = 50$
. Furthermore,
$P(G_1)=100> D$
in case (c),
$P(G_1)=112> D$
in case (d), and
$P(G_1)=416> D$
in case (e). Thus, in all cases, we have

where
$S:= G_1/{\mathbf Z}(G_1)$
is simple.
By [Reference Kleidman and LiebeckKlL1, Theorem 3], the smallest degree
$e(G_1)$
of any nontrivial projective representation of
$G_1$
(over
${\mathbb C}$
) is at least the smallest degree
$e(S)$
of any nontrivial projective representation of S (over
${\mathbb C}$
). According to [Reference Kleidman and LiebeckKlL2, Table 5.3.A],

in case (a), unless
$D \leq 12$
. If
$6 \leq D \leq 12$
but
$V|_{G_1}$
is irreducible, then
$q^n \geq 11$
(as
$D \geq 6$
) and

If
$6 \leq D \leq 12$
but
$V|_{G_1}$
is reducible, then
$D=q^n \leq 11$
, which is excluded by our hypothesis. Similarly,

in case (b), and

in cases (c)–(e). Moreover, in all cases, the smallest nontrivial projective representation of S is also a projective representation of
$G_1$
, so in fact, we have

(A1) By assumption, G acts irreducibly on
$V := {\mathbb C}^{D}$
, and
$G_1$
is quasisimple. Suppose that G fixes an imprimitive decomposition

with
$m> 1$
. Then
$1 < m|D$
and
$\dim (V_i) = D/m$
for all i.
Suppose we are in the case
$P(G_1)> D$
of (6.8.1). Then every homomorphism
$G_1 \to \mathsf {S}_m$
is trivial, and so the action of
$G_1$
on the m summands
$V_i$
is trivial. In other words,
$G_1$
stabilizes every
$V_i$
. If in addition
$V|_{G_1}$
is irreducible, then
$V_1$
, being fixed by
$G_1$
, is equal to V, contrary to
$m> 1$
. So
$V|_{G_1}$
is reducible. In particular, we are either in (a) with
$D=q^n$
, in which case
$V|_{G_1}$
is a sum
$W_1 \oplus W_2$
of two simple summands,
$W_1$
of dimension
$d:=(q^n-1)/2$
and
$W_2$
of dimension
$d+1=(q^n+1)/2$
, or in (b), in which case
$V|_{G_1}$
is a sum
$\oplus ^s_{i=1}$
of
$s> 1$
simple summands,
$W_1$
of dimension
$d:=(q^n-q)/(q+1)$
and
$W_2, \ldots ,W_s$
all of dimension
$d+1=(q^n+1)/(q+1)$
. In either case, because
$W_1$
is a simple summand of multiplicity one in
$V|_{G_1}$
, we may assume that
$W_1$
occurs in
$V_1$
(and only in
$V_1$
). Since each
$W_i$
with
$i> 1$
has dimension
$d+1$
, we see that

Thus,
$\dim (V_1) \neq \dim (V_2)$
, a contradiction.
Suppose now that we are in the case
$P(G_1) \leq D$
of (6.8.1), so that
$G_1/{\mathbf Z}(G_1) = \mathrm {PSU}_3(5)$
. The same arguments as above show that
$G_1$
cannot act trivially on the set
$\{V_1, \ldots ,V_m\}$
. As
$P(G_1) = 50$
, we must have that
$m \geq 50$
, and so
$\dim (V_i) \leq 2$
as
$D \leq 5^3$
. But the simple summands of
$V|_{G_1}$
has dimension
$20$
or
$21$
, so no
$V_i$
can be fixed by
$G_1$
. Using [Reference Conway, Curtis, Norton, Parker and WilsonAtlas] we can check that every proper subgroup of
$G_1$
has index
$50$
or
$\geq 126> m$
. It follows that every
$G_1$
-orbit on
$\{V_1, \ldots ,V_m\}$
is of length
$50$
, and hence,
$50|D$
. However, by (6.8.1) and hypothesis,
$D \in \{62,83,104,125\}$
, a contradiction.
(A2) We have shown that G acts primitively on V. Let
$\Phi $
denote the representation of G on V. Next suppose that
$G_1$
fixes a tensor decomposition

with
$1 < \dim (V_1) \leq \dim (V_2)$
. Then the quasisimple group
$G_1$
admits a projective representation on
$V_1$
, of dimension
$\leq \sqrt {D}$
, whose image is either trivial, or a quasisimple cover of the simple group
$S:=G_1/{\mathbf Z}(G_1)$
. By (6.8.6), every composition factor of the projective representation of
$G_1$
on
$V_1$
is trivial, and so the corresponding image of
$G_1$
in
$\mathrm {PGL}(V_1)$
is contained in a Borel subgroup which is solvable. As
$G_1$
is quasisimple, this image is trivial (i.e.,
$G_1$
acts via scalars on
$V_1$
). Pulling the constants to the action of
$G_1$
on
$V_2$
, for every
$g \in G_1$
, we can write

for a unique matrix
$\Psi (g) \in \mathrm {GL}(V_2)$
. Since
$\Phi |_{G_1}$
is a linear representation, it follows that
$\Psi $
is a linear representation
$G_1 \to \mathrm {GL}(V_2)$
. Thus,
$\Phi |_{G_1}$
is the sum of
$\dim (V_1)> 1$
copies of the representation
$\Psi $
. In particular, every simple summand of
$V|_{G_1}$
occurs with multiplicity
$\geq \dim (V_1)> 1$
. But this contradicts our hypothesis on
$V|_{G_1}$
.
(A3) We have shown that
$G_1$
cannot fix any tensor decomposition of V. Finally, suppose
$G_1$
fixes a tensor induced decomposition

with
$m \geq 2$
and
$\dim (V_i) \geq 2$
. Then

the latter inequality because of (6.8.1). In such a case,
$G_1$
must fix every tensor factor
$V_i$
, and hence,
$V|_{G_1}$
is tensor decomposable, contrary to the preceding result.
Note that
${\mathbf Z}(G)^\circ \leq {\mathbf Z}(G^\circ )$
is finite since
$G^\circ $
is semisimple. Thus,
${\mathbf Z}(G)$
is finite. We have shown that
$(V,G)$
satisfies condition
$\mathrm {(\mathbf {S+})}$
. By [Reference Katz and TiepKT3, Lemma 1.1],
$G^\circ $
is a simple algebraic group acting irreducibly on V.
(B) By Schur’s lemma,
${\mathbf C}_{G}(G^\circ ) = {\mathbf Z}(G)$
is cyclic. Furthermore,
$\mathrm {Out}(G^\circ )$
is a subgroup of
$\mathsf {S}_3$
, and hence solvable. It follows that
$G/G^\circ $
is solvable. But
$G_1$
is perfect, so
$G_1 < G^\circ $
.
Let r denote the rank of the simple algebraic group
$G^\circ $
. We will now analyze each of the possibilities for
$G^\circ $
.
(B1) Suppose
$G^\circ $
is of type
$A_r$
. In this case,
$G^\circ $
admits an irreducible projective complex representation
$\Theta $
of dimension
$r+1$
with finite cyclic kernel. Thus,
$\Theta |_{G_1}$
is now a nontrivial projective representation, and, arguing as in (A), we see that

by (6.8.6). It follows that
$r \geq 4$
and
$D \leq 4(r+1)^2/11 < r(r+1)/2$
. Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we conclude that the D-dimensional module V of
$G^\circ $
is the natural module, or its dual, and so
$G^\circ = \mathrm {SL}_D$
.
(B2) Suppose
$G^\circ $
is of type
$B_r$
with
$r \geq 2$
. In this case,
$G^\circ $
admits an irreducible projective complex representation
$\Theta $
of dimension
$2r+1$
with finite cyclic kernel and with image contained in
$\mathrm {PSO}_{2r+1}$
. Thus,
$\Theta |_{G_1}$
is now a nontrivial projective representation, and hence,

by (6.8.6). Now, if
$r=2$
, then
$e(G_1)=5$
, and so by (6.8.2)–(6.8.5), we must have that
$G_1/{\mathbf Z}(G_1)=\mathrm {PSL}_2(11)$
. Since
$\mathrm {SL}_2(11)$
is the universal cover of
$\mathrm {PSL}_2(11)$
,
$\Theta $
lifts to a
$5$
-dimensional orthogonal representation of
$\mathrm {SL}_2(11)$
, which is impossible. So

Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we see that either the D-dimensional module V of
$G^\circ $
is the natural module of dimension
$2r+1$
and so
$G^\circ = \mathrm {SO}_D$
, or
$3 \leq r \leq 5$
and
$D=2^r$
. It remains to look at the latter possibilities.
Note that if
$D=8$
, then
$G_1=\mathrm {SL}_2(17)$
and
$V|_{G_1}$
is irreducible of symplectic type. (Indeed, if we are in case (a), then, as
$q^n \neq 8$
, we have

whence
$(n,q)=(1,17)$
and so
$G_1 = \mathrm {SL}_2(17)$
in an irreducible Weil representation. If we are in case (b), then
$8=D \geq (q^n-q)/(q+1)$
, which is at least
$10$
if
$n \geq 5$
, and at least
$12$
if
$n = 3$
but
$q \geq 4$
. So
$(n,q) = (3,3)$
, and we quickly reach a contradiction.) However, in this case, we have
$e(G_1) = 8$
and
$r=3$
, contrary to (6.8.7).
Similarly, if
$D=16$
,
$32$
, or
$64$
, then either
$(D,G_1)=(16,\mathrm {SL}_2(31))$
,
$(64,\mathrm {SL}_2(127))$
and
$V|_{G_1}$
is irreducible, or
$D=32$
and
$G_1=\mathrm {SU}_5(2)$
, or
$D=64$
and
$G_1 = \mathrm {SU}_3(4)$
. (Indeed, if we are in case (a), then, as
$q^n \neq 16,32,64$
, we have
$(q^n \pm 1)/2 = D = 16$
,
$32$
, or
$64$
, whence
$(D,n,q)=(16,1,31)$
or
$(64,1,127)$
and
$G_1$
acts in an irreducible Weil representation. If we are in case (b), then
$D \geq (q^n-q)/(q+1)$
, which is at least
$72$
if
$n \geq 9$
, or
$n=7$
but
$q \geq 3$
, or
$n = 5$
but
$q \geq 4$
, or
$n=3$
but
$q \geq 9$
. If
$(n,q) = (7,2)$
, then
$D \in \{42,43,\geq 85\}$
. If
$(n,q) = (5,3)$
, then
$D \in \{60,61,\geq 121\}$
. If
$(n,q) = (5,2)$
, then
$D \in \{10,11,21,32\}$
by hypothesis, so
$D=32$
. If
$(n,q) = (3,8)$
, then
$D \in \{56,57,\geq 113\}$
. If
$(n,q) = (3,7)$
, then
$D \in \{42,43,\geq 85\}$
. If
$(n,q) = (3,5)$
, then
$D \in \{20,21,41,62, \geq 83\}$
by hypothesis. If
$(n,q) = (3,4)$
, then by hypothesis
$D \in \{12,13,25,38,51,64\}$
, so
$D=64$
. If
$(n,q) = (3,3)$
, then
$D=6,7$
by hypothesis.)
Now, in the case
$D=16$
and
$G_1= \mathrm {SL}_2(31)$
, we have
$e(G_1) = 15$
and
$r=4$
, contrary to (6.8.7).
In the case
$D=32$
and
$G_1=\mathrm {SU}_5(2)$
, we have
$G^\circ = \mathrm {Spin}_{11}$
. The projection
$G^\circ \to \mathrm {SO}_{11}$
has kernel of order
$2$
which must then intersect
$G_1$
trivially since
$G_1$
is simple. It follows that
$G_1$
embeds in
$\mathrm {SO}_{11}$
. But this is a contradiction, since every nontrivial complex representation of degree
$11$
of
$\mathrm {SU}_5(2)$
is either irreducible non-self-dual, or a direct sum of a trivial representation and an irreducible representation of symplectic type (of degree
$10$
) see [Reference Conway, Curtis, Norton, Parker and WilsonAtlas].
(B3) Suppose
$G^\circ $
is of type
$D_r$
with
$r \geq 4$
. In this case,
$G^\circ $
admits an irreducible projective complex representation
$\Theta $
of dimension
$2r$
with finite cyclic kernel and with image contained in
$\mathrm {PSO}_{2r}$
. Thus,
$\Theta |_{G_1}$
is now a nontrivial projective representation, and hence,

by (6.8.6), so

Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we see that either the D-dimensional module V of
$G^\circ $
is (quasi-equivalent in the case
$r=4$
to) the natural module of dimension
$2r$
and so
$G^\circ \cong \mathrm {SO}_D$
, or
$4 \leq r \leq 7$
and
$D=2^{r-1}$
. It remains to look at the latter possibilities.
If
$(r,D) = (4,8)$
, then
$G_1=\mathrm {SL}_2(17)$
and
$V|_{G_1}$
is of symplectic type, as shown in (B2). However,
$V|_{G^\circ }$
is quasi-equivalent to the natural module, so it is of orthogonal type, a contradiction.
If
$(r,D) = (5,16)$
, then
$G_1=\mathrm {SL}_2(31)$
as shown in (B2). However, in this case,
$e(G_1)=15$
and
$r=5$
, contrary to (6.8.8).
If
$(r,D) = (6,32)$
, then
$G_1=\mathrm {SU}_5(2)$
as shown in (B2), and
$G^\circ = \mathrm {Spin}_{12}$
. The projection
$G^\circ \to \mathrm {SO}_{12}$
has kernel of order
$2$
which must then intersect
$G_1$
trivially since
$G_1$
is simple. It follows that
$G_1$
embeds in
$\mathrm {SO}_{12}$
. But this is a contradiction, since every nontrivial complex representation of degree
$12$
of
$\mathrm {SU}_5(2)$
is either a sum of of a trivial representation and a non-self-dual irreducible representation (of degree
$11$
), or a sum of two copies of the trivial representation and an irreducible representation of symplectic type (of degree
$10$
); see [Reference Conway, Curtis, Norton, Parker and WilsonAtlas].
If
$(r,D) = (7,64)$
, then
$G_1=\mathrm {SU}_3(4)$
as shown in (B2), and
$G^\circ = \mathrm {Spin}_{14}$
. The projection
$G^\circ \to \mathrm {SO}_{14}$
has kernel of order
$2$
which must then intersect
$G_1$
trivially since
$G_1$
is simple. It follows that
$G_1$
embeds in
$\mathrm {SO}_{14}$
. But this is a contradiction, since every nontrivial complex representation of degree
$14$
of
$\mathrm {SU}_3(4)$
is either a sum of of a trivial representation and a non-self-dual irreducible representation (of degree
$13$
), or a sum of two copies of the trivial representation and an irreducible representation of symplectic type (of degree
$12$
), see [Reference Conway, Curtis, Norton, Parker and WilsonAtlas].
(B4) Suppose
$G^\circ $
is of type
$C_r$
with
$r \geq 3$
. In this case,
$G^\circ $
admits an irreducible projective complex representation
$\Theta $
of dimension
$2r$
with finite cyclic kernel and with image contained in
$\mathrm {PSp}_{2r}$
. Thus,
$\Theta |_{G_1}$
is now a nontrivial projective representation, and hence, (6.8.8) holds by (6.8.6), and so

Applying [Reference Kleidman and LiebeckKlL2, Proposition 5.4.11], we see that either the D-dimensional module V of
$G^\circ $
is the natural module of dimension
$2r$
and so
$G^\circ \cong \mathrm {Sp}_D$
, or
$3 \leq r \leq 5$
and
$D=2^r$
, or
$(r,D)=(3,14)$
. It remains to look at the latter possibilities.
If
$D=8$
, then
$G_1=\mathrm {SL}_2(17)$
as shown in (B2). However, in this case, we have
$e(G_1) = 8$
and
$r=3$
, contrary to (6.8.8).
If
$D=16$
, then
$G_1=\mathrm {SL}_2(31)$
as shown in (B2). However, in this case, we have
$e(G_1) = 15$
and
$r=4$
, contrary to (6.8.8).
If
$D=32$
, then
$G_1=\mathrm {SU}_5(2)$
as shown in (B2), and this is recorded in conclusion (ii).
If
$(r,D) = (3,14)$
, then this violates (6.8.8).
(B5) Suppose
$G^\circ = G_2$
. Then
$G^\circ < \mathrm {SL}_7$
, and so

whence
$D \leq 17$
by (6.8.6). Since
$V|_{G^\circ }$
is irreducible, we must have
$D = 7$
or
$14$
by [Reference LübeckLu].
Suppose
$D=7$
. In case (a), (since
$q^n \geq 13$
when
$D=q^n$
), we have
$(q^n \pm 1)/2 = D = 7$
, and so
$(n,q)=(1,13)$
and
$G_1 = \mathrm {PSL}_2(13)$
. In case (b), since
$D \geq (q^n-q)/(q+1)$
is at least
$10$
when
$n \geq 5$
or
$n=3$
but
$q \geq 4$
, we have
$(n,q) = (3,3)$
and
$G_1=\mathrm {SU}_3(3)$
. These two possibilities are recorded in conclusion (iii).
Suppose
$D=14$
. In case (a), we have
$(q^n \pm 1)/2 = D = 14$
, and so
$(n,q)=(1,27)$
and
$G_1 = \mathrm {SL}_2(27)$
. This violates (6.8.9) since
$e(G_1)=13$
. In case (b),
$D \geq (q^n-q)/(q+1)$
is at least
$20$
when
$n \geq 7$
, or
$n=5$
but
$q \geq 3$
, or
$n=3$
but
$q \geq 5$
. Now if
$(n,q) = (5,2)$
, then
$D \in \{10,11,\geq 21\}$
. If
$(n,q) = (3,4)$
, then
$D \in \{12,13,\geq 25\}$
. If
$(n,q) = (3,3)$
, then
$D \in \{6,7\}$
by hypothesis.
(B6) Suppose
$G^\circ = F_4$
. Then
$G^\circ < \mathrm {SL}_{26}$
, and so

