Let be a finite field of order
, and let
be an absolutely irreducible smooth projective curve defined over
(and hence over the algebraic closure
of that field). For instance,
could be the projective elliptic curve
in the projective plane , where
are coefficients whose discriminant
is non-vanishing, which is the projective version of the affine elliptic curve
To each such curve one can associate a genus
, which we will define later; for instance, elliptic curves have genus
. We can also count the cardinality
of the set
of
-points of
. The Hasse-Weil bound relates the two:
The usual proofs of this bound proceed by first establishing a trace formula of the form
for some complex numbers independent of
; this is in fact a special case of the Lefschetz-Grothendieck trace formula, and can be interpreted as an assertion that the zeta function associated to the curve
is rational. The task is then to establish a bound
for all
; this (or more precisely, the slightly stronger assertion
) is the Riemann hypothesis for such curves. This can be done either by passing to the Jacobian variety of
and using a certain duality available on the cohomology of such varieties, known as Rosati involution; alternatively, one can pass to the product surface
and apply the Riemann-Roch theorem for that surface.
In 1969, Stepanov introduced an elementary method (a version of what is now known as the polynomial method) to count (or at least to upper bound) the quantity . The method was initially restricted to hyperelliptic curves, but was soon extended to general curves. In particular, Bombieri used this method to give a short proof of the following weaker version of the Hasse-Weil bound:
Theorem 2 (Weak Hasse-Weil bound) If
is a perfect square, and
, then
.
In fact, the bound on can be sharpened a little bit further, as we will soon see.
Theorem 2 is only an upper bound on , but there is a Galois-theoretic trick to convert (a slight generalisation of) this upper bound to a matching lower bound, and if one then uses the trace formula (1) (and the “tensor power trick” of sending
to infinity to control the weights
) one can then recover the full Hasse-Weil bound. We discuss these steps below the fold.
I’ve discussed Bombieri’s proof of Theorem 2 in this previous post (in the special case of hyperelliptic curves), but now wish to present the full proof, with some minor simplifications from Bombieri’s original presentation; it is mostly elementary, with the deepest fact from algebraic geometry needed being Riemann’s inequality (a weak form of the Riemann-Roch theorem).
The first step is to reinterpret as the number of points of intersection between two curves
in the surface
. Indeed, if we define the Frobenius endomorphism
on any projective space by
then this map preserves the curve , and the fixed points of this map are precisely the
points of
:
Thus one can interpret as the number of points of intersection between the diagonal curve
and the Frobenius graph
which are copies of inside
. But we can use the additional hypothesis that
is a perfect square to write this more symmetrically, by taking advantage of the fact that the Frobenius map has a square root
with also preserving
. One can then also interpret
as the number of points of intersection between the curve
Let be the field of rational functions on
(with coefficients in
), and define
,
, and
analogously )(although
is likely to be disconnected, so
will just be a ring rather than a field. We then (morally) have the commuting square
if we ignore the issue that a rational function on, say, , might blow up on all of
and thus not have a well-defined restriction to
. We use
and
to denote the restriction maps. Furthermore, we have obvious isomorphisms
,
coming from composing with the graphing maps
and
.
The idea now is to find a rational function on the surface
of controlled degree which vanishes when restricted to
, but is non-vanishing (and not blowing up) when restricted to
. On
, we thus get a non-zero rational function
of controlled degree which vanishes on
– which then lets us bound the cardinality of
in terms of the degree of
. (In Bombieri’s original argument, one required vanishing to high order on the
side, but in our presentation, we have factored out a
term which removes this high order vanishing condition.)
To find this , we will use linear algebra. Namely, we will locate a finite-dimensional subspace
of
(consisting of certain “controlled degree” rational functions) which projects injectively to
, but whose projection to
has strictly smaller dimension than
itself. The rank-nullity theorem then forces the existence of a non-zero element
of
whose projection to
vanishes, but whose projection to
is non-zero.