whence
$D \leq 245$
by (6.8.6). Since
$V|_{G^\circ }$
is irreducible, we must have
$D = 26$
or
$52$
by [Reference LübeckLu].
Suppose
$D=26$
. In case (a), we have
$(q^n \pm 1)/2 = D = 26$
, and so
$G_1 = \mathrm {SL}_2(53)$
. But
$\mathrm {SL}_2(53)$
cannot be embedded in
$F_4$
by [Reference Griess and RybaGrR]. In case (b),
$D \geq (q^n-q)/(q+1)$
is at least
$42$
when
$n \geq 7$
, or
$n=5$
but
$q \geq 3$
, or
$n=3$
but
$q \geq 7$
. Now if
$(n,q) = (5,2)$
, then
$D \in \{10,11,21,32\}$
. If
$(n,q) = (3,5)$
, then
$D \in \{20,21,\geq 41\}$
. If
$(n,q) = (3,4)$
, then
$D \in \{12,13,25,\geq 38\}$
. If
$(n,q) = (3,3)$
, then
$D \in \{6,7\}$
by hypothesis.
Suppose
$D=52$
. In case (a), we have
$(q^n \pm 1)/2 = D = 52$
, and so
$G_1 = \mathrm {SL}_2(103)$
. But
$\mathrm {SL}_2(103)$
cannot be embedded in
$F_4$
by [Reference Griess and RybaGrR]. In case (b),
$D \geq (q^n-q)/(q+1)$
is at least
$56$
if
$n \geq 9$
, or
$n=7$
but
$q \geq 3$
, or
$n = 5$
but
$q \geq 3$
, or
$n=3$
but
$q \geq 8$
. If
$(n,q) = (7,2)$
, then
$D \in \{42,43,\geq 85\}$
. If
$(n,q) = (5,2)$
or
$(3,3)$
, then
$D \leq 32$
. If
$(n,q) = (3,7)$
, then
$D \in \{42,43,\geq 85\}$
. If
$(n,q) = (3,5)$
, then
$D \in \{20,21,41,\geq 62\}$
by hypothesis. If
$(n,q) = (3,4)$
, then by hypothesis,
$D \in \{12,13,25,38,51,64\}$
.
(B7) Suppose
$G^\circ = E_6$
. Then
$G^\circ $
admits an irreducible projective representation of degree
$27$
, and so

whence
$D \leq 265$
by (6.8.6). Since
$V|_{G^\circ }$
is irreducible, we must have
$D = 27$
or
$78$
by [Reference LübeckLu].
In case (a),
$D=q^n$
or
$(q^n \pm 1)/2$
, so
$G_1 = \mathrm {SL}_2(27)$
as recorded in (iv), or
$\mathrm {SL}_2(53)$
, which cannot be projectively embedded in
$E_6$
by [Reference Griess and RybaGrR]. So we are in case (b). Since no
$\mathrm {PSU}_n(q)$
with
$n \geq 5$
, or
$n=3$
but
$q \geq 9$
, can be embedded in
$E_8$
by [Reference Griess and RybaGrR], we have
$n=3$
and
$q \leq 8$
. If
$(n,q) = (3,8)$
, then by hypothesis,
$D \in \{56,57\}$
or
$D \geq 113\}$
or
$170 \leq D \leq 512$
. If
$(n,q) = (3,7)$
, then by hypothesis,
$D \in \{42,43, \geq 85\}$
. If
$(n,q) = (3,5)$
, then
$D \in \{20,21,41,62, \geq 83\}$
by hypothesis. If
$(n,q) = (3,4)$
, then by hypothesis,
$D \in \{12,13,25\}$
or
$38 \leq D \leq 64$
.
(B8) Suppose
$G^\circ = E_7$
. Then
$G^\circ $
admits an irreducible projective representation of degree
$56$
, and so

whence
$D \leq 1140$
by (6.8.6). Since
$V|_{G^\circ }$
is irreducible, we must have
$D = 56$
,
$133$
, or
$912$
by [Reference LübeckLu].
In case (a),
$D=q^n$
or
$(q^n \pm 1)/2$
, so
$G_1 = \mathrm {SL}_2(113)$
or
$\mathrm {SL}_2(1813)$
, neither of which can be projectively embedded in
$E_7$
by [Reference Griess and RybaGrR]. So we are in case (b). Since no
$\mathrm {PSU}_n(q)$
with
$n \geq 5$
, or
$n=3$
but
$q \geq 9$
, can be embedded in
$E_8$
by [Reference Griess and RybaGrR], we have
$n=3$
and
$q \leq 8$
. If
$(n,q) = (3,8)$
, then by hypothesis,
$D \in \{56,57,113\}$
or
$170 \leq D \leq 512$
, so
$D=56$
and
$G_1=\mathrm {PSU}_3(8)$
as recorded in (v). If
$(n,q) = (3,7)$
, then by hypothesis,
$D \in \{42,43,85,128\}$
or
$171 \leq D \leq 343$
. If
$(n,q) = (3,5)$
, then
$D \leq 41$
or
$62 \leq D \leq 125$
by hypothesis. If
$(n,q) = (3,4)$
, then by hypothesis,
$D \leq 51$
or
$D=64$
.
(B9) Finally, suppose
$G^\circ = E_8$
. Then
$G^\circ < \mathrm {SL}_{248}$
, and so

whence
$D \leq 22365$
by (6.8.6). Since
$V|_{G^\circ }$
is irreducible, we must have
$D = 248$
or
$3875$
by [Reference LübeckLu].
In case (a),
$D=q^n$
or
$(q^n \pm 1)/2$
, which is impossible. So we are in case (b). Since no
$\mathrm {PSU}_n(q)$
with
$n \geq 5$
, or
$n=3$
but
$q \geq 9$
, can be embedded in
$E_8$
by [Reference Griess and RybaGrR], we have
$n=3$
and
$q \leq 8$
. Now if
$q\leq 5$
, then
$D \leq 125$
. If
$(n,q) = (3,8)$
, then by hypothesis,
$D \leq 227$
or
$284 \leq D \leq 512$
. If
$(n,q) = (3,7)$
, then by hypothesis,
$D \leq 214$
or
$257 \leq D \leq 343$
. None of these values can fit the values
$248$
or
$3875$
.
Now we can prove the main result of this section. Recall that the systems with finite
$G_{\mathrm {geom}}$
are already classified in Theorem 11.2.4 of [Reference Katz and TiepKT6].
Theorem 6.9. Consider the local system with
$r \geq 1$
subject to (4.0.1). Suppose that
$A \geq 7$
and that
$G_{\mathrm {geom},{\mathcal F}}$
is infinite. Then the following statements hold.
-
(i) Suppose that
$AB_1 \ldots B_r$ is even. Then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
-
(ii) Suppose that
$AB_1 \ldots B_r$ is odd. Then
$G_{\mathrm {geom},{\mathcal F}} =\mathrm {Sp}_{A-1}$ .
Proof. If
$r=1$
, then the result follows from Theorems 10.2.4 of 10.3.21 of [Reference Katz and TiepKT6]. Hence, we will assume
$r \geq 2$
. Since
$\gcd (A,B_1, \ldots ,B_r)=1$
, there must be some i such that

Now if has infinite monodromy group, then we are done by Theorem 6.7.
It remains to consider the case in which has finite geometric monodromy group
$G_1$
. By Theorem 5.4(ii), we may assume that

(b) Let the prime p be the characteristic of
${\mathcal F}$
, and consider the case where
$(A,B_r)=(p^n+1,1)$
for some
$n \in {\mathbb Z}_{\geq 1}$
. Then
${\mathcal F}$
is the system of Theorem 2.4, with its
$f(x)$
taken to be
$x^A$
, and its
$b_1,\ldots ,b_n$
taken to be
$\{B_1,\ldots ,B_{r-1},B_r=1\}$
, and
${\mathcal F}$
has

where
$\Sigma $
is the locus
$\Sigma _1 = \Sigma _A = \Sigma _{B_1} = \ldots = \Sigma _{B_{r-1}}=0$
.
According to [Reference Katz and TiepKT6, Theorem 11.2.3], when
$A=p^n+1$
, the local system
in characteristic
$p \nmid AB_1 \ldots B_r$
can have finite
$G_{\mathrm {geom}}$
only in the ‘van-der-Geer–van-der-Vlugt’ situations, that is, when
$B_j = p^{m_j}+1$
for
$1 \leq j \leq r-1$
, and either
$B_r=p^{m_r}+1$
with
$m_r \geq 0$
, or
$B_r = 1$
.
We apply Proposition 3.5 to
${\mathcal F}$
. Suppose
$p=2$
; in particular,
$2 \nmid AB_1 \ldots B_r$
. In the case of 3.5(b), we have
$M_{2,2}=3$
and hence (because
$G_{\mathrm {geom},{\mathcal F}}$
is infinite and
${\mathcal F}$
is symplectically self dual),
$G=\mathrm {Sp}_{A-1}$
. In the case of 3.5(d), which is ‘vdG-vdV’, G is finite.
Suppose
$p>2$
. In the case of 3.5(a), we have
$M_{2,2}=2$
, and hence (again because
$G_{\mathrm {geom},{\mathcal F}}$
is infinite),
$G^\circ = \mathrm {SL}_{A-1}$
. In the case of 3.5(c), which is again ‘vdG-vdV’, G is finite.
(c) In the rest of the proof, we will assume that

Moreover, using [Reference Katz and TiepKT6, Theorem 11.2.3], we may assume that some
$B_j$
is neither
$1$
nor a
$2$
-power plus one when
$p=2$
. Replacing
$B_i$
by this
$B_j$
, we may furthermore assume that

Let V denote the underlying representation of G, and apply Theorems 10.3.14 and 11.2.4 of [Reference Katz and TiepKT6] to
$G_1$
.
Suppose
$C> 1$
. Then, since C is odd, we are in case (iii) of [Reference Katz and TiepKT6, Theorem 11.2.4]. In particular,
$V|_{G_1}$
is a submodule of the total Weil module of
$\mathrm {SU}_n(q)$
that contains the submodule
of dimension
$(q^n-q)/(q+1)$
, for an odd integer
$n \geq 3$
and a power
$q=p^f$
. Furthermore,
$C|(q+1)$
by [Reference Katz and TiepKT6, Theorem 11.2.4(iii)], and this rules out the possibility
$(n,q) = (3,3)$
. If, moreover,
$(n,q)=(3,2)$
, then we have
$(A/C,B_i/C) = (3,1)$
, contrary to (6.9.1). Hence,
$(n,q) \neq (3,2)$
. Thus, we fulfill hypothesis (b) of Theorem 6.8.
Suppose
$C=1$
. Then [Reference Katz and TiepKT6, Theorem 11.2.4] implies that
is as described in Theorems 10.2.6 and 10.3.13 of [Reference Katz and TiepKT6]. Next, assumption (6.9.2) rules out possibility (iv) of [Reference Katz and TiepKT6, Theorem 10.2.6], and assumption (6.9.3) rules out possibility (iii) of [Reference Katz and TiepKT6, Theorem 10.3.13]. Furthermore, in case (ii) of [Reference Katz and TiepKT6, Theorem 10.2.6], we have
$(n,q) \neq (3,2)$
because
$A \geq 7$
. Thus,
$G_1$
satisfies the hypothesis of Theorem 6.8 when
$C=1$
as well, by [Reference Katz and TiepKT6, Theorems 10.2.7 and 10.3.13].
It follows that the semisimple group
$G^\circ =G_{\mathrm {geom},{\mathcal F}}^\circ $
satisfies one of the conclusions of Theorem 6.8. In particular,
$G^\circ> G_1$
acts irreducibly on V. Hence, by Proposition 6.4,
$V|_{G^\circ }$
is not self-dual in case (i). Next, we observe that none of the possibilities (ii) and (iv) of Theorem 6.8 cannot occur. Indeed, in the case of 6.8(ii), we have
$(D,p,G_1,C) = (32,2,\mathrm {SU}_5(2),3)$
. In such a case, by [Reference Katz and TiepKT6, Theorem 11.2.4],
$A=33$
and
$B_i \in \{3,9\}$
, which is forbidden by (6.9.3). In the case of 6.8(iv), we have
$(D,p,G_1,C) = (27,3,\mathrm {SL}_2(27),2)$
, which is ruled out since C is odd.
Suppose
$G^\circ $
satisfies Theorem 6.8(i), that is,
$G^\circ = \mathrm {SL}(V)$
,
$\mathrm {Sp}(V)$
, or
$\cong \mathrm {SO}(V)$
. In case (i),
$V|_{G^\circ }$
is not self-dual, so we must have
$G^\circ = \mathrm {SL}_D$
. In case (b), V is symplectic self-dual, so
$G^\circ = \mathrm {Sp}_D$
.
Suppose
$G^\circ $
satisfies Theorem 6.8(iii). Here,
$A=8$
, so
$V|_{G^\circ }$
is not self-dual, contradicting the fact that the
$7$
-dimensional module of
$G_2$
is self-dual.
Finally, we consider the case when
$G^\circ $
satisfies Theorem 6.8(v). Then we have
$A=57$
,
$C=1$
,
$G_1 = \mathrm {PSU}_3(8)$
, and Theorems 10.2.4 and 10.3.21 of [Reference Katz and TiepKT6] imply that
$p=2$
and
$B_i=1$
(and so
$i=r$
). As
$p=2$
, all
$B_j$
are odd, and hence,
$G \leq \mathrm {Sp}_{56}$
. We will derive a contradiction by showing that
$G=\mathrm {Sp}_{56}$
in this case. Indeed, recalling
$r \geq 2$
, we have that
$B_1> 1=B_i$
and
$\gcd (A,B_1)$
is odd. Replacing
$(A,B_i)$
by
$(A,B_1)$
, we have
$G^\circ = \mathrm {Sp}_{56}$
by the already established result.
Finally, we remove the restriction
$A \geq 7$
in Theorem 6.9:
Theorem 6.10. Consider the local system with
$r \geq 1$
subject to (4.0.1). Suppose that
$3 \leq A \leq 6$
and that
$G_{\mathrm {geom},{\mathcal F}}$
is infinite. Then the following statements hold.
-
(i) Suppose that
$AB_1 \ldots B_r$ is even. Then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_{A-1}$ .
-
(ii) Suppose that
$AB_1 \ldots B_r$ is odd. Then
$G_{\mathrm {geom},{\mathcal F}} =\mathrm {Sp}_{A-1}$ .
Proof. Denote
$G:=G_{\mathrm {geom},{\mathcal F}}$
. If
$r=1$
, then the result follows from Theorems 10.2.4 and 10.3.21 of [Reference Katz and TiepKT6]. Also, if
$A=3$
, then
$G^\circ \leq \mathrm {GL}_2$
is a semisimple algebraic group, whence
$G^\circ = \mathrm {SL}_2$
. Henceforth, we assume

Suppose that for some i, we have
$A=2B_i$
or
$A = 3B_i$
. In the latter case, we have
$A = 6$
because of (6.10.1). Hence, Theorem 3.7 implies that
$M_{2,2}=2$
in both cases, when
$G^\circ = \mathrm {SL}_{A-1}$
by Larsen’s Alternative.
Next, suppose that
$B_1=A-1$
; in particular,
$2|AB_1$
. If
has infinite geometric monodromy group H, then
$H^\circ = \mathrm {SL}_{A-1}$
by [Reference Katz and TiepKT6, Theorem 10.3.21(i)]. As
is a pullback of
${\mathcal F}$
, it follows that
$G^\circ =\mathrm {SL}_{A-1}$
. If
has finite geometric monodromy group, then
$(A,B_1,p) = (5,4,3)$
by [Reference Katz and TiepKT6, Theorem 10.3.13]. In this case, we have
$\{B_2, \ldots ,B_r\} \subseteq \{1,2\}$
, and G is finite by [Reference Katz and TiepKT6, Theorem 11.2.3(vii)].
We now assume that

and analyze the remaining cases.
(a) Suppose
$A = 4$
. Then
$B_i \neq 2,3$
and so
$k< 2$
, contrary to (6.10.1).
(b) Suppose
$A=5$
. First, we consider the case that
$B_i=2$
for some i, and let H denote the geometric monodromy group of
. If H is infinite, then
$H^\circ = \mathrm {SL}_{A-1}$
by [Reference Katz and TiepKT6, Theorem 10.3.21(i)], whence
$G^\circ = \mathrm {SL}_{A-1}$
. If H is finite, then
$(H,p) = (\mathrm {Sp}_4(3),3)$
or
$(2\mathsf {A}_7,7)$
by [Reference Katz and TiepKT6, Theorem 10.3.13]. In both cases, we have
$M_{2,2}(H)=2$
, whence
$M_{2,2}({\mathcal F})=2$
and
$G^\circ = \mathrm {SL}_4$
.
In the remaining case,
$B_i \neq 2,4$
, so
$k \geq 2$
forces
$(A,B_1, \ldots ,B_r) = (5,3,1)$
and
$p \neq 3,5$
. If
$p=2$
, then G is finite by [Reference Katz and TiepKT6, Theorem 11.2.3(ii)]. If
$p> 5$
, then
has
$\mathrm {Sp}_4$
as its
$G_{\mathrm {geom}}$
by [Reference Katz and TiepKT6, Lemma 10.3.20]. As
${\mathcal F}$
is symplectic self-dual, we conclude that
$G=\mathrm {Sp}_4$
.
(c) Finally, assume
$A=6$
. Then
$B_i \neq 2,3,5$
, so the inequality
$k \geq 2$
forces
$(A,B_1, \ldots ,B_r) = (6,4,1)$
and
$p \geq 5$
. Let H denote the geometric monodromy group of
. If in addition,
$p \neq 5,11$
, then H is infinite by [Reference Katz and TiepKT6, Theorem 10.2.6], whence
$H=\mathrm {SL}_5$
by [Reference Katz and TiepKT6, Theorem 10.2.4(i)], and we conclude that
$G^\circ = \mathrm {SL}_5$
.
Suppose
$p=5$
. Then
$M_{2,2}=2$
by Proposition 3.5, whence
$G^\circ = \mathrm {SL}_5$
.
In the remaining case, we have
$p=11$
. Recall that G acts irreducibly on the underlying representation
$V={\mathbb C}^5$
, of prime dimension
$D=5$
. Since
$G^\circ \neq 1$
is semisimple – in particular, non-abelian – it must have some simple submodule of dimension
$>1$
on V, and so Clifford’s theorem implies that
$G^\circ $
is irreducible on V as well. Now
$D=5$
being prime forces
$G^\circ $
to be simple, of rank
$\leq 4$
. An inspection of [Reference LübeckLu] or use of Gabber’s theorem [Reference KatzKa2, 1.8] shows that either
$G^\circ = \mathrm {SL}_5$
, or
$G^\circ = \mathrm {SL}_2$
, or
$G^\circ = \mathrm {SO}_5$
. In the latter two cases,
$V|_{G^\circ }$
is self-dual. Note that
$G/{\mathbf C}_G(G^\circ ) \hookrightarrow \mathrm {Out}(G^\circ )=1$
and
${\mathbf C}_G(G^\circ )$
is abelian by Schur’s lemma. So
$G/G^\circ $
is abelian, and hence, the simple
$H = \mathrm {PSL}_2(11)$
must embed in
$G^\circ $
. But this is a contradiction since
$V|_H$
is non-self-dual. Hence, we conclude that
$G^\circ = \mathrm {SL}_5$
. (Alternatively, by considering the pullback
of
${\mathcal F}$
and its decomposition as
, we see by that
projects onto
$\mathrm {SL}_3$
. This rules out the possibilities
$\mathrm {SL}_2$
and
$\mathrm {SO}_5$
for
$G^\circ $
).
7 Multiparameter local systems with infinite monodromy. II
In this section, we are given a (possibly trivial) multiplicative character
$\chi $
of (the multiplicative group of) a finite extension
$L/{\mathbb F}_p$
. We consider a local system
${\mathcal F}_\chi $
on
${\mathbb A}^r/L$
defined as follows. We are given a list of integers

as in (4.0.1). For
$E/L$
a finite extension, and
$(t_1,\ldots , t_r) \in E^r$
,

Here, we make a choice of
$\sqrt {p} \in \overline {{\mathbb Q}_\ell }$
, and define
$\sqrt {\# L}:=\sqrt {p}^{\deg (L/{\mathbb F}_p)}$
. We adopt the usual convention that
$\chi (0)=0$
if
, but
. We will name this
${\mathcal F}_\chi $
as

when confusion about which
${\mathcal F}_\chi $
?’ is possible. Recall from [Reference Katz and TiepKT5, 2.6] that such an
${\mathcal F}_\chi $
is geometrically irreducible.
In the previous sections, we determined
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ $
for any
whose
$G_{\mathrm {geom}}$
is infinite. We now do the same for any
${\mathcal F}_\chi $
with
whose
$G_{\mathrm {geom},{\mathcal F}_\chi }$
is infinite.
We begin with the ‘easy’ case.
Theorem 7.1. Let
$\chi $
be nontrivial. Suppose that for given data

with
$A \ge 3$
,
$r \geq 2$
, and both
,
${\mathcal F}_\chi $
have infinite
$G_{\mathrm {geom}}$
. Then we have the following results.
-
(i) If
$A\prod _iB_i$ is even, then
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$ .
-
(ii) If
$A\prod _iB_i$ is odd,
$p \neq 2$ , and
$\chi $ is the quadratic character, then we have
$G_{\mathrm {geom},{\mathcal F}_\chi }=\mathrm {SO}_A$ .
-
(iii) If
$A\prod _iB_i$ is odd, and
, then
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$ .
Proof. If
$A\prod _iB_i$
is even and
is infinite, then
by Theorems 6.9 and 6.10. Therefore,
. By Theorem 2.4, we have
. But for any local system of rank
$>1$
,
$M_{2,2}\ge 2$
. Therefore,
$M_{2,2}({\mathcal F}_\chi ) =2$
. Given that
$G_{\mathrm {geom},{\mathcal F}_\chi }$
is infinite, we must have
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$
by Larsen’s Alternative [Reference KatzKa3, 1.1.6].
If
$A\prod _iB_i$
is odd and
is infinite, then
. Therefore,
. Therefore,
$M_{2,2}({\mathcal F}_\chi ) \le 3$
, so either
$M_{2,2}({\mathcal F}_\chi ) =2$
or
$M_{2,2}({\mathcal F}_\chi ) =3$
. If p is odd and
$\chi $
is the quadratic character, then
${\mathcal F}_\chi $
is orthogonally self-dual (being self-dual because its traces are real, and being geometrically irreducible of odd rank). Thus, we have an a priori inclusion
$G_{\mathrm {geom},{\mathcal F}_\chi } \le \mathrm {O}_A$
. Given that
$G_{\mathrm {geom},{\mathcal F}_\chi }$
is infinite, we must have
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SO}_A$
by Larsen’s Alternative [Reference KatzKa3, 1.1.6]. Thus, we have
$\mathrm {SO}_D \le G \le \mathrm {O}_D$
. But
$\det ({\mathcal F})$
is lisse on
${\mathbb A}^k$
of order dividing
$2$
, and so must be geometrically trivial as
$p \neq 2$
.
Finally, we must treat the case when
$A\prod _iB_i$
is odd,
, and
. When
and A and all
$B_i$
are odd, we have
by Theorem 2.4. Therefore,
$M_{2,2}({\mathcal F}_\chi ) =2$
in this case, and we have
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$
by Larsen’s Alternative [Reference KatzKa3, 1.1.6].
It remains to treat cases with
$\chi $
nontrivial in which
has finite
$G_{\mathrm {geom}}$
but
${\mathcal F}_\chi $
has infinite
$G_{\mathrm {geom}}$
.
Theorem 7.2. Consider the case of p arbitrary,
$q=p^f$
for some
$f \ge 1$
,
$r \ge 2$
,

integers with
$\gcd (n,m_1,\ldots ,m_{r-1})=1$
and
${\mathcal F}_\chi $
,
, formed with

Then
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$
.
Proof. The only purpose of the
$\gcd $
hypothesis is to ensure that our choice of q is correct. The fact that
$B_r=1$
ensures the geometric irreducibility. We compute
as the number of geometrically irreducible components of dimension
$2$
of the intersection
$\Sigma _{A,B_1,\ldots ,B_r}$
of the Fermat surfaces

as d runs over the exponents
$(A,B_1,\ldots ,B_r)$
. We have the obvious inclusion

Using the equation
$\Sigma _1 =x+y-z-w=0$
, we may solve for w as
$w=x+y-z$
, and rewrite
$ \Sigma _{1+q^n,1}$
as the locus in the
${\mathbb A}^3$
of
$x,y,z$
of

Let us temporarily write

According to [Reference Katz and TiepKT6, Lemma 12.3.2], we have the following factorization in
${\mathbb F}_{Q^2}[x,y,z]$
:

In the special case
$p=2$
, we get the identity in
${\mathbb F}_{Q}[x,y,z]$

Going back to
$x,y,z,w$
, these linear factors give the following
$Q+1$
affine planes in
${\mathbb A}^4$
:

together with the
$Q-1$
planes
$P_A$
, one for each
$A \in {\mathbb F}_{Q^2}$
with
$A^{Q-1}=-1$
, of equation

By [Reference Katz and TiepKT6, Theorem 11.2.3],
$G_{\mathrm {geom},{\mathcal F}_\chi }$
is infinite, so it suffices to show that
$M_{2,2}({\mathcal F}_\chi )=2$
. The geometrically irreducible components of
$\Sigma _{A,B_1,\ldots ,B_r}$
are then among the planes above. So it suffices to show that for each
$A \in {\mathbb F}_{Q^2}$
with
$A^{Q-1}=-1$
, the limsup over extensions E of L dies:

We readily calculate on
$P_A$
, with coordinates
$y,z$
,

This is an expression in the quantity

namely,

Thus,

So it suffices to show that for every
$A \in {\mathbb F}_{Q^2}$
with
$A^{Q-1}=-1$
, this sum is
$O(\sqrt {\#E})$
.
Suppose first that p is odd. Then
$Q+1$
is even, and hence, A is neither
$1$
nor
$-1$
. Then the local system

is lisse of rank one on
${\mathbb P}^1 \smallsetminus \{0, \infty , A/(A+1), (A-1)/A\}$
, extended by direct image across the missing points, at each of which the ramification is tame but nontrivial. Then by the usual Weil estimate, this sum has absolute value at most
$2\sqrt {\#E}$
.
Suppose next that
$p=2$
. Then for
$A \neq 1$
, the above argument gives the same bound
$2\sqrt {\#E}$
. In the case
$A=1$
, local system is just
${\mathcal L}_{\overline {\chi }(T^2)} ={\mathcal L}_{\overline {\chi }^2(T)}$
, But
$\chi $
, being nontrivial in characteristic
$2$
, has odd order, so
, and in this case, the sum vanishes.
Theorem 7.3. Suppose given
$r \ge 2$
and integers

with
$\gcd (n,m_1,\ldots ,m_r)=1$
and
$2|n\prod _im_i$
. Let p be a prime,
$q=p^f$
with
$f \ge 1$
,
$\kappa := \gcd (p-1,2)$
, and form the data

If
$p=2$
, make the further assumption that
$m_r \ge 1$
. Then for
${\mathcal F}:={\mathcal F}(A,B_1,\ldots ,B_r)$
and any
$\chi $
with
, we have
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$
.
Proof. Because
$\gcd (n,m_1,\ldots ,m_r)=1$
and
$2|n\prod _im_i$
, there is some
$m_i$
whose parity is different from that of n: that is, if n is even the
$\gcd $
condition forces some
$m_i$
to be odd, and if n is odd, the evenness condition forces some
$m_i$
to be even. Pick one such
$m:=m_i$
such that n and m have opposite parities.
Next, we show that

Indeed,
$e:=\gcd (n,m)$
is odd as n and m have different parities. Let
$k \in \{n,m\}$
be the one that is even. Then
$q^k \equiv 1 \ \ (\mathrm {mod} \ {4})$
when
$p>2$
, and so
$(q^k+1)/\kappa $
is always odd, and hence,
$2 \nmid d$
. Suppose
$d> 1$
, and let
$\ell> 2$
be any prime divisor of d. Then
$\ell $
divides

and so
$\ell |(q^e-1)$
or
$\ell |(q^e+1)$
. In the former case, as
$e|n$
and
$\ell>2$
, we have
$\ell |(q^n-1)/\kappa $
, and so
$(q^n+1)/\kappa \equiv 2/\kappa \ \ (\mathrm {mod} \ {\ell })$
, a contradiction. In the latter case, as
$2 \nmid e$
and
$2|k$
, we have
$k=2le$
for some
$l \in {\mathbb Z}_{\geq 1}$
. Now
$\ell |(q^{2e}-1)$
and
$(q^{2e}-1)|(q^{2le}-1)$
, so we again have
$(q^k+1)/\kappa \equiv 2/\kappa \ \ (\mathrm {mod} \ {\ell })$
, a contradiction.
By [Reference Katz and TiepKT6, Theorem 11.2.3],
$G_{\mathrm {geom},{\mathcal F}(A,B_i,\chi )}$
is infinite. Using (7.3.1) and applying Theorems 10.2.4 and 10.3.21 of [Reference Katz and TiepKT6], we obtain
$G_{\mathrm {geom},{\mathcal F}(A,B_i,\chi )}^\circ = \mathrm {SL}_A$
. Since
${\mathcal F}(A,B_i,\chi )$
is a pullback of
${\mathcal F}_\chi $
, we conclude that
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ = \mathrm {SL}_A$
.
We now begin preparation for the
$\mathrm {SU}$
case. We begin with an ‘axiomatic’ result, which reveals the simple underlying mechanism.
Theorem 7.4. Let p be a prime
$A> B \ge 1$
a pair of odd, prime to p integers,
$C:=\gcd (A,B)$
. Write
$(A,B);=(A_0C,B_0C)$
. Suppose that
$\chi $
is a multiplicative character with
, with the following property: For every multiplicative character
$\rho $
with
$\rho ^C=\chi $
, the local system
${\mathcal F}(A_0,B_0,\rho )$
has infinite
$G_{\mathrm {geom}}$
. [Indeed, it has
$G_{\mathrm {geom}} =\mathrm {SL}_{A_0}$
, in view of Theorems 10.2.4 and 10.3.21 of [Reference Katz and TiepKT6]. Notice that
$A_0,B_0$
are both odd, so
$A_0-B_0 \ge 2$
.] Then the local system

has

Proof. For each
$\rho $
, pick a multiplicative character
$\sigma _\rho $
with

Then
$ \bigoplus _{\rho :\rho ^C=\chi }{\mathcal F}(A_0,B_0,\rho )$
is geometrically isomorphic to the Kummer
$[A_0]^\star $
pullback of the direct sum of hypergeometric sheaves

Each constituent hypergeometric sheaf is of type
$(A_0,B_0)$
, of odd rank
$A_0 \ge 3$
. As Kummer pullback does not change
$G_{\mathrm {geom}}^\circ $
, we see that each constituent hypergeometric sheaf has its
$G_{\mathrm {geom},{\mathcal H}}^\circ =\mathrm {SL}_{A_0}$
. So it suffices to show that

For this, we apply Goursat-Kolchin-Ribet in the form [Reference KatzKa2, 8.11.7.2]. We must show that for
$\rho _1 \neq \rho _2$
, there is no Kummer sheaf
${\mathcal L}_\Lambda $
such that the sheaf
${\mathcal L}_\Lambda \otimes {\mathcal H}_{big,A_0,B_0,\sigma _{\rho _1}}$
is geometrically isomorphic to either
${\mathcal H}_{big,A_0,B_0,\sigma _{\rho _2}}$
or to its dual. [Notice that because
$A_0-B_0$
is even, the dual of
$ {\mathcal H}_{big,A_0,B_0,\sigma _{\rho _2}}$
is (with the same
$\psi $
) geometrically isomorphic to
$ {\mathcal H}_{big,A_0,B_0,\overline {\sigma _{\rho _2}}}$
.]
We argue by contradiction. Suppose that

Looking at the
$I(0)$
-representations of the two hypergeometrics, which are each
$\mathsf {Char}(A_0)$
, we first conclude that
. From the definition of
${\mathcal H}_{big,A_0,B_0,\sigma _{\rho _1}}$
(cf. [Reference Katz and TiepKT4, §3]), we see that

So if the purported isomorphism holds, then
$\sigma _{\rho _1}/\Lambda ^{B_0}=\sigma _{\rho _2}$
. But their
$A_0$
powers are
$\rho _1$
and
$\rho _2$
, respectively, (because
). But
$\rho _1 \neq \rho _2$
, the desired contradiction.
If instead we have

then we get the equality
$\sigma _{\rho _1}/\Lambda ^{B_0}=\overline {\sigma _{\rho _2}}.$
But their
$A_0$
powers are
$\rho _1$
and
$\overline {\rho _2}$
. These cannot be equal because their C powers are
$\chi $
and
$\overline {\chi }$
, respectively, which are not equal, precisely because
.
With this ‘axiomatic’ result in hand, we now turn to the
$\mathrm {SU}$
case directly. In preparation, observe that for any prime power
$q>1$
and any odd integer
$n \ge 1$
, the ratio
$(q^n+1)/(q+1)$
is odd; indeed for
$n \ge 3$
, it is
$1$
mod
$q(q-1)$
.
Proposition 7.5. Let p be a prime,
$q=p^f$
with
$f \ge 1$
,
$r \ge 2$
, and

a sequence of odd integers with
$\gcd (n,m_1,\ldots ,m_r)=1$
. Define

Consider
${\mathcal F}:={\mathcal F}(A,B_1,\ldots ,B_r,\chi )$
where
. We have the following results.
-
(i) If
$\gcd (n,m_i)=1$ for some i, then
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_A$ .
-
(ii) In general, with
$c:=\gcd (n,m_i)$ and
$C:=(q^c+1)/(q+1)$ , we have
$$ \begin{align*}G_{\mathrm{geom},{\mathcal F}(A,B_i,\chi)} =\prod_{\rho:\,\rho^C =\chi}\mathrm{SL}_{A/C}.\end{align*} $$
$G_{\mathrm {geom},{\mathcal F}(A,B_i,\chi )}$ acts on
${\mathcal F}$ with C simple summands, none of which is self-dual and any two of which are neither isomorphic nor dual to each other.
Proof. The first assertion is easy since
$\gcd (A,B_i)=1$
, and so already the pullback
${\mathcal F}(A,B_i, \chi )$
has
$G_{\mathrm {geom}}=\mathrm {SL}_A$
by Theorems 10.2.4 and 10.3.21 of [Reference Katz and TiepKT6]. For the second assertion, with
$c=\gcd (n,m_i)$
and
$Q=q^c$
, we have

where

It remains only to remark that if and
$\rho ^C =\chi $
, then
. Hence,
${\mathcal F}(A_0,B_0,\rho )$
indeed has infinite
$G_{\mathrm {geom}}$
. Now if
$p> 2$
, then
$2|(q+1)$
and so
, and if
$p = 2$
, then
implies
. The formula for
$G_{\mathrm {geom},{\mathcal F}(A,B_i,\chi )}$
then follows from Theorem 7.4. The last statement also follows since each of the C simple summands
${\mathcal F}(A_0,B_0,\rho )$
is acted on by exactly one of the C simple factors
$\mathrm {SL}_{A_0}$
as on its natural module (or its dual), and
$A_0 \geq Q^2-Q+1 \geq 3$
.
Now we can complete the
$\mathrm {SU}$
case:
Theorem 7.6. Let p be a prime,
$q=p^f$
with
$f \ge 1$
,
$r \ge 2$
, and

a sequence of odd integers with
$\gcd (n,m_1,\ldots ,m_r)=1$
. Define

Consider
${\mathcal F}:={\mathcal F}(A,B_1,\ldots ,B_r,\chi )$
with any
$\chi $
where
. Then
${\mathcal F}$
has
$G_{\mathrm {geom},{\mathcal F}}^\circ =\mathrm {SL}_A$
.
Proof. (a) Let
$G :=G_{\mathrm {geom},{\mathcal F}}$
. If there is some i such that
$\gcd (n,m_i)=1$
, then we are done by Proposition 7.5(i). Hence, we may assume that

for all i. Since
$r \geq 2$
and
$2 \nmid nm_1 \ldots m_r$
, this implies that

for
$c:=c_r$
. (Indeed, if
$n < 15$
, then either n is a prime or
$n=9$
. In the former case,
$\gcd (n,m_1) = 1$
, and in the latter case,
$\gcd (n,m_2)=1$
, both violating (7.6.1).)
We know by [Reference Katz and TiepKT5, 2.6] that
${\mathcal F}$
is geometrically irreducible (i.e., that G is an irreducible subgroup of
$\mathrm {GL}_{A}=\mathrm {GL}(V)$
with
$V:={\mathcal F}_{\overline {\eta }}$
). By Proposition 7.5(ii), for each
$1 \leq i \leq r$
, G contains a semisimple subgroup

of rank

In particular,

Furthermore, the
$H_i$
-module V is a direct sum of
$(q^{c_i}+1)/(q+1)$
pairwise non-isomorphic simple summands, all of dimension

(b) Because
$G^\circ \lhd G$
, by Clifford’s theorem, we may express
$V|_{G^\circ } =e(\oplus ^m_{j=1} W_j)$
as the sum of e copies each of pairwise non-isomorphic simple summands
$W_1, \ldots ,W_m$
. Note that
$G^\circ \geq H_i$
for all i. Now if
$e> 1$
, then some simple summand of
$V|_{H_k}$
has multiplicity
$\geq e$
, contradicting the discussion in (a). Hence,
$e=1$
.
Next, the summands
$W_j$
are transitively permuted by G, so all have the same dimension

Since
$G^\circ \geq H_i$
and all simple summands of
$V|_{H_i}$
have the same dimension
$D_i$
, we must have that
$D_i|M$
; equivalently,
$(q^{c_i}+1)/m(q+1) \in {\mathbb Z}$
for all i. In turn, this implies that
$m(q+1)$
divides

where
$e := \gcd (c_1,c_2, \ldots ,c_r)$
. As e divides
$c_i=\gcd (n,m_i)$
, we have e divides n and each
$m_i$
, and thus,
$e|\gcd (n,m_1, \ldots ,m_r)=1$
. Thus,
$e=1$
and so
$m=1$
. We have shown that
$G^\circ $
acts irreducibly on V.
(c) Recall from (7.6.3) that the semisimple group
$G^\circ $
has rank
$R \geq R_r> 2A/3$
. As shown in (b),
$G^\circ $
acts irreducibly on V of dimension
$A < 3R/2$
, and

by (7.6.2). Arguing as in part (a1) of the proof of Theorem 6.6, we conclude that
$G^\circ $
is simple. Arguing as in part (b) of the proof of Theorem 6.6, we then see that
$G^\circ = \mathrm {SL}(V)$
,
$\mathrm {Sp}(V)$
, or
$\mathrm {SO}(V)$
. In the two latter cases, the
$G^\circ $
-module V is self-dual. Restricting to
$H_r$
, we see that some simple summand of the
$H_r$
-module V is either self-dual, or dual to another simple summand. This is, however, impossible by Proposition 7.5(ii). Hence,
$G^\circ = \mathrm {SL}_A$
.
Now we consider the remaining cases of an
${\mathcal F}_\chi $
on
${\mathbb A}^r/\overline {{\mathbb F}_p}$
with finite
and with
$r \geq 2$
. These remaining cases are listed in [Reference Katz and TiepKT6, Theorem 11.2.3]. They are
-
(i)
$p=2$ ,
$r=2$ ,
$A=13$ ,
$B_1=3$ ,
$B_2=1$ , and
$G = 2 \cdot G_2(4)$ .
-
(ii)
$p=3$ ,
$r=2,3$ ,
$A=7$ ,
$\{B_1, \ldots ,B_r\} \subseteq \{4,2,1\}$ , and
$G = 6_1 \cdot \mathrm {PSU}_4(3)$ .
-
(iii)
$p=3$ ,
$r=2,3$ ,
$A=5$ ,
$\{B_1, \ldots ,B_r\} \subseteq \{4,2,1\}$ . Furthermore,
$G = \mathrm {Sp}_4(3) \times 3$ if some
$B_i$ is
$4$ , and
$G=\mathrm {Sp}_4(3)$ otherwise.
-
(iv)
$p=5$ ,
$A=3$ ,
$B_1=2$ ,
$B_2=1$ , and
$G = \mathrm {SL}_2(5) \times 5$ .
Each of these cases, with the exception of
${\mathcal F}(5,2,1)$
in characteristic
$p=3$
, has the following property: for any
,
${\mathcal F}_\chi $
has infinite
$G_{\mathrm {geom}}$
. This is immediate from [Reference Katz and TiepKT6, Theorem 11.2.3], which lists all cases of an
${\mathcal F}_\chi $
with finite
$G_{\mathrm {geom}}$
. In the exceptional case of
${\mathcal F}:={\mathcal F}(5,2,1)$
in characteristic
$p=3$
, we have a Weil representation of degree
$4$
of
$\mathrm {Sp}_4(3)$
. In this case,
${\mathcal F}_{\chi _2}$
yields a Weil representation of degree
$5$
of
$\mathrm {PSp}_4(3)$
.
Theorem 7.7. For any of the
${\mathcal F}$
listed above other than
${\mathcal F}(5,2,1)$
in characteristic
$p=3$
, and any
,
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$
. In the exceptional case of
${\mathcal F}(5,2,1)$
in characteristic
$p=3$
, the same is true for any
$\chi $
with
.
Proof. In cases (ii)–(iv), has
$M_{2,2}=2$
, whence the same holds for
${\mathcal F}_\chi $
. As
$G_{\mathrm {geom},{\mathcal F}_\chi }$
is infinite by the discussion preceding the theorem, we conclude
$G_{\mathrm {geom},{\mathcal F}_\chi }^\circ =\mathrm {SL}_A$
. In case (i), the pullback
${\mathcal F}(13,3,\chi )$
has
$\mathrm {SL}_{13}$
as its
$G_{\mathrm {geom}}$
by Theorems 10.3.13 and 10.3.21 of [Reference Katz and TiepKT6], so we are done again.
Now we can prove the first main result of the paper. Recall that local systems
${\mathcal F}(A,B_1,\ldots ,B_r,\chi )$
with finite
$G_{\mathrm {geom}}$
(and the corresponding
$G_{\mathrm {geom}}$
) have been determined in [Reference Katz and TiepKT6, Theorem 11.2.3].
Theorem 7.8. Consider the local system
${\mathcal F}_\chi :={\mathcal F}(A,B_1,\ldots ,B_r,\chi )$
over
${\mathbb A}^r/\overline {{\mathbb F}_p}$
with
$r \geq 1$
subject to (4.0.1), of dimension
$D=A-1$
if
, and
$D=A$
otherwise. Suppose that
$D \geq 2$
and that
$G:=G_{\mathrm {geom},{\mathcal F}_\chi }$
is infinite. Then the following statements hold.
-
(i) If
$AB_1 \ldots B_r$ is even, then
$G^\circ =\mathrm {SL}_D$ .
-
(ii) If
$AB_1\ldots B_r$ is odd and
, then
$G^\circ =\mathrm {SL}_D$ .
-
(iii) If
$AB_1 \ldots B_r$ is odd and
, then
$G =\mathrm {Sp}_D$ .
-
(iv) Suppose
$AB_1 \ldots B_r$ is odd,
$p \neq 2$ , and
$\chi = \chi _2$ . Then
$G =\mathrm {SO}_D$ , unless
$(r,A,B_r) = (1,7,1)$ , in which case we have
$G =G_2$ .
Proof. If
$r=1$
and
$A \geq 3$
, then the result follows from Theorems 10.2.4 of 10.3.21 of [Reference Katz and TiepKT6]. If
$A=2$
, then
${\mathcal F} = {\mathcal F}(2,1,\chi )$
, and
$1 \neq G^\circ \leq \mathrm {GL}_2$
is semisimple, so
$G^\circ = \mathrm {SL}_2$
.
We next treat the cases
$r \geq 2$
when
has infinite
$G_{\mathrm {geom}}$
. These cases result from Theorem 7.1.
Finally, assume that
$r \geq 2$
and
has finite
$G_{\mathrm {geom}}$
. Applying [Reference Katz and TiepKT6, Theorem 11.2.3], we arrive at one of the possibilities considered in Theorems 7.2, 7.3, 7.6, and 7.7.
We now consider the following variant. Given a finite field L of characteristic p, a multiplicative character
$\chi $
of
$L^\times $
, and data
$(A,B_1, \ldots ,B_r)$
subject to (4.0.1), we denote by
${\mathcal F}^\sharp (A,B_1,\ldots ,B_r,\chi )$
the local system on
$({\mathbb G}_m\times {\mathbb A}^r)/L$
whose trace function is given as follows:
For
$E/L$
a finite extension, and
$(s,t_1,\ldots ,t_r) \in E^\times \times E^r$
,

Theorem 7.9. Consider
${\mathcal F}^\sharp _\chi :={\mathcal F}^\sharp (A,B_1,\ldots ,B_r,\chi )$
with
$r \geq 1$
subject to (4.0.1), of dimension
$D=A-1$
if
, and
$D=A$
otherwise. Suppose that
$D \geq 2$
and that
$G:=G_{\mathrm {geom},{\mathcal F}^\sharp _\chi }$
is infinite. Then the following statements hold.
-
(i) If
$AB_1 \ldots B_r$ is even, then
$G^\circ =\mathrm {SL}_D$ .
-
(ii) If
$AB_1\ldots B_r$ is odd and
, then
$G^\circ =\mathrm {SL}_D$ .
-
(iii) If
$AB_1 \ldots B_r$ is odd and
, then
$G =\mathrm {Sp}_D$ .
-
(iv) Suppose
$AB_1 \ldots B_r$ is odd,
$p \neq 2$ , and
$\chi = \chi _2$ . Then
$G =\mathrm {O}_D$ , unless
$(r,A,B_r) = (1,7,1)$ , in which case we have
$G =\{\pm 1\}\times G_2$ .
Proof. We follow the idea behind [Reference Katz and TiepKT6, 8.5.1]. After the partial Kummer covering of
${\mathbb G}_m\times {\mathbb A}^r$
by itself,

the change of variable
$x \mapsto x/s$
, and the reparameterization
$s \mapsto s, t_i \mapsto t_is^{B_i}$
, this pullback is just (the restriction to
${\mathbb G}_m\times {\mathbb A}^r$
of) the external tensor product
$ {\mathcal L}_{\overline {\chi }(s)} \otimes {\mathcal F}(A,B_1,\ldots ,B_r,\chi )$
. Finite pullback does not change
$G^\circ $
, nor does tensoring with a Kummer sheaf of finite order. In the case when
and
$AB_1\ldots B_r$
is odd,
${\mathcal F}^\sharp _\chi $
is symplectic. So on the one hand, its
$G^\circ =\mathrm {Sp}_D$
, while we also have
$G \le \mathrm {Sp}_D$
. In the case when
$p \neq 2$
,
$\chi =\chi _2$
and
$AB_1\ldots B_r$
is odd,
${\mathcal F}^\sharp _\chi $
is orthogonal. So its
$G^\circ =\mathrm {SO}_D$
, while we also have
$G \le \mathrm {O}_D$
. However, after the partial Kummer pullback
$[A,\mathrm {Id}]^\star $
, we obtain
${\mathcal L}_{\chi _2}(s)\otimes {\mathcal F}(A,B_1,\ldots ,B_r,\chi _2)$
. Here,
${\mathcal F}_{\chi _2}$
has odd rank A and trivial determinant, so this
$[A,\mathrm {Id}]^\star {\mathcal F}^\sharp _{\chi _2}$
pullback has nontrivial determinant. Therefore,
${\mathcal F}^\sharp _{\chi _2}$
must have nontrivial determinant.
8
$M_{2,2}$
and finite symplectic and special unitary groups
In this section, we will determine the subgroups of
$G=\mathrm {Sp}_{2n}(q)$
with
$2 \nmid q$
, and
$G=\mathrm {SU}_n(q)$
with
$2 \nmid n$
, which have the same
$M_{2,2}$
on an irreducible Weil representation of G. These results will allow us to determine
$G_{\mathrm {geom}}$
for
${\mathcal F}(f,A,B)$
, as defined in (1.0.4), in §11.
Let p be any odd prime and
$q=p^f$
. Then
$G=\mathrm {Sp}_{2n}(q)$
has two total Weil representations of degree
$q^n$
, with characters
$\xi + \eta $
, and
$\xi ^* + \eta ^*$
, where
$\xi \in \mathrm {Irr}(G)$
has degree
$(q^n+1)/2$
,
$\eta \in \mathrm {Irr}(G)$
has degree
$(q^n-1)/2$
, and
$^*$
denotes the action of the outer automorphism of
$G_n$
induced by the conjugation by an element in
$\mathrm {CSp}_{2n}(q) \smallsetminus \mathrm {Sp}_{2n}(q){\mathbf Z}(\mathrm {CSp}_{2n}(q))$
; cf. [Reference Tiep and ZalesskiiTZ2], [Reference Katz and TiepKT1].
Theorem 8.1. [Reference Katz and TiepKT7, Theorem 2.1] Assume
$(n,q) \neq (1,3)$
. Then the following statements hold for any irreducible Weil character
$\theta =\xi ,\xi ^*,\eta ,\eta ^*$
of
$G=\mathrm {Sp}_{2n}(q)$
.
-
(i) If
$n \geq 2$ , or if
$n=1$ but
$\theta \in \{\xi ,\xi ^*\}$ , then
$$ \begin{align*}M_{2,2}(\theta) = \left\{\begin{array}{ll}(q+7)/4, & q \equiv 1 \ \ \ \ (\mathrm{mod} \ {4}),\\ (q+5)/4, & q \equiv 3 \ \ \ \ (\mathrm{mod} \ {4}).\end{array}\right.\end{align*} $$
-
(ii) If
$n=1$ but
$\theta \in \{\eta ,\eta ^*\}$ , then
$M_{2,2}$ drops by one; that is,
$$ \begin{align*}M_{2,2}(\theta) = \left\{\begin{array}{ll}(q+3)/4, & q \equiv 1 \ \ \ \ (\mathrm{mod} \ {4}),\\ (q+1)/4, & q \equiv 3 \ \ \ \ (\mathrm{mod} \ {4}).\end{array}\right.\end{align*} $$
Theorem 8.2. Let
$q=p^f$
be a power of an odd prime p,
$n \geq 1$
, and
$(n,q) \neq (1,3)$
. Let H be a subgroup of
$G=\mathrm {Sp}_{2n}(q)$
and
$\theta $
be an irreducible Weil character of G, and suppose that

Then either
$H=G$
, or one of the following cases occurs.
-
(i)
$(G,H,\theta (1))=(\mathrm {Sp}_2(5),\mathrm {SL}_2(3),2)$ .
-
(ii)
$(G,H,\theta (1)) = (\mathrm {Sp}_4(3),2^{1+4}_- \cdot \mathsf {A}_5,4)$ .
Proof. We argue by contradiction. If
$H < G$
, there exists a subgroup M with
$H \leq M <G$
and M a maximal subgroup of G. We will show that this leads to a contradiction except in the two specified exceptional cases. For brevity, in this proof,
$\boldsymbol {\langle } a \boldsymbol {\rangle }$
(or
$\boldsymbol {\langle } a\boldsymbol {\rangle }_i$
with some subscript i) will denote an irreducible character of G of degree
$a \in {\mathbb Z}_{\geq 1}$
. We will freely use the fact that the equality

implies that H, and so M, is irreducible on any irreducible constituent
$\alpha $
of the G-character
$\theta \bar \theta $
. Moreover,
$M_{1,1}(H,\theta )=M_{1,1}(G,\theta )=1$
by [Reference Guralnick and TiepGT2, Lemma 3.1], so
$\theta $
is irreducible over H and M as well.
(a) Here, we consider the case
$n =1$
. First, suppose that
$q=5$
. If
$\theta \in \{\xi ,\xi ^*\}$
, then

as one can check using [GAP]. This implies that
$\boldsymbol {\langle } 5 \boldsymbol {\rangle }$
is irreducible over M, and so
$|M| \geq 26$
, which is impossible by [Reference Conway, Curtis, Norton, Parker and WilsonAtlas]. If
$\theta \in \{\eta ,\eta ^*\}$
, then
$\theta \bar \theta = \boldsymbol {\langle } 1 \boldsymbol {\rangle }+\boldsymbol {\langle } 3 \boldsymbol {\rangle }$
by [GAP]. Then
$\boldsymbol {\langle } 3 \boldsymbol {\rangle }$
is irreducible on G, and so
$|G| \geq 10$
and
$3$
divides
$|G|$
, whence
$G \cong \mathrm {SL}_2(3)$
by [Reference Conway, Curtis, Norton, Parker and WilsonAtlas], as stated in (i).
Assume now that
$9 \leq q \equiv 1 \ \ (\mathrm {mod} \ {4})$
. Using the character table of G [Reference DornhoffDo], one can check that

Similarly, if
$7 \leq q \equiv 3 \ \ (\mathrm {mod} \ {4})$
, then

In both cases, some
$\boldsymbol {\langle } q+1 \boldsymbol {\rangle }$
is irreducible over M, and so
$|M|> (q+1)^2$
, which is impossible by [Reference Bray, Holt and Roney-DougalBHR, Tables 8.1, 8.2].
(b) From now on, we will assume
$n \geq 2$
. According to [Reference Katz and TiepKT7, formulas (2.1.7) and (2.1.11)], we have

when
$q \equiv 1 \ \ (\mathrm {mod} \ 4)$
, and

when
$q \equiv 3 \ \ (\mathrm {mod} \ {4})$
. In particular, there exist (not necessarily distinct) characters
$\gamma ,\delta \in \mathrm {Irr}(G)$
such that

Indeed, if
$\theta \in \{\eta ,\eta ^*\}$
, then we can choose

Suppose
$\theta \in \{\xi ,\xi ^*\}$
. Then we can choose
$\gamma = \theta $
and

Assume in addition that
$n = 2$
. Then (8.2.3) implies that
$q(q^2+1)/2$
divides
$|M|$
. Using [Reference Bray, Holt and Roney-DougalBHR, Tables 8.12, 8.13], we now see that either
$M = \mathrm {Sp}_2(q^2) \rtimes C_2$
, or
$q=3$
and
$M = 2^{1+4}_- \cdot \mathsf {A}_5$
. In the former case, the degree of any irreducible character of M has p-part equal to
$1$
or
$q^2$
, contrary to the existence of
$\delta $
in (8.2.3). In the latter case, suppose
$\theta = \xi $
. Then (8.2.2) shows that
$\boldsymbol {\langle } 24 \boldsymbol {\rangle }$
is irreducible on the image
$2^4 \cdot \mathsf {A}_5$
of M in
$G/\mathrm {Ker}(\theta )$
. Hence,
$24$
divides
$|\mathsf {A}_5|=60$
by Ito’s theorem [Reference IsaacsIs, (6.15)], a contradiction. Thus,
$\theta (1)=\eta (1)=4$
, and using [GAP], we can check that
$M_{2,2}(M,\theta )=3=M_{2,2}(G,\theta )$
. Now, as
${\mathbf O}_2(M)H$
is irreducible on
$\boldsymbol {\langle } 15 \boldsymbol {\rangle }$
,
${\mathbf O}_2(M)H/{\mathbf O}_2(M)$
is a subgroup of
$\mathsf {A}_5$
of order divisible by
$15$
, whence
$H{\mathbf O}_2(M) = M$
. Working in
$M/Z$
where
$Z := {\mathbf Z}({\mathbf O}_2(M)) \cong C_2$
and noting that
$\mathsf {A}_5$
is irreducible on
${\mathbf O}_2(M)/Z \cong {\mathbb F}_2^4$
, we see that either
$ZH=M$
or
$|ZH|=2|\mathsf {A}_5|=120$
. The latter is, however, impossible as H is irreducible on
$\boldsymbol {\langle } 15 \boldsymbol {\rangle }$
. So
$ZH=M$
, whence
$H \geq [ZH,ZH] \geq [{\mathbf O}_2(M),{\mathbf O}_2(M)]=Z$
. Thus,
$H=ZH=M$
, and we arrive at (ii).
Next, we consider the case
$G = \mathrm {Sp}_6(5)$
. By the choice of
$\gamma $
in (8.2.3),
$|M|$
is divisible by
$7 \cdot 31$
. Inspecting [Reference Bray, Holt and Roney-DougalBHR, Tables 8.28, 8.29], we see that
$M = \mathrm {Sp}_2(5^3) \rtimes C_3$
. In this case, the degree of any irreducible character of M has p-part equal to
$1$
or
$q^3$
, contrary to the existence of
$\delta $
in (8.2.3).
Assume now that
$G = \mathrm {Sp}_6(3)$
. As noted in the proof of (8.2.3) for
$q=3$
, both
$\theta $
and
$\theta \bar \theta -1_G$
are irreducible over M. This implies that
$|M|$
is divisible by
$7 \cdot 13$
and
$\mathrm {Irr}(M)$
contains a character of degree
$\geq 168$
. Inspecting [Reference Bray, Holt and Roney-DougalBHR, Tables 8.28, 8.29], we arrive at a contradiction.
(c) In the rest of the proof, we may assume that

Then
$p^{2nf}-1$
admits a large primitive prime divisor
$\ell $
in the sense of [Reference FeitF]. Note that
$Q:=(q^{2n}-1)_\ell $
divides
$(q^n+1)/2$
, and so Q divides
$|M|$
by (8.2.3). Now we can apply [Reference Katz and TiepKT1, Theorem 4.6] to the subgroup
$M < \mathrm {GL}_{2n}(q)$
. If in addition

is abelian, then again by Ito’s theorem the irreducibility of
$\gamma |_M$
implies that
$\gamma (1)$
divides
$|M/L|$
, and hence,
$\ell \nmid \gamma (1)=(q^n+1)/2$
, a contradiction. Hence, by [Reference Katz and TiepKT1, Theorem 4.6], there is a divisor
$j<2n$
of
$2n$
such that we are in one of the following cases for L.
(c1)
$j \leq 2n/3$
and
$L \cong \mathrm {SL}_{2n/j}(q^j)$
. Note that if
$q^j=3$
, then
$2n/j> 6$
by (8.2.4), and so
$L \not \cong \mathrm {PSL}_4(3)$
. Hence, the smallest degree
$d(L)$
of nontrivial irreducible complex characters of L satisfies

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1]. This forces the quasisimple subgroup L of G to be in the kernel of the Weil character
$\eta $
, which is absurd since
$\mathrm {Ker}(\eta ) \leq {\mathbf Z}(G) = C_2$
.
(c2)
$j|n$
,
$j \leq n/2$
, and
$L \cong \Omega ^-_{2n/j}(q^j)$
. Now if
$j \leq n/4$
, then

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1], which leads to the contradiction
$L \leq \mathrm {Ker}(\eta ) \leq C_2$
as in (c1).
Suppose
$j = n/3$
. Then
$(2n/j,q^j) \neq (6,3)$
by (8.2.4). Hence,
$L = \Omega ^-_6(q^j)$
is a cover of
$\mathrm {PSU}_4(q^j) \not \cong \mathrm {PSU}_4(3)$
, and so

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1]. This again yields the contradiction
$L \leq \mathrm {Ker}(\eta ) \leq C_2$
as in (c1).
In the remaining case
$j = n/2$
, we have
$L \cong \Omega ^-_4(q^{n/2}) \cong \mathrm {PSL}_2(q^n)$
; see [Reference Kleidman and LiebeckKlL2, Proposition 2.9.1(v)]. Now,
$d(\mathrm {PSL}_2(q^n)) = (q^n+1)/2> \eta (1)$
(as
$q^n> 27$
by (8.2.4)), and this again forces
$L \leq \mathrm {Ker}(\eta ) \leq C_2$
, a contradiction.
(c3)
$j|n$
,
$L \cong \mathrm {Sp}_{2n/j}(q^j)$
, and
$L \lhd M \leq {\mathbf N}_G(L) = L \rtimes C_j$
. Then we look at the character
$\delta $
in (8.2.3). First, suppose that
$\epsilon =-$
. As
$n \geq 3$
by (8.2.4),
$p^{(2n-2)f}-1$
has a primitive prime divisor
$\ell _1$
by [Reference ZsigmondyZs], and then
$\ell _1$
divides both
$\delta (1)$
and
$|M|$
. Note that
$\ell _1 \geq 2n-1> j$
, so in fact
$\ell _1$
divides
$|L|$
. Hence, we can find some
$1 \leq i \leq n/j$
such that
$\ell _1|(q^{2ij}-1)$
. The primitivity of
$\ell _1$
implies that
$(n-1)|ij$
, but
$ij \leq n < 2(n-1)$
. Thus,
$ij=n-1$
, and so
$j|\gcd (n,n-1)=1$
. We conclude that
$j=1$
and
$L = G$
, a contradiction.
Next, we consider the case
$\epsilon =+$
. As before,
$L < G$
implies that
$j> 1$
. Suppose first that
$j=n$
. Then
$\mathrm {Sp}_2(q^n) = L \lhd M \leq L \cdot C_n$
. It follows that the maximum degree of any
$\alpha \in \mathrm {Irr}(M)$
is at most

contrary to (8.2.3). So we have
$j < n$
; in particular,
$n \geq 4$
. Hence,
$p^{(n-1)f}-1$
has a primitive prime divisor
$\ell _2$
by [Reference ZsigmondyZs]. Now
$\ell _2$
divides both
$\delta (1)$
and
$|M|$
. Note that
$\ell _2 \geq n> j$
, so in fact,
$\ell _2$
divides
$|L|$
. Hence, we can find some
$1 \leq i \leq n/j$
such that
$\ell _2|(q^{2ij}-1)$
. The primitivity of
$\ell _2$
implies that
$(n-1)|2ij$
, but
$2ij \leq 2n < 3(n-1)$
. Thus,
$ij=(n-1)/2$
or
$n-1$
. It follows that
$j|\gcd (n,n-1)=1$
, and so
$j=1$
, again a contradiction.
(c4)
$j=2j_0 \in 2{\mathbb Z}$
,
$n/j_0 \geq 3$
is odd,
$L \cong \mathrm {SU}_{n/j_0}(q^{j_0})$
, and

First, suppose that
$\theta (1)=\xi (1)$
, and so
$\epsilon =-$
in (8.2.3). As
$n \geq 3$
by (8.2.4),
$p^{(2n-2)f}-1$
has a primitive prime divisor
$\ell _1$
by [Reference ZsigmondyZs], and then
$\ell _1$
divides both
$\delta (1)$
and
$|M|$
. Note that
$\ell _1 \geq 2n-1> j$
, so in fact,
$\ell _1$
divides
$|L|$
. Hence, we can find some
$1 \leq i \leq n/j_0$
such that
$\ell _1|(q^{ij_0}-(-1)^i)$
. The primitivity of
$\ell _1$
implies that
$2(n-1)|2ij_0$
(i.e.,
$(n-1)|ij_0$
). But
$ij_0 \leq n < 2(n-1)$
, so
$ij_0=n-1$
, and
$j_0|\gcd (n,n-1)=1$
. In this case,
$j_0=1$
, and
$i=n-1$
is even. Hence,
$\ell _1|(q^{ij_0}-1)$
, and so
$2n-2$
divides
$ij_0=n-1$
, a contradiction.
In the remaining case, we have
$\theta (1)=\eta (1)$
. Since
$L < \mathrm {Sp}_{2n/j_0}(q^{j_0})$
,
$\theta |_L$
is the restriction to L of a Weil character of degree
$(q^n-1)/2$
of
$\mathrm {Sp}_{2n/j_0}(q^{j_0})$
, and so it is a sum of the unipotent Weil character of degree
$(q^n-q^{j_0})/(q^{j_0}+1)$
and
$(q^{j_0}-1)/2$
irreducible Weil characters, each of degree
$(q^n+1)/(q^{j_0}+1)$
. Since these characters are not of the same degree,
$\theta |_M$
cannot be irreducible, a contradiction.
(c5)
$(p,nf) = (3,9)$
and
$L/{\mathbf Z}(L) = \mathrm {PSL}_2(37)$
. Here, since the smallest degree of nontrivial irreducible representations of L over
$\overline {{\mathbb F}_3}$
is
$\geq 18$
, we must have that
$G = \mathrm {Sp}_{18}(3)$
,
$L = \mathrm {SL}_2(37)=M$
. But then M cannot be irreducible on
$\theta $
of degree
$\geq (3^9-1)/2$
.
(c6)
$(p,nf) = (17,6)$
and
$L/{\mathbf Z}(L) = \mathrm {PSL}_2(13)$
. Here, since the smallest degree of nontrivial irreducible representations of L over
$\overline {{\mathbb F}_{17}}$
is
$\geq 6$
, we must have that
$G = \mathrm {Sp}_{6}(17)$
,
$L = \mathrm {SL}_2(13)=M$
. But then M cannot be irreducible on
$\theta $
of degree
$\geq (17^3-1)/2$
.
Now let p be any prime,
$q=p^f$
, and
$2 \nmid n \geq 5$
. Then
$G=\mathrm {SU}_{n}(q)$
has a total Weil representation of degree
$q^n$
, with character
$\omega _n=\sum ^q_{i=0}\zeta _{i,n}$
, where
$\zeta _{i,n} \in \mathrm {Irr}(G)$
has degree
$(q^n+q(-1)^n)/(q+1)$
when
$i=0$
and
$(q^n-(-1)^n)/(q+1)$
when
$1 \leq i \leq q$
see, for example, [Reference Tiep and ZalesskiiTZ2] and [Reference Katz and TiepKT2].
Theorem 8.3 [Reference Katz and TiepKT7, Theorem 3.4].
Assume
$2 \nmid n$
and
$n \geq 5$
. Then for the irreducible Weil character
$\theta =\zeta _{i,n}$
of
$\mathrm {SU}_{n}(q)$
, of degree
$(q^n-q)/(q+1)$
if
$i=0$
and
$(q^n+1)/(q+1)$
if
$1 \leq i \leq q$
, we have

Theorem 8.4. Let
$q=p^f$
be a power of a prime p,
$2 \nmid n \geq 3$
odd, and
$(n,q) \neq (3,2)$
. Let H be a subgroup of
$G=\mathrm {SU}_{n}(q)$
and
$\theta $
be an irreducible Weil character of G, and suppose that

Then
$H=G$
.
Proof. As in the proof of Theorem 8.2, we will assume that
$H < G$
and let
$H \leq M <G$
for a maximal subgroup M of G. We will also use the fact that
$M_{2,2}(H,\theta ) = M_{2,2}(G,\theta )$
implies that H, and so M, is irreducible on any irreducible constituent
$\alpha $
of the G-characters
$\theta ^2$
and
$\theta \bar \theta $
, as well as on
$\theta $
itself.
(a) Here, we consider the case
$n =3$
. First, suppose that
$q=3$
, respectively
$q=4$
. Using [GAP], we can check that
$\theta \bar \theta $
has an irreducible constituent
$\alpha $
with
$\alpha (1) \geq 21$
, respectively
$\alpha (1)=65$
. However,
$|M| \leq 216$
, respectively
$|M| \leq 960$
by [Reference Conway, Curtis, Norton, Parker and WilsonAtlas], so
$\alpha |_M$
is reducible, a contradiction.
Assume now that
$q \geq 5$
. First, we consider the case
$\theta (1)=q^2-q+1$
. Then
$\theta (1)$
is divisible by
$\ell $
, a primitive prime divisor of
$p^{6f}-1$
by [Reference ZsigmondyZs]. Using [Reference Bray, Holt and Roney-DougalBHR, Tables 8.5, 8.6], we see that
$|M|$
can be divisible by
$\ell $
only when
$M = C_{\gcd 3,(q+1)} \times \mathrm {PSL}_2(7)$
,
$3 \cdot \mathsf {A}_6$
,
$3 \cdot \mathsf {A}_6 \cdot 2_3$
, or
$q=5$
and
$M=3 \cdot \mathsf {A}_7$
. The first three cases are, however, impossible because M cannot have an irreducible character of odd degree
$q^2-q+1 \geq 21$
. In the last case,
$\theta \bar \theta $
contains an irreducible constituent
$\alpha $
of degree
$126$
, and hence,
$\alpha $
is reducible over M by [Reference Conway, Curtis, Norton, Parker and WilsonAtlas].
It remains to consider the case
$\theta (1)=q^2-q$
. Then
$\theta |_M$
is irreducible; in particular,
$|M|> q^2(q-1)^2$
. Again using [Reference Bray, Holt and Roney-DougalBHR, Tables 8.5, 8.6], we can check that M must be a Borel subgroup of G. Note that the degree of any irreducible character of M is then equal to
$1$
or divisible by a fixed prime divisor r of
$(q-1)/\gcd (3,q-1)$
[Reference GeckGeck]. However, any irreducible constituent of
$\theta \bar \theta -1_G$
has degree
$>1$
, and at least one of them, say
$\beta $
, has degree coprime to r. Thus,
$\beta |_M$
is reducible, again a contradiction.
(b) From now on, we may assume
$n \geq 5$
, and write
$\theta =\zeta _{i,n}$
with
$0 \leq i \leq q$
. Then the proof of Theorem 8.3 in [Reference Katz and TiepKT7] shows that
$\theta ^2$
has an irreducible constituent

when
$i \neq 0$
, and

when
$i=0$
. As
$\gamma |_M$
is irreducible, we always have

As
$n \geq 5$
,
$p^{2nf}-1$
admits a large primitive prime divisor
$\ell $
in the sense of [Reference FeitF]. Note that
$Q:=(q^{2n}-1)_\ell $
divides
$\gamma (1)$
, and so Q divides
$|M|$
by (8.4.1). Now we can apply [Reference Katz and TiepKT1, Theorem 4.6] to the subgroup
$M < \mathrm {Sp}_{2n}(q)$
. If, in addition,

is abelian, then again by Ito’s theorem, the irreducibility of
$\gamma |_M$
implies that
$\gamma (1)$
divides
$|M/L$
, and hence,
$\ell \nmid \gamma (1)$
, a contradiction. Hence, by [Reference Katz and TiepKT1, Theorem 4.6], there is a divisor
$j<2n$
of
$2n$
such that we are in one of the following cases for L.
(b1)
$j \leq 2n/3$
and
$L \cong \mathrm {SL}_{2n/j}(q^j)$
. Note that if
$q^j=3$
, then
$2n/j =2n \geq 10$
, and so
$L \not \cong \mathrm {PSL}_4(3)$
. Hence, as in the proof of Theorem 8.2, we have

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1]. This forces the quasisimple subgroup L of G to be in the kernel of the Weil character
$\theta $
, which is absurd since
$\mathrm {Ker}(\theta ) \leq {\mathbf Z}(G)$
.
(b2)
$j|n$
and
$L \cong \mathrm {Sp}_{2n/j}(q^j)$
. Here,
$j \neq n/2$
as
$2 \nmid n$
; furthermore,
$q^{j} \geq 2^5$
if
$j=n$
, and
$q^{j} \geq 2^3$
if
$j = n/3$
(as
$2 \nmid n \geq 5$
). Hence,

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1], which leads to the contradiction
$L \leq \mathrm {Ker}(\theta ) \leq {\mathbf Z}(G)$
as in (b1).
(b3)
$j|n$
,
$j < n/2$
(recall
$2 \nmid n$
), and
$L \cong \Omega ^-_{2n/j}(q^j)$
. Now if
$j \leq n/4$
, then

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1], which again leads to the contradiction that

Suppose
$j = n/3$
. Then
$q^j \geq 2^3$
as
$2 \nmid n \geq 5$
. Hence,
$L = \Omega ^-_6(q^j)$
is a cover of
$\mathrm {PSU}_4(q^j) \not \cong \mathrm {PSU}_4(3)$
, and so

by [Reference Tiep and ZalesskiiTZ1, Theorem 1.1]. This again yields the contradiction
$L \leq \mathrm {Ker}(\theta ) \leq {\mathbf Z}(G)$
.
(b4)
$j=2j_0 \in 2{\mathbb Z}$
,
$n/j_0 \geq 3$
is odd,
$L \cong \mathrm {SU}_{n/j_0}(q^{j_0})$
, and

As
$M < G = \mathrm {SU}_n(q)$
, we have
$j_0> 1$
. In particular, n is not prime, and so we may assume
$n \geq 9$
. It follows that
$p^{(n-1)f}-1$
has a primitive prime divisor
$\ell _1$
[Reference ZsigmondyZs], which then divides
$|M|$
by (8.4.1). As
$\ell _1 \geq n> j$
,
$\ell _1$
divides
$|\mathrm {GU}_{n/j_0}(q^{j_0})|$
. Hence, we can find some
$1 \leq i \leq n/j_0$
such that
$\ell _1|(q^{ij_0}-(-1)^i)$
. The primitivity of
$\ell _1$
implies that
$(n-1)|2ij_0$
. But
$2ij_0 \leq 2n < 3(n-1)$
, so
$ij_0=n-1$
or
$(n-1)/2$
, and thus,
$j_0|\gcd (n,n-1)=1$
, a contradiction.
(b5)
$(p,nf) = (3,9)$
and
$L/{\mathbf Z}(L) = \mathrm {PSL}_2(37)$
. This case cannot, however, occur since the smallest degree of nontrivial irreducible representations of L over
$\overline {{\mathbb F}_3}$
is
$\geq 18$
, and hence, L cannot embed in
$G = \mathrm {SU}_9(3)$
.
9
$M_{2,2}$
and intersections of Fermat hypersurfaces
In this section, we fix a set
$S = \{B_0, B_1, \ldots ,B_r\}$
of integers

We will sometimes write

when we wish to emphasize the largest
$B_i$
. We work in characteristic
$p\nmid \prod _iB_i$
and choose a prime
$\ell \neq p$
so that we can speak of
$\ell $
-adic local systems. [For example, one might take for
$\ell $
a prime which divides
$\prod _iB_i$
.]
In [Reference Katz and TiepKT6, 11.2.6], given a multiplicative character
$\chi $
of
$k^\times $
for
$E/{\mathbb F}_p$
a finite extension, we introduced the local system

on
$({\mathbb G}_m \times {\mathbb A}^r)/E$
whose trace function is

We will denote this

The pullback of
${\mathcal F}^\sharp $
to
$s=1$
is the local system
${\mathcal F}(A,B_1,\ldots ,B_r,\chi )$
on
${\mathbb A}^r/E$
whose trace function is

We will denote this

As shown in §2, there is an intimate relationship between the
$M_{2,2}$
of
${\mathcal F}(S,\chi )$
and the number
$N(S,p)$
of geometrically irreducible components Z of dimension
$2$
of the
$\overline {{\mathbb F}_p}$
-locus

where
$\Sigma _{B_i}$
is the Fermat hypersurface
$x^{B_i} +y^{B_i} =z^{B_i} +w^{B_i}$
in
${\mathbb A}^4(x,y,z,w)$
. As an application of the results of the preceding sections, we will be able to completely determine this invariant
$N(S,p)$
.
In fact,
$N(S,p)$
is related to
$M_{2,2}$
of a more general kind of multi-parameter local system. Consider a partition of S as

and a polynomial
$f(x)=\sum _ic_ix^i \in E[x]$
for which

In a more cumbersome expression, we assume that

We now consider the two-parameter family

on
${\mathbb A}^2/E$
if
$A:=B_0\in S_0$
, respectively on
$({\mathbb G}_m\times {\mathbb A}^1)/E$
if
$a<b =A:=B_0$
, whose trace function at L-valued points is

The following theorem is a recapitulation of Theorems 2.3 and 2.6; see also Corollary 2.5. Remember that
$\#S \ge 3$
in this section.
Theorem 9.1. For any
$\chi $
, any partition
$S =S_0 \sqcup T$
as above and any f whose set of exponents is
$S_0$
, the following three local systems

have the same geometric
$M_{2,2}$
as each other. This common
$M_{2,2}$
is the number
$N(S,\chi )$
of geometrically irreducible components Z of dimension
$2$
of the
$\overline {{\mathbb F}_p}$
-locus
$\Sigma (S)$
with the property that on the dense open set
$xyzw \neq 0$
of Z, the rank one local system
${\mathcal L}_{\chi (xy)\overline {\chi }(zw)}$
is geometrically trivial. In particular, when
, the common
$M_{2,2}$
is
of geometrically irreducible components Z of dimension
$2$
of the
$\overline {{\mathbb F}_p}$
-locus
$\Sigma (S)$
.
Recall the definitions 4.1 and 4.2 of a data
$S = \{B_0, \ldots ,B_r\}$
to be p-finite, respectively strongly p-finite.
Theorem 9.2. Given a set
$S = \{B_0,B_1, \ldots ,B_r\}$
subject to (9.0.1) and a prime
$p \nmid \prod ^k_{i=0}B_i$
. The following statements holds for the number
$N(S,p)$
of geometrically irreducible components Z of dimension
$2$
of the
$\overline {{\mathbb F}_p}$
-locus
$\Sigma (S)$
.
-
(i) Suppose that S is either strongly p-finite or not p-finite. Then
$N(S,p)$ is
$2$ if
$2|\prod ^k_{i=0}B_i$ , and
$3$ otherwise.
-
(ii) Suppose that S is p-finite, but not strongly p-finite, that is, we are in 4.1(i) with
$q \geq 7$ , 4.1(ii) with
$q>2$ and furthermore
$2 \nmid nm_1 \ldots m_{r-1}$ if
$p> 2$ , 4.1(iii) with
$q>2$ , or 4.1(iv) with
$q> 2$ . In the case of 4.1(i),
$N(S,p)$ equals
$(q+7)/4$ if
$q \equiv 1 \ \ (\mathrm {mod} \ {4})$ and
$(q+5)/4$ if
$q \equiv 3 \ \ (\mathrm {mod} \ {4})$ . In the cases of 4.1(ii)–4.1(iv),
$N(S,p)=q+1$ .
Proof. By Theorem 9.1,
$N(S,p)$
is just the
$M_{2,2}$
of the local system
of rank
$D=B_0-1$
. Now the statements follow from Theorem 4.3 if S is strongly p-finite.
Suppose next that S is not p-finite. By Theorem 4.3, has infinite
$G_{\mathrm {geom}}$
, whence
$G_{\mathrm {geom}}^\circ =\mathrm {SL}_D$
if
$2|\prod ^r_{i=0}B_i$
and
$G_{\mathrm {geom}}=\mathrm {Sp}_D$
otherwise (note that in the latter case,
$B_0 \geq 5$
as
$k \geq 2$
, and hence,
$D \geq 4$
). It follows that the conclusion of (i) holds.
Finally, we consider the case where S is p-finite, but not strongly p-finite. By Theorem 4.3, has finite
$G_{\mathrm {geom}}$
, which is determined in [Reference Katz and TiepKT6, Theorem 11.2.3]. In the case of 4.1(i), we have
$B_0 = (q^n+1)/2$
with
$n \geq 2$
, and
$G_{\mathrm {geom}}$
is the image of
$\mathrm {Sp}_{2n}(q)$
in a Weil representation of degree
$D=(q^n-1)/2$
by [Reference Katz and TiepKT6, Theorem 11.2.3(i)], so the conclusion of (ii) follows from Theorem 8.1. In the case of 4.1(iv), we have
$B_0 = (q^n+1)/(q+1)$
with
$2 \nmid n \geq 5$
, and
$G_{\mathrm {geom}}$
is the image of
$\mathrm {SU}_{n}(q)$
in a Weil representation of degree
$D=(q^n-q)/(q+1)$
by [Reference Katz and TiepKT6, Theorem 11.2.3(iii)], whence
$N(S,p)=M_{2,2}=q+1$
by Theorem 8.3. In the case of 4.1(ii), we have
$B_0= q^n+1$
,
$B_i = q^{m_i}+1$
for
$1 \leq i \leq r-1$
,
$B_r=1$
, and furthermore,
$2 \nmid nm_1 \ldots m_{r-1}$
if
$p> 2$
. In this case,
$N(S,p) = q+1$
by Corollary 3.6. In the case of 4.1(iii), we have
$B_0= q^n+1$
and
$B_i = q^{m_i}+1$
for
$1 \leq i \leq k$
with
$q=2^f>2$
and
$n> 2$
. In this case,
$G_{\mathrm {geom}} = 2^{1+2nf}_- \cdot \Omega ^-_{2n}(q)$
by [Reference Katz and TiepKT6, Theorem 11.2.3(ii)], and the proof of [Reference Guralnick and TiepGT2, Lemma 5.1] shows that
$N(S,p)=M_{2,2}=q+1$
, the number of
$\Omega ^-_{2n}(q)$
orbits on the vectors of its natural module
${\mathbb F}_q^{2n}$
.
10 Two-parameter specializations of multi-parameter local systems
In this and the next sections, we will use our results on
$M_{2,2}$
to determine the geometric monodromy groups of the two-parameter families
${\mathcal F}(f,a,b)$
,
$1 \le a <b < \deg (f)$
, with f monic and Artin-Schreier reduced, obtained as the specializations of the multi-parameter local systems
${\mathcal F}(A,B_1, \ldots ,B_r)$
, as defined in (1.0.4) given the data (1.0.2).
Theorem 10.1. Let
$p=2$
. Consider the data (1.0.2) with
$r \geq 3$
,
$\gcd (n,m_1, \ldots ,m_r)=1$
, and

Then the following statements hold for the geometric monodromy group
$G=G_{\mathrm {geom}}$
of the local system
${\mathcal F}={\mathcal F}(f,a,b)$
defined in (1.0.4), with
$a=B_i < b=B_j$
.
-
(i) Either
$G= 2^{1+2n}_- \cdot \Omega ^-_{2n}(2)$ or
$G=2^{1+2n}_- \cdot \mathrm {SU}_n(2)$ .
-
(ii) If
$B_r=1$ and
$2 \nmid nm_1 \ldots m_{r-1}$ , then
$G=2^{1+2n}_- \cdot \mathrm {SU}_n(2)$ .
-
(iii) If
$2|n$ , then
$G=2^{1+2n}_- \cdot \Omega ^-_{2n}(2)$ .
Proof. Note that
${\mathcal F}$
is a pullback of the local system
$\tilde {\mathcal F}:={\mathcal F}(A,B_1, \ldots ,B_r)$
; furthermore,

By Theorems 9.1 and 9.2, both
${\mathcal F}$
and
$\tilde {\mathcal F}$
have
$M_{2,2}=3$
; moreover, G embeds in the (finite) geometric monodromy group
$\tilde {G} < \mathrm {Sp}_{2^n}({\mathbb C})$
of
$\tilde {\mathcal F}$
. Now we can apply [Reference Guralnick and TiepGT2, Theorem 1.5] and use the assumption
$n \geq 3$
to conclude that

furthermore,
$G/E \leq \mathrm {O}(V)$
is transitive on the set of
$2^{n-1}(2^n+1)$
(nonzero) isotropic vectors and the set of
$(2^n+1)(2^{n-1}-1)$
anisotropic vectors of the natural module
$V={\mathbb F}_2^{2n}$
of
$\mathrm {O}^-_{2n}(2)$
. In particular,

Moreover, the semidirect product
$V \rtimes (G/E)$
acts on the point set of V as a rank
$3$
affine permutation group with point stabilizer
$G/E$
. By [Reference LiebeckLi, Theorem], we arrive at one of the following possibilities for
$G/E$
.
(a)
$G/E$
is in one of the ‘exceptional’ cases listed in [Reference LiebeckLi, Theorem, part (C)]. Here,
$n= 3$
,
$4$
or
$6$
, so the lengths of the orbits of
$G/E$
on
$V \smallsetminus \{0\}$
, which are the so-called subdegrees for
$V \rtimes (G/E)$
, must be
$27,36$
, or
$119,136$
, or
$2015,2080$
, respectively. But those subdegrees do not match the subdegrees listed in [Reference LiebeckLi, Table 14].
(b)
$G/E$
is in one of the ‘extraspecial’ cases listed in [Reference LiebeckLi, Theorem, part (B)]. Here, we have
$n= 3$
, so
$(A,B_1, \ldots ,B_r) = (9,5,3,1)$
by (10.1.1). Furthermore,
$G/E$
is a subgroup of
$\mathrm {O}^-_6(2)$
that normalizes an extraspecial
$3$
-group
$3^{1+2}_{\pm }$
; in particular,
$G/E$
cannot contain
$\Omega ^-_6(2)$
. Using the list of maximal subgroups of
$\mathrm {O}^-_6(2)$
[Reference Conway, Curtis, Norton, Parker and WilsonAtlas] and the fact that
$|G/E|$
is divisible by
$27$
, we now see that
$G/E$
is solvable, and hence, G is solvable. Next, since
${\mathcal F}$
is a pullback of the local system
${\mathcal F}_{9531}$
, by Theorem 4.4(ii), we have

where
$E_3 \cong E = 2^{1+6}_-$
, and
${\mathbf Z}(E_3) = {\mathbf Z}(E)$
acts via
$\pm 1$
in the underlying representation. Since G is solvable,
$E_3G$
is a solvable subgroup of
$\mathrm {Sp}_{2^n}({\mathbb C})$
, for which we have

and hence,
$M_{2,2}(E_3G)=3$
. Now the arguments in part (d) of the proof of Theorem 4.4, with
$G_1$
replaced by G and E replaced by
$E_3$
, show that, first,
$E_3G = E_3 \cdot \mathrm {SU}_3(2)$
, and, secondly, either
$E_3 \cap G = {\mathbf Z}(E_3) = {\mathbf Z}(E)$
or
$G \geq E_3$
. In the former case,

which is a contradiction since G contains E of order
$2^7$
. So
$G \geq E_3$
, and hence,
$G=E_3 \cdot \mathrm {SU}_3(2)$
.
(c)
$G/E$
is in one of the infinite families listed in [Reference LiebeckLi, Theorem, part (A)]. First, we may have that

in particular,
$4 \nmid |G/E|$
if
$2 \nmid n$
. This rules out the case
$2 \nmid n \geq 3$
since
$2^{n-1}$
divides
$|G/E|$
by (10.1.3). Assume now that
$2|n \geq 4$
. By [Reference ZsigmondyZs],
$2^{n-1}-1$
admits a primitive prime divisor
$\ell $
, for which we have
$\ell> n$
,
$\ell $
divides
$|G/E|$
by (10.1.3), but not
$2n(2^{2n}-1)$
, a contradiction.
In the imprimitive case, by [Reference LiebeckLi, Table 12], the subdegrees are
$(2^n-1)^2$
and
$2(2^n-1)$
, none of which is divisible by
$4$
, whereas one of the subdegrees of
$G/E$
is divisible by
$2^{n-1}$
.
In the tensor product case, according to [Reference LiebeckLi, Table 12], the subdegrees are
$(q+1)(q^m-1)$
and
$q(q^{m-1}-1)(q^m-1)$
with
$q^m=2^n$
. Since the even subdegree of
$G/E$
has
$2$
-part equal to
$2^{n-1}$
, we get
$2^{n-1}=q$
. As
$n \geq 3$
, we have
$q^{2m}=2^{2n} \leq 2^{3(n-1)}=q^3$
, whence
$m=1=q$
, a contradiction.
In all the remaining cases, we again match up the subdegrees listed in [Reference LiebeckLi, Table 12] to the ones of
$G/E$
and compare the
$2$
-part of the even subdegree. First, in the case
$G/E \rhd \mathrm {SL}_a(q)$
, we either have
$q^{2a}=2^{2n}$
and
$q=2^{n-1}$
, which is impossible as shown in the preceding case, or
$a=2$
,
$q^6=2^{2n}$
, and
$q=2^{n-1}$
, which is also impossible, or
$a=5$
,
$q^{10}=2^{2n}$
, and
$q=2^{n-1}$
, which is absurd.
In the case
$G/E \rhd {}^2\! B_2(q)$
, we have
$q^4=2^{2n}$
and
$q=2^{n-1}$
, which is impossible since
$n \geq 3$
.
In the case
$G/E \rhd \Omega ^+_{10}(q)$
, we have
$q^{16}=2^{2n}$
and
$q=2^{n-1}$
, which is impossible.
Suppose
$G/E \rhd \mathrm {Sp}_6(q)$
. Then
$q^8=2^{2n}$
and
$q=2^{n-1}$
, whence
$(n,q) = (4,2)$
. But then the subdegrees are
$135,120$
but not
$136,119$
.
Suppose
$G/E \rhd \Omega ^\epsilon _{2a}(q)$
. Then
$q^{2a}=2^{2n}$
and
$q^{a-1}=2^{n-1}$
, in which case
$(a,q) = (n,2)$
. Now the even subdegree is
$2^{n-1}(2^n-\epsilon )$
, so
$\epsilon =-$
.
Suppose
$G/E \rhd \mathrm {SU}_{a}(q)$
. Then
$q^{2a}=2^{2n}$
and
$q^{a-1}=2^{n-1}$
, in which case
$(a,q) = (n,2)$
. Now the even subdegree is
$2^{n-1}(2^n-(-1)^n)$
, so
$2 \nmid n$
.
To summarize, with replacing E by
$E_3$
in the case
$(A,B_1, \ldots ,B_r) = (9,5,3,1)$
if necessary, we have shown that

Now, suppose that we have the first possibility in (10.1.5). Then

by (10.1.2). However, G injects in the geometric monodromy group
$\tilde {G}$
of
$\tilde {\mathcal F}$
, which is isomorphic to a subgroup of
$2^{1+2n}_- \cdot \Omega ^-_{2n}(2)$
by [Reference Katz and TiepKT6, Theorem 11.2.3(ii)] when
$n \geq 4$
and Theorem 4.4(ii) when
$n=3$
. Comparing the orders of G and
$\tilde {G}$
, we conclude that
$G/E = \Omega ^-_{2n}(2)$
and that
$\tilde {G} \cong 2^{1+2n}_- \cdot \Omega ^-_{2n}(2)$
. The latter conclusion implies by by [Reference Katz and TiepKT6, Theorem 11.2.3(ii)] when
$n \geq 4$
and Theorem 4.4(ii) when
$n=3$
that either
$B_r> 1$
, or
$B_r=1$
but
$2|nm_1 \ldots m_r$
.
Next, suppose that
$2 \nmid nm_1 \ldots m_{r-1}$
and
$B_r=1$
; in particular,
$n \geq 4$
since
$n,r \geq 3$
. Then G injects in the geometric monodromy group
$\tilde {G}$
of
$\tilde {\mathcal F}$
, which is isomorphic to
$2^{1+2n}_- \cdot \mathrm {SU}_{n}(2)$
, by [Reference Katz and TiepKT6, Theorem 11.2.3(ii)]. Again comparing the orders of G and
$\tilde {G}$
, we see that
$G/E = \mathrm {SU}_n(2)$
in (10.1.5), and hence, (ii) follows.
Finally, assume that we have the second possibility in (10.1.5), so
$2 \nmid n$
, and, in addition, either
$B_r> 1$
, or
$B_r=1$
but
$2|m_1 \ldots m_{r-1}$
. Then G injects in the geometric monodromy group
$\tilde {G}$
of
$\tilde {\mathcal F}$
, which is
$E_3 \cdot S$
by [Reference Katz and TiepKT6, Theorem 11.2.3(ii)] when
$n \geq 4$
and Theorem 4.4(ii) when
$n=3$
, where
$E_3 \cong E = 2^{1+2n}_-$
and
$S \cong \Omega ^-_{2n}(2)$
. Certainly,
$E_3G \leq \tilde {G} < \mathrm {Sp}_{2^n}({\mathbb C})$
still has
$M_{2,2}=3$
; see (10.1.4). So the preceding arguments but applied to
$E_3G$
show that (10.1.5) also holds for
$E_3G$
:

In the former case, we have
$E_3G=\tilde {G}$
, and so the composition factors of G are
$\Omega ^-_{2n}(2)$
and
$C_2$
, all present. But this contradicts the fact that
$G/E \rhd \mathrm {SU}_n(2)$
(which yields a composition factor
$\mathrm {PSU}_n(2)$
when
$n \geq 4$
and
$C_3$
when
$n=3$
). So we must have that

Recall that
$E_3G/E_3$
is a subgroup of
$S = \Omega (W)$
, where
$W:=E_3/{\mathbf Z}(E_3) = {\mathbb F}_2^{2n}$
carries the quadratic form

and symplectic form

both invariant under the normal subgroup
$G_1:=\mathrm {SU}_n(2)$
of
$E_3G/E_3$
. Assuming
$n> 3$
and applying [Reference Katz and TiepKT6, Proposition 8.4.1], we obtain that

where
$W_1 := {\mathbb F}_4^n$
. Working from a standard basis for the Hermitian form on
$W_1$
(over
${\mathbb F}_4$
) back to a Witt basis of W (over
${\mathbb F}_2$
), one can readily check that the Galois automorphism
$\alpha \mapsto \alpha ^2$
of
${\mathbb F}_4$
induces (in that standard basis) an element of
$\mathrm {O}(W)$
which is a product of n reflections. Since
$2 \nmid n$
, this element is not in
$S=\Omega (W)$
. However, since
$\mathrm {O}(W)$
has index
$2$
over S and
$2 \nmid n \geq 5$
,
$\mathrm {GU}(W_1) \leq S$
. It follows that
$E_3G/E_3 \leq {\mathbf N}_S(G_1) \cong \mathrm {GU}_n(2)$
. Now we can use the fact that
$G={\mathbf O}^{2'}(G)$
to conclude that

Suppose now that
$n=3$
. Then, using (10.1.6) and the fact that
$E_3G/E_3$
is transitive on both the nonzero singular vectors and the non-singular vectors of the quadratic space
${\mathbb F}_2^6$
, and arguing as in part (b) (recalling that
${\mathbf O}_3(\mathrm {SU}_3(2)) = 3^{1+2}_+$
), we see that (10.1.7) holds in this case as well.
We have therefore shown that
$|G| \leq |E_3G| = |E_3|\cdot |\mathrm {SU}_n(2)| = |E| \cdot |\mathrm {SU}_n(2)|$
. But
$G/E \rhd \mathrm {SU}_n(2)$
by (10.1.5), so in fact,
$G/E=\mathrm {SU}_n(2)$
.
In some special instances of the case where
$2|m_1 \ldots m_r$
but
$2 \nmid n$
of Theorem 10.1, we can also prove that
${\mathcal F}(f,a,b)$
has
$G_{\mathrm {geom}} = E \cdot \Omega ^-_{2n}(2)$
. [Also see Theorem 11.7 about the ‘generic’ situation.] To do this, we first prove a general statement.
Proposition 10.2. Let
$k/{\mathbb F}_p$
be an finite extension,
$f(x)\in k[x]$
a polynomial of degree A with
$p \nmid A$
, and a an integer

Denote by
${\mathcal F}_{f,a}$
the lisse sheaf on
${\mathbb A}^1$
whose trace function at a point
$t \in L$
, for
$L/k$
a finite extension, is

Then the following statements hold for its
$G_{\mathrm {geom}}$
.
-
(i)
$G_{\mathrm {geom}}$ , indeed its
$I(\infty )$ , contains elements of order a. In particular,
$G_{\mathrm {geom}}$ is not a finite p-group.
-
(ii) Assume in addition that
$\gcd (A,a)=1$ . Then
$G_{\mathrm {geom}}$ contains a subquotient of order
$(A-a)_{p'}$ .
Proof. (i) Up to a Tate twist
$(1/2)$
,
${\mathcal F}_{f,a}$
is the Fourier transform of the Kummer direct image
$[a]_\star ({\mathcal L}_{\psi (f)})$
:

The sheaf
${\mathcal G}$
is lisse of rank a on
${\mathbb G}_m$
, its
$I(0)$
-representation is
, and its
$I(\infty )$
-representation has all slopes
$A/a> 1$
. By Laumon’s theory of local Fourier Transform (cf. [Reference KatzKa2, 7.4.2, 7.4.4(2)]), the
$I(\infty )$
-representation of
${\mathcal F}_{f,a}$
is the direct sum

The first factor is . Thus, the subgroup
$I(\infty ) \le G_{\mathrm {geom}}$
contains elements of order a.
(ii) The
$I(\infty )$
-representation of
${\mathcal G}$
has rank a, and all slopes
$A/a$
. By Laumon’s result [Reference KatzKa2, 7.4.1(1)],the second factor
$\mathrm {FT}_\psi \mathrm { loc}(\infty ,\infty )({\mathcal G}_{|I(\infty )})$
has rank
$A-a$
and all slopes
$A/(A-a)$
. If
$\gcd (A,a)=1$
, one knows [Reference KatzKa1, 1.1.4] that the second factor is
$I(\infty )$
-irreducible, and one knows further that denoting by
$(A-a)_{p'}$
the
$p'$
part of
$A-a$
, the second factor is the Kummer induction
$[(A-a)_{p'}]_\star W$
of an irreducible
$I(\infty )$
-representation of dimension the p part of
$A-a$
. This description of the second factor makes visible the group
$\mu _{(A-a)_{p'}}$
as a quotient of the wild part of the
$I(\infty )$
-representation of
${\mathcal F}_{f,a}$
.
Corollary 10.3. Suppose that we are in the case
$2 \nmid n$
of Theorem 10.1 and some
$m \in \{m_i,m_j\}$
is even and strictly positive. Then the local system
${\mathcal F}(f,a,b)$
has
$G_{\mathrm {geom}} = 2^{1+2n}_- \cdot \Omega ^-_{2n}(2)$
.
Proof. For definiteness, we will assume
$m=m_i$
, so that
$a=2^m+1$
. By Theorem 10.1, it suffices to prove that
$|G_{\mathrm {geom}}|$
is divisible by some odd prime which does not divide
$|\mathrm {SU}_n(2)|$
.
First, consider the case
$m> n/2$
. Applying Proposition 10.2(i) to the pullback
$t=1$
of
${\mathcal F}(f,a,b)$
, we see that
$|G_{\mathrm {geom}}|$
is divisible by
$2^{m}+1$
. Since
$2m_i \neq 6$
,
$2^{2m}-1$
has a primitive prime divisor
$\ell $
by [Reference ZsigmondyZs]. Then
$\ell $
certainly divides both
$2^m+1$
and
$|G_{\mathrm {geom}}|$
. Suppose
$\ell $
divides
$|\mathrm {SU}_n(2)|$
. Then there is some
$1 \leq k \leq n$
such that
$\ell $
divides
$2^k-(-1)^k$
. In particular,
$\ell |(2^{2k}-1)$
. The primitivity of
$\ell $
implies that
$2m$
divides
$2k$
. But
$2m> n$
and
$2k \leq 2n$
, so
$k=m$
. It follows that
$\ell $
divides
$2^k-(-1)^k=2^{m}-1$
, contradicting the choice of
$\ell $
. Thus
$\ell $
does not divide
$|\mathrm {SU}_n(2)|$
, as desired.
Assume now that
$2 \leq m < n/2$
. Suppose that some prime r divides both
$2^n+1$
and
$2^m+1$
. Then r divides
$\gcd (2^{2n}-1,2^{2m}-1) = 2^{2e}-1$
for
$e:=\gcd (n,m)$
. As
$2 \nmid n$
, e is odd, and so
$2e$
divides m. But in this case, r divides
$2^m-1$
and so cannot divide
$2^m+1$
, a contradiction. Thus,
$2^n+1$
and
$2^m+1$
are coprime. Hence, by Proposition 10.2(ii) applied to the pullback
$t=1$
of
${\mathcal F}(f,a,b)$
,
$|G_{\mathrm {geom}}|$
is divisible by
$2^{n-m}-1$
. Note that
$n \geq 3$
and
$n-m>n/2$
is odd, so
$n-m \geq 3$
. By [Reference ZsigmondyZs],
$2^{n-m}-1$
admits a primitive prime divisor
$\ell _1$
. Suppose
$\ell _1$
divides
$|\mathrm {SU}_n(2)|$
. Then there is some
$1 \leq k \leq n$
such that
$\ell _1$
divides
$2^k-(-1)^k$
. In particular,
$\ell _1|(2^{2k}-1)$
. The primitivity of
$\ell _1$
implies that
$n-m$
divides
$2k$
, and hence,
$n-m$
divides k since
$n-m$
is odd. But
$2(n-m)> n \geq k$
, so
$k=n-m$
. It follows that
$\ell _1$
divides
$2^k-(-1)^k=2^{n-m}+1$
, contradicting the choice of
$\ell _1$
. Thus,
$\ell _1$
does not divide
$|\mathrm {SU}_n(2)|$
, and we are done in this case as well.
11 Semicontinuity
First, we recall some results from [Reference KatzKa2, 8.17, 8.18].
The situation we consider is the following. We are given a normal connected affine noetherian scheme
$S=\mathrm {Spec \,}(A)$
with A a noetherian normal integral domain with fraction field K, and a chosen algebraic closure
$\overline {K}$
of K. Thus,
$\mathrm {Spec \,}(K)$
is a generic point
$\eta $
of S, and
$\mathrm {Spec \,}(\overline {K})$
is a geometric point
$\overline {\eta }$
of S. We are given
$X/S$
a smooth S-scheme of relative dimension D, with geometrically connected fibres, and
$\phi \in X(S)$
a section of
$X/S$
. Then
$\phi (\overline {\eta })$
is a geometric point of X. We are given a finite group G and a surjective homomorphism

For each geometric point s of S,
$\phi (s)$
is a geometric point of
$X_s$
(and also of X). We have a continuous group homomorphism

This last isomorphism is only canonical up to inner automorphism of the target group
$ \pi _1(X,\phi (\overline {\eta }))$
. By composition, we get a group homomorphism

which is well defined up to inner automorphism of G. This applies in particular with s taken to be
$\overline {\eta }$
. We are interested in how the image of
$\pi _1(X_s,\phi (s)) $
in G compares with the image of
$\pi _1(X_{\overline {\eta }},\phi (\overline {\eta }))$
in G: when are these two subgroups of G conjugate in G? Let us denote these image groups
$G_{s}$
and
$G_{\overline {\eta }}$
.
Theorem 11.1. There exists a dense open set
$U \subset S$
such that for any geometric point
$s \in U$
,
$G_s$
and
$G_{\overline {\eta }}$
are conjugate subgroups of G. Moreover, for any geometric point
$s \in S$
,
$G_{s}$
is conjugate to a subgroup of
$G_{\overline {\eta }}$
.
Proof. We first reduce to the case when
$G_{\overline {\eta }}=G$
.
Consider the scheme
$X_{\eta }$
, a smooth K-scheme, and compare it to the smooth
$\overline {K}$
-scheme
$X_{\overline {\eta }}$
. We have the
$\pi _1$
short exact sequence

The scheme
$X_{\eta }$
has the same function field as X, so the canonical map is surjective:

Thus, the image of
$\pi _1(X_{\eta },\phi (\overline {\eta }))$
is G, while the image of
$\pi _1(X_{\overline {\eta }},\phi (\overline {\eta }))$
is a normal subgroup H of G, with
$G/H$
the Galois group of some finite Galois extension
$L/K$
. View
$X_{\overline {\eta }}$
as
$X\otimes _A \overline {K}$
. Then for the finite Galois extension
$L/K$
,
$ \pi _1(X_{\overline {\eta }},\phi (\overline {\eta }))$
and
$ \pi _1(X\otimes _A L,\phi (\overline {\eta }))$
have the same image H.
Now replace S by T:=the normalization of S in L (i.e., the
$\mathrm {Spec \,}$
of the integral closure of A in L), replace X by
$X_T:=X\times _ST$
, and replace
$\phi $
by the section
$\phi _T$
(in terms of the finite map
$f:T \rightarrow S$
,
$\phi _T$
is
$(\phi \circ f)\times id_T$
as map to
$X\times _ST$
). In this new situation, the image H of
$\pi _1((X_T)_{\phi _T(\overline {\eta })},\phi _T(\overline {\eta }))$
is equal to the image of
$\pi _1(X_T,\phi _T(\overline {\eta }))$
. Because
$L/K$
is separable (being Galois), one knows that
$f:T \rightarrow S$
is both finite and surjective. Being finite, it is proper. Thus, f is closed. Hence, the image of a dense open set
$V=T \smallsetminus Z$
of T contains the dense open set
$U:=S \smallsetminus f(Z)$
of S.
Returning to the original notation
$(X,S,\phi ,G)$
, this completes reduction to the case when
$G_{\overline {\eta }}=G$
, for G the image of
$\pi _1(X,\phi (\overline {\eta }))$
. In this case, every
$G_s$
is visibly (conjugate to) a subgroup of G, by the homomorphism

Let
$E \rightarrow X$
denote the finite etale G-covering classified by the surjection

Precisely because this is a surjection, E is connected. Being finite etale over X, which is in turn smooth over the normal scheme S, we see that E is itself smooth over S, of relative dimension d. Let us denote by

the structural morphism.
Then
$E_s$
is a finite etale G-covering of
$X_s$
, but it may not be connected. One has
$G_s=G$
if and only if
$E_s$
, which is smooth over s of dimension d, is itself connected ( or equivalently geometrically irreducible, being smooth over s). [Indeed, the index of
$G_s$
in G is the number of geometrically irreducible components of
$E_s$
.]
For any prime
$\ell $
invertible on S, the
${\mathbb F}_\ell $
-rank of the stalk at s of
$R^{2d}g_!({\mathbb F}_\ell )$
is the number of geometrically irreducible components of
$E_s$
. By general constructibility theorems,
$R^{2d}g_!({\mathbb F}_\ell )$
is a constructible sheaf on S, so on a dense open set has constant rank. But at the generic point
$\overline {\eta }$
, the rank is one (precisely because
$G_{\overline {\eta }}=G$
). Therefore, the rank is one on some dense open set U. Thus, for every
$s \in U$
, we have
$G_s=G$
. [If there is no prime
$\ell $
invertible on S, pick any two distinct primes, say
$2,3$
, and work separately on
$S[1/2]$
and on
$S[1/3]$
.] For a more direct proof, see [Reference GrothendieckEGA, 9.7.8] or [StPr, Lemma 37.27.5].
Corollary 11.2. Hypotheses and notations as in Theorem 11.1, suppose that for some geometric point
${s_0} \in S$
,
$G_{s_0}=G$
. Then
$G_{\overline {\eta }}=G_{s_0} =G$
, and hence, there exists a dense open set U of S such that we have
$G_s=G$
for every geometric point
$s \in S$
.
Proof. We have the inclusion, up to conjugation,
$G_{s_0} \subset G_{\overline {\eta }}$
. We also have the inclusion
$G_{\overline {\eta }} \subset G$
, simply via the map
$\pi _1(X_{\overline {\eta }},\phi (\overline {\eta })) \rightarrow \pi _1(X,\phi (\overline {\eta }))$
. Thus,
$G=G_{s_0} \subset G_{\overline {\eta }} \subset G$
, whence
$G_{\overline {\eta }} = G$
, and we apply Theorem 11.1.
Here is a particular instance of Corollary 11.2.
Proposition 11.3. Let p be a prime,
$q=p^f$
,
$\chi $
a (possibly trivial) multiplicative character of
${\mathbb F}_q^\times $
,
$r \ge 2$
an integer, and let
$A> B_1 > \ldots > B_r \ge 1$
be integers with
$\gcd (A,B_1, \ldots ,B_r)=1$
and
$p \nmid AB_1 \ldots B_r$
. Consider the local system
${\mathcal F}(A,B_1, \ldots ,B_r,\chi )$
on
${\mathbb A}^r/{\mathbb F}_q$
with trace function for any finite extension
$L/{\mathbb F}_q$

in characteristic p, of rank
$D=A-1$
if
and
$D=A$
otherwise, with geometric monodromy group
$G=G_{\mathrm {geom}}$
. Given a choice
$i_0 \in [1,r]$
and a polynomial
$f(x) \in \overline {{\mathbb F}_p}[x]$
of the form

denote by
${\mathcal F}(A,B_{i_0},f,\chi )$
the local system on
${\mathbb A}^1/K_f$
with

whose trace function, for any finite extension
$L/K_f$
, is

and by
${\mathcal F}(A,B_{i_0},f=0,\chi )$
, the local system on
${\mathbb A}^1/{\mathbb F}_q$
whose trace function, for any finite extension
$L/{\mathbb F}_q$
, is

Suppose that
${\mathcal F}(A,B_1, \ldots ,B_r,\chi )$
has finite geometric monodromy group G and that the specialized local system
${\mathcal F}(A,B_{i_0},f=0,\chi )$
has the same geometric monodromy group G. Then in the
${\mathbb A}^{r-1}/{\mathbb F}_p$
of possible f, there is an open dense set
$U \subset {\mathbb A}^{r-1}$
such that for any
$f \in U$
, the specialized local system
${\mathcal F}(A,B_{i_0},f=0,\chi )$
has the same geometric monodromy group G.
Here are some examples. In the first two of these examples, we are given
$r+1$
integers

with
$2|nm_1 \ldots m_r$
,
$\gcd (n,m_1, \ldots ,m_r)=1$
.
-
(i)
$p=2$ ,
$q=2^f$ ,
$A=q^n+1$ ,
$r \geq 2$ ,
$B_i = q^{m_i}+1$ for
$1 \leq i < r$ , and either (
$m_r>0$ ,
$B_r = q^{m_r}+1$ ) or (
$B_r=1$ ,
$m_r=0$ , and
$2|nm_1 \ldots m_{r-1}$ ). Suppose that
$2|nm_{i_0}$ and
$\gcd (n,m_{i_0})=1$ . Then
has the same geometric monodromy group G as does
– namely, the group
$2^{1+2nf}_{-} \cdot \Omega ^-_{2n}(q)$ . Simplest example:
$i_0=1$ and
$m_1=n-1$ . The calculations of the monodromy groups are Theorem 11.2.3 (ii) and Theorem 10.3.13(iii) of [Reference Katz and TiepKT6] for
$q^n> 8$ and Theorem 4.4 for
$q^n=4,8$ .
-
(ii)
$p> 2$ ,
$q=p^f$ ,
$\chi $ is either
or the quadratic character
$\chi _2$ ,
$A = (q^n+1)/2$ ,
$B_i = (q^{m_i}+1)/2$ ,
$1 \leq i \leq k$ , where
$n> m_1 > \ldots > m_r \geq 0$ are integers with
$2|nm_1 \ldots m_r$ ,
$\gcd (n,m_1, \ldots ,m_r)=1$ , and
or
$\chi = \chi _2$ . Suppose that
$2|nm_{i_0}$ and
$\gcd (n,m_{i_0})=1$ . Then
${\mathcal F}(A,B_{i_0},f=0,\chi )$ has the same geometric monodromy group G as does
${\mathcal F}(A,B_1, \ldots ,B_r,\chi )$ , namely the image of
$\mathrm {Sp}_{2n}(q)$ in one of its irreducible Weil representations of degree D, with
$D=A-1$ for
and
$D=A$ for
$\chi =\chi _2$ . Simplest example:
$i_0=1$ and
$m_1=n-1$ . The calculations of the monodromy groups are Theorem 11.2.3 (i) and Theorem 10.3.13(i) of [Reference Katz and TiepKT6].
-
(iii) p arbitrary,
$q=p^f$ . In this third example,
$n> m_1 > \ldots > m_r \geq 1$ are all odd, and
$\gcd (n,m_1, \ldots ,m_r)=1$ ,
$\chi $ is a character of
${\mathbb F}_{q^2}^\times $ of order dividing
$q+1$ . Suppose that
$\gcd (n,m_{i_0})=1$ . Then
${\mathcal F}(A,B_{i_0},f=0,\chi )$ has the same geometric monodromy group G as does
${\mathcal F}(A,B_1, \ldots ,B_r,\chi )$ , namely the image of
$\mathrm {SU}_n(q)$ in a Weil representation of degree D, with
$D=A-1$ for
and
$D=A$ for
. Simplest example:
$i_0=1$ and
$m_1=n-2$ . The calculations of the monodromy groups are Theorem 11.2.3 (iii) and Theorem 10.3.13(ii) of [Reference Katz and TiepKT6].
Remark 11.4. In the above examples, we need the existence of an index
$i_0$
such that
$\gcd (n, m_{i_0})=1$
. So we have nothing to say about one-parameter specializations in cases such as
$(n,m_1,\ldots m_r)=(6,3,2)$
or
$(15,6,5,3)$
or
$(30,5,3,2)$
.
A second problem is that in the examples, although we know
$G_{\mathrm {geom}}$
for an open dense set U of f’s, we do not know which subgroups of
$G_{\mathrm {geom}}$
can occur for f’s not in U, nor for which f these smaller groups occur.
Next, we consider some one- and two-parameter systems in characteristic
$p=2$
. We begin with a lemma on generalized Pink–Sawin sheaves.
Lemma 11.5. Let p be a prime,
$n \ge 1$
an integer,
$k/{\mathbb F}_p$
a finite extension, and
$f(x) \in k[x]$
a polynomial of the form

Denote by
${\mathcal F}_f$
the lisse sheaf on
${\mathbb A}^1/k$
whose trace function at a point
$t \in L$
, for
$L/k$
a finite extension, is

that is,
${\mathcal F}_f$
is, up to the Tate twist
$(1/2)$
which makes it pure of weight zero, the Fourier Transform
$FT_\psi ({\mathcal L}_{\psi (f)})$
. Then there exists an explicit finite extension
$L_0/k$
such that for every finite extension
$L_1/L_0$
and every
$t \in L_1$
,
$|\mathrm {Trace}(\mathrm {Frob}_{t,L_1}|{\mathcal F}_f)|$
is either
$0$
or
$p^{n}$
.
Proof. This is an instance of the argument of [vG-vV, Section 5]. Write

for
$R(x)$
the additive polynomial
$\sum _{i=1}^n a_ix^{p^i}$
. Then

(substituting
$(x,y) \mapsto (x+y,y)$
and remembering that
$R(x+y)=R(x)+R(y)$
),


For the inner sum, the
$\mathrm {Trace}_{L_1/{\mathbb F}_p}$
of
$yR(x)+xR(y)$
is equal to the
$\mathrm {Trace}_{L_1/{\mathbb F}_p}$
of

Let us denote by

Equivalently,
$W_R(L_1)$
is the set of zeroes in
$L_1$
of the additive polynomial

The sum

which is
$1$
if
$P_R(x)=0$
, and zero otherwise.
Take for
$L_0$
a field containing
${\mathbb F}_{p^2}$
and each of the
$p^{2n}$
zeroes of
$P_R(x)$
. [Notice that the highest-degree term of
$P_R(x)$
is
$a_n^{p^n}x^{p^{2n}}$
and its lowest-degree term is
$a_nx$
, so its derivative is the nonzero constant
$a_n$
, and hence,
$P_R(x)$
has
$p^{2n}$
distinct zeroes over
$\overline {{\mathbb F}_p}$
]. Then

One checks that the map
$x \mapsto \psi _{L_1}(xR(x)+tx)$
is a
$\mu _p$
-valued character of the finite abelian group
$W_R(L_1)$
, so the sum
$\sum _{x\in W_R(L_1)}\psi _{L_1}(xR(x)+tx)$
is either
$0$
, if the character is nontrivial, or is
$\#W_R(L_1)$
. But over any extension
$L_1/L_0$
,
$W_R(L_1)= W_R(L_0)$
, whose cardinality is
$p^{2n}$
.
Corollary 11.6. Keep the notation and assumption of Lemma 11.5. For every finite extension
$L_1/L_0$
and every
$t \in L_1$
,
$\mathrm {Trace}(\mathrm {Frob}_{t,L_1}|{\mathcal F}_f)$
is either
$0$
or
$\pm p^n\zeta $
for some
$\zeta \in \mu _p$
.
Proof. The trace lies in
${\mathbb Z}[\zeta _p]$
and divides
$p^{2n}$
in that ring, so is a unit at all places outside p, while at the unique place over p of
${\mathbb Q}(\zeta _p)$
, it and its complex conjugate each have absolute value
$p^n$
. By the product formula, this trace, divided by
$p^n$
, is an element of
${\mathbb Z}[\zeta _p]$
all of whose absolute values (at all places) are
$1$
, and hence is a root of unity in
${\mathbb Z}[\zeta _p]$
.
Theorem 11.7. Let
$p=2$
,
$q=p^f$
,
$r \geq 2$
,
$n> m_1 > \ldots > m_r \ge 0$
,
$\gcd (n,m_1, \ldots ,m_r)=1$
,
$2|nm_1 \ldots m_r$
, and set
$A=q^n+1$
,
$B_i=q^{m_i}+1$
,
$1 \leq i \leq r-1$
, and either
$ B_r=q^{m_r}+1$
with
$m_r \geq 1$
or
$(B_r,m_r) = (1,0)$
. Recall (see [Reference Katz and TiepKT6, Theorem 11.2.3(ii)] and Theorem 4.4) that the local system

has
$G_{\mathrm {geom},{\mathcal F}_{\mathrm {up}}}=:G_{\mathrm {up}}$
equal to
$2^{1+2nf}_- \rtimes \mathrm {SU}_n(q)$
if
$B_r=1$
and
$2 \nmid nm_1 \ldots m_{r-1}$
, and
$2^{1+2nf}_{-} \cdot \Omega ^-_{2n}(q)$
otherwise. Assume in addition that
$(q,r,n,m_1,m_2) \neq (2,2,3,1,0)$
. Fix a choice of
$1 \le i \leq j \leq r$
. If
$i=j$
, set
$d:=1$
. If
$i <j$
, set
$d:=2$
and assume
$r \ge 3$
. For f in the space
${\mathbb A}^{r-d}$
of all polynomials

denote by
${\mathcal F}(A,B_i,B_j,f)$
the local system on
${\mathbb A}^d$
whose trace function is

when
$i=j$
and

when
$i<j$
. Then one of the following statements holds.
-
(i) There is an open dense set
$U \subset {\mathbb A}^{r-d}$ such that for any
$f \in U$ ,
${\mathcal F}(A,B_i,B_j,f)$ has
$G_{\mathrm {geom}}$ the group
$G_{\mathrm {up}}$ .
-
(ii)
$i=j$ , and for all
$f \in {\mathbb G}_m^{r-1}$ ,
${\mathcal F}(A,B_i,B_i,f)$ has
$G_{\mathrm {geom}}$ the extraspecial
$2$ -group
$2^{1+2nf}_-$ .
In particular, conclusion (i) holds if
$i<j$
. Moreover, conclusion (ii) holds if and only if
$i=j$
and
$B_i=1$
.
Proof. We first note that each
${\mathcal F}(A,B_i,B_j,f)$
is a pullback of
${\mathcal F}_{\mathrm {up}}$
, so its geometric monodromy group
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
is a subgroup of
$G_{\mathrm {up}}$
, well defined up to conjugacy in
$G_{\mathrm {up}} = E \cdot S$
, where
$E = 2^{1+2nf}_-$
and
$S = \mathrm {SU}_n(q)$
, respectively
$S = \Omega ^-_{2n}(q)$
. We further note that, so long as all coefficients of f are nonzero, the group
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
is an irreducible subgroup of
$\mathrm {Sp}_{q^n}({\mathbb C})$
; cf. [Reference Katz and TiepKT4, Prop. 2.4]. By the specialization Theorem 11.1, there is a subgroup
$G_0 \le G_{\mathrm {up}}$
, well defined up to conjugacy in
$G_{\mathrm {up}}$
, and a dense open set
$U \subset {\mathbb A}^{r-d}$
such that for every
$f \in U$
,
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
is conjugate to
$ G_0$
. Concretely, there is a nonzero polynomial
$P(x_k \mid 1 \leq k \leq r,~k \neq i,j)$
in
$r-d$
variables over
$\overline {{\mathbb F}_q}$
such that any
$f(x)=\sum _{1 \leq i \leq m,\,i \neq i,js}c_ix^{B_i}$
with
$P(c_k \mid 1 \leq k \leq r,~k \neq i,j) \neq 0$
lies in U. Let us denote
$U_{\mathrm {up}} \subset {\mathbb A}^{r}$
(with coordinates
$(s_1,\ldots , s_r)$
) the dense open set on which
$P(s_k \mid 1 \leq k \leq r,~k \neq i,j) \neq 0$
. Replacing P by
$P\prod _{k \ne i,j}x_k$
, we reduce to the case when every
$f \in U$
has all coefficients nonzero, and hence, for every
$f \in U$
,
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
is an irreducible subgroup of
$\mathrm {Sp}_{q^n}({\mathbb C})$
. In particular, the group
$G_0$
is an irreducible subgroup of
$\mathrm {Sp}_{q^n}({\mathbb C})$
.
Because
$U_{\mathrm {up}} \subset {\mathbb A}^r$
is a dense open set,
${\mathcal F}_{\mathrm {up}}$
on
${\mathbb A}^r$
and
${\mathcal F}_{\mathrm {up}}|_{U_{\mathrm {up}}}$
on
$U_{\mathrm {up}}$
have the same
$G_{\mathrm {geom}}$
– namely,
$G_{\mathrm {up}}$
. Both
$G_{\mathrm {up}}$
and the arithmetic group
$G_{\mathrm {arith},{\mathcal F}_{\mathrm {up}},{\mathbb F}_2}$
are finite, with
$G_{\mathrm {up}} \lhd G_{\mathrm {arith},{\mathcal F}_{\mathrm {up}}} \le \mathrm {Sp}_{q^n}({\mathbb C})$
, with the quotient
$G_{\mathrm {arith},{\mathcal F}_{\mathrm {up}},{\mathbb F}_2}/G_{\mathrm {up}}$
a finite cyclic group. In the case
$S=\Omega ^-_{2n}(q)$
, one knows that

contains
$G_{\mathrm {up}}$
with index dividing
$2f$
. In the case
$S=\mathrm {SU}_{n}(q)$
, our assumptions imply that
$(n,q) \neq (3,2)$
, whence S is simple and

contains
$G_{\mathrm {up}}$
with index dividing
$2f(n+1)$
; see [Reference Katz and TiepKT6, Proposition 8.4.1(b2)]. [For completeness, we note that when
$S=\mathrm {SU}_n(q)$
with
$(n,q)=(3,2)$
,
$G_{\mathrm {arith},{\mathcal F}_{\mathrm {up}},{\mathbb F}_2}$
has index
$2$
over
$G_{\mathrm {geom},{\mathcal F}_{\mathrm {up}}}$
by Theorem 4.4(iii).]
Thus, over any extension
$L/{\mathbb F}_{q^{2n+2}}$
,
$G_{\mathrm {arith},{\mathcal F}_{\mathrm {up}},L}=G_{\mathrm {up}}$
. By the finite group version [Reference Katz and SarnakKaS, Theorem 9.7.13] of Deligne’s equidistribution theorem, applied to
${\mathcal F}_{\mathrm {up}}|_{U_{\mathrm {up}}}$
, over any sufficiently large finite extension
$L/{\mathbb F}_{q^{2n+2}}$
, every element
$\gamma \in G_{\mathrm {up}}$
is conjugate to some Frobenius
$\mathrm {Frob}_{(s_1,\ldots s_r),L}$
with
$(s_1,\ldots ,s_r) \in U_{\mathrm {up}}(L)$
. Such a Frobenius is
$\mathrm {Frob}_{s_i,s_j,L}$
on
${\mathcal F}(A,B_i,B_j,f)$
for

Now view
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
as a subgroup of
$G_{\mathrm {up}}$
. Then
$\mathrm {Frob}_{s_i,s_j,L}$
lies in
$G_{\mathrm {arith},{\mathcal F}(A,B_i,B_j,f)}$
, so normalizes
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
. But
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
is conjugate in
$G_{\mathrm {up}}$
to
$G_0$
, and hence, every conjugacy class in
$G_{\mathrm {up}}$
contains an element that normalizes
$G_0$
. Thus, the normalizer
${\mathbf N}_{G_{\mathrm {up}}}(G_0)$
of
$G_0$
in
$G_{\mathrm {up}}$
meets every conjugacy class in
$G_{\mathrm {up}}$
. Therefore,

whence

In particular,
$EG_0/E$
is a normal subgroup of the simple group
$G_{\mathrm {up}}/E \cong S$
, whence
$EG_0=E$
or
$EG_0 =G_{\mathrm {up}}$
. Note that any proper subgroup of E has order
$\leq q^{2n}$
and so cannot be irreducible on
${\mathbb C}^{q^n}$
, and thus, the only irreducible subgroup of E is E itself. Furthermore,
$M_{2,2}(E)=q^{2n}> q+1$
, whereas
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
with
$i<j$
has the same
$M_{2,2}$
as that of
$G_{\mathrm {up}}$
, which is equal to
$q+1$
, by Theorems 9.1 and 9.2. Hence, in the former case, we must have that
$G_0 = E$
,
$i=j$
, and furthermore,
$B_i=1$
by Proposition 10.2, and thus, conclusion (ii) holds by Theorem 11.1.
In the latter case,
$(E \cap {\mathbf Z}(E)G_0)/{\mathbf Z}(E)$
is a normal subgroup of

that is contained in
$E/{\mathbf Z}(E)$
. But S acts irreducibly on
$E/{\mathbf Z}(E) = {\mathbb F}_2^{2nf}$
, so either
$E \cap {\mathbf Z}(E)G_0 = {\mathbf Z}(E)$
or
${\mathbf Z}(E)G_0 \geq E$
. However, since
$EG_0=G_{\mathrm {up}}$
and
$G_0 \lhd G_{\mathrm {up}}$
, the first possibility leads to
$G_{\mathrm {up}}/{\mathbf Z}(E) \cong E/{\mathbf Z}(E) \times S$
, which is impossible. So
${\mathbf Z}(E)G_0 \geq E$
, in which case we have

(since
${\mathbf Z}(E) = {\mathbf Z}(G_{\mathrm {up}})$
), whence
$G_0={\mathbf Z}(E)G_0=EG_0=\tilde {G}$
and (i) holds.
Assume now that
$i=j$
and
$B_i=1$
. By Corollary 11.6,
$\varphi (x) \in \{\pm q^n,0\}$
for all
$x \in G_0$
, where
$\varphi $
denotes the character of the underlying representation. It follows that
$[\varphi ,\varphi ]_{G_0} = q^{2n}|{\mathbf Z}(G_0)|/|G_0|$
. As
${\mathbf Z}(G_0)={\mathbf Z}(E) \cong C_2$
and
$\varphi \in \mathrm {Irr}(G_0)$
, we conclude that
$|G_0|=2q^{2n}=|E|$
, and hence,
$G_0=E$
.
Here is the odd-p analogue of the above result:
Theorem 11.8. Let
$p>2$
,
$q=p^f$
,
$r \geq 2$
,
$n> m_1 > \ldots > m_{r-1} > 0$
,
$\gcd (n,m_1, \ldots ,m_{r-1})=1$
, and
$A=q^n+1$
,
$B_i=q^{m_i}+1$
,
$1 \leq i \leq r-1$
, and
$B_r=1$
. Recall (see [Reference Katz and TiepKT6, Theorem 11.2.3(i-bis)]) that the local system

has
$G_{\mathrm {geom},{\mathcal F}_{\mathrm {up}}}=:G_{\mathrm {up}}$
equal to

if
$2 \nmid nm_1 \ldots m_{r-1}$
, and

otherwise. Fix a choice of
$1 \leq i \leq j \leq r$
. If
$i=j$
, set
$d:=1$
. If
$i <j$
, set
$d:=2$
and assume
$r \ge 3$
. For f in the space
${\mathbb A}^{r-d}$
of all polynomials

denote by
${\mathcal F}(A,B_i,B_j,f)$
the local system on
${\mathbb A}^d$
whose trace function is

when
$i=j$
and

when
$i<j$
. Then one of the following statements holds.
-
(i) There is an open dense set
$U \subset {\mathbb A}^{r-d}$ such that for any
$f \in U$ ,
${\mathcal F}(A,B_i,B_j,f)$ has
$G_{\mathrm {geom}}$ the group
$G_{\mathrm {up}}$ .
-
(ii)
$i=j$ , and for all
$f \in {\mathbb G}_m^{r-1}$ ,
${\mathcal F}(A,B_i,B_i,f)$ has
$G_{\mathrm {geom}}$ the extraspecial p-group
$p^{1+2nf}_+$ .
In particular, conclusion (i) holds if
$i<j$
. Moreover, conclusion (ii) holds if and only if
$i=j$
and
$B_i=1$
.
Proof. We can follow the proof of Theorem 11.7 almost verbatim. Note that since
$n \geq 2$
,
$S =\mathrm {Sp}_{2n}(q)$
, respectively
$\mathrm {SU}_n(q)$
with
$2 \nmid n$
, is quasisimple. We also use the fact that
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f)}$
has no nontrivial
$p'$
-quotient to show that if
$G_0$
is contained in
$E{\mathbf Z}(S)$
, then
$G_0 \leq E$
for
$E = p^{1+2nf}_+$
.
We can be much more precise in the quasisimple case:
Theorem 11.9. Let p be a prime,
$q=p^f$
,
$r \geq 2$
. Suppose that either
-
(a)
$p> 2$ ,
$n> m_1 > \ldots > m_{r} \geq 0$ with
$2|nm_1 \ldots m_r$ ,
$\gcd (n,m_1, \ldots ,m_{r})=1$ ,
$A=(q^n+1)/2$ ,
$B_i=(q^{m_i}+1)/2$ ,
$1 \leq i \leq r$ , and
or
$\chi _2$ ; or
-
(b)
$n> m_1 > \ldots > m_{r} \geq 1$ with
$2 \nmid nm_1 \ldots m_r$ ,
$\gcd (n,m_1, \ldots ,m_{r})=1$ ,
$A=(q^n+1)/(q+1)$ ,
$B_i=(q^{m_i}+1)/(q+1)$ ,
$1 \leq i \leq r$ , and
.
Recall (see [Reference Katz and TiepKT6, Theorem 11.2.3(i), (iii)]) that the local system

has
$G_{\mathrm {geom},{\mathcal F}_{\mathrm {up}}}=:G_{\mathrm {up}}$
equal to the image of
$S:=\mathrm {Sp}_{2n}(q)$
in case (a) and
$S:= \mathrm {SU}_n(q)$
in case (b), in a Weil representation of degree
$D=\mathrm {rank}({\mathcal F}_{\mathrm {up}})$
. Fix a choice of
$1 \leq i \leq j \leq r$
. If
$i=j$
, set
$d:=1$
. If
$i <j$
, set
$d:=2$
and assume
$r \ge 3$
. For f in the space
${\mathbb A}^{r-d}$
of all polynomials

denote by
${\mathcal F}(A,B_i,B_j,f,\chi )$
the local system on
${\mathbb A}^d$
whose trace function is

when
$i=j$
and

when
$i<j$
. Then we have the following results.
-
(i) If
$i=j$ , there is an open dense set
$U \subset {\mathbb A}^{r-1}$ such that for any
$f \in U$ ,
${\mathcal F}(A,B_i,B_j,f,\chi )$ has
$G_{\mathrm {geom}}$ the group
$G_{\mathrm {up}}$ .
-
(ii) In the case
$i < j$ , for any
$f \in ({\mathbb G}_m)^{r-2}$ (i.e., for any f having all coefficients nonzero),
${\mathcal F}(A,B_i,B_j,f,\chi )$ has
$G_{\mathrm {geom}}$ the group
$G_{\mathrm {up}}$ .
Proof. To prove (i), we follow the proof of Theorem 11.7 almost verbatim. In the
$\mathrm {Sp}$
case, we have
$n> m_1 >m_2 \ge 0$
, so
$n \ge 2$
, and
$\mathrm {Sp}_{2n}(q)$
is quasisimple for any odd q. In the
$\mathrm {SU}$
case, we have
$n> m_1 > m_2 \geq 1$
are all odd, so
$n \ge 5$
, and
$\mathrm {SU}_n(q)$
is again quasisimple. We also use the fact that
$G_{\mathrm {geom},{\mathcal F}(A,B_i,B_j,f,\chi )}$
is irreducible on
${\mathcal F}(A,B_i,B_j,f,\chi )$
of rank
$D>1$
to see that
$G_0$
cannot be contained in the image of
${\mathbf Z}(S)$
.
To prove (ii), we use the fact that when
$i < j$
, for any f all of whose coefficients are nonzero,
${\mathcal F}(A,B_i,B_j,f,\chi )$
has the same
$M_{2,2}$
as
${\mathcal F}_{\mathrm {up}}$
; cf. Theorem 2.3 and Corollary 2.5. The result is then immediate from Theorem 8.2 in the
$\mathrm {Sp}$
case (since
$r \geq 3$
implies
$n \geq 3$
here), and from Theorem 8.4 in the
$\mathrm {SU}$
case (since
$r \geq 3$
implies
$n \geq 7$
here).
Acknowledgements
Part of this work was done when the second author visited the Department of Mathematics, Princeton University. It is a pleasure to thank Princeton University for its generous hospitality and its stimulating environment. We also thank János Kollár for helpful conversations about intersections of Fermat hypersurfaces. We are grateful to the referees for careful reading and numerous comments which helped greatly improve the exposition of the paper.
Competing interest
The authors have no competing interests to declare.
Financial support
The second author gratefully acknowledges the support of the NSF (grants DMS-1840702 and DMS-2200850), the Simons Foundation, and the Joshua Barlaz Chair in Mathematics.