Now we build . Pick a
point
of
, which we will think of as being a point at infinity. (For the purposes of proving Theorem 2, we may clearly assume that
is non-empty.) Thus
is fixed by
. To simplify the exposition, we will also assume that
is fixed by the square root
of
; in the opposite case when
has order two when acting on
, the argument is essentially the same, but all references to
in the second factor of
need to be replaced by
(we leave the details to the interested reader).
For any natural number , define
to be the set of rational functions
which are allowed to have a pole of order up to
at
, but have no other poles on
; note that as we are assuming
to be smooth, it is unambiguous what a pole is (and what order it will have). (In the fancier language of divisors and Cech cohomology, we have
.) The space
is clearly a vector space over
; one can view intuitively as the space of “polynomials” on
of “degree” at most
. When
,
consists just of the constant functions. Indeed, if
, then the image
of
avoids
and so lies in the affine line
; but as
is projective, the image
needs to be compact (hence closed) in
, and must therefore be a point, giving the claim.
For higher , we have the easy relations
The former inequality just comes from the trivial inclusion . For the latter, observe that if two functions
lie in
, so that they each have a pole of order at most
at
, then some linear combination of these functions must have a pole of order at most
at
; thus
has codimension at most one in
, giving the claim.
From (3) and induction we see that each of the are finite dimensional, with the trivial upper bound
Riemann’s inequality complements this with the lower bound
thus one has for all but at most
exceptions (in fact, exactly
exceptions as it turns out). This is a consequence of the Riemann-Roch theorem; it can be proven from abstract nonsense (the snake lemma) if one defines the genus
in a non-standard fashion (as the dimension of the first Cech cohomology
of the structure sheaf
of
), but to obtain this inequality with a standard definition of
(e.g. as the dimension of the zeroth Cech cohomolgy
of the line bundle of differentials) requires the more non-trivial tool of Serre duality.
At any rate, now that we have these vector spaces , we will define
to be a tensor product space
for some natural numbers which we will optimise in later. That is to say,
is spanned by functions of the form
with
and
. This is clearly a linear subspace of
of dimension
, and hence by Rieman’s inequality we have
Observe that maps a tensor product
to a function
. If
and
, then we see that the function
has a pole of order at most
at
. We conclude that
and in particular by (4)
We will choose to be a bit bigger than
, to make the
image of
smaller than that of
. From (6), (10) we see that if we have the inequality
(together with (7)) then cannot be injective.
On the other hand, we have the following basic fact:
Proof: From (3), we can find a linear basis of
such that each of the
has a distinct order
of pole at
(somewhere between
and
inclusive). Similarly, we may find a linear basis
of
such that each of the
has a distinct order
of pole at
(somewhere between
and
inclusive). The functions
then span
, and the order of pole at
is
. But since
, these orders are all distinct, and so these functions must be linearly independent. The claim follows.
This gives us the following bound:
Proposition 4 Let
be natural numbers such that (7), (11), (12) hold. Then
.
Proof: As is not injective, we can find
with
vanishing. By the above lemma, the function
is then non-zero, but it must also vanish on
, which has cardinality
. On the other hand, by (8),
has a pole of order at most
at
and no other poles. Since the number of poles and zeroes of a rational function on a projective curve must add up to zero, the claim follows.
If , we may make the explicit choice
and a brief calculation then gives Theorem 2. In some cases one can optimise things a bit further. For instance, in the genus zero case (e.g. if
is just the projective line
) one may take
and conclude the absolutely sharp bound
in this case; in the case of the projective line
, the function
is in fact the very concrete function
.
Remark 1 When
is not a perfect square, one can try to run the above argument using the factorisation
instead of
. This gives a weaker version of the above bound, of the shape
. In the hyperelliptic case at least, one can erase this loss by working with a variant of the argument in which one requires
to vanish to high order at
, rather than just to first order; see this survey article of mine for details.
— 1. Additional notes —
One can get a “cheap” proof of Riemann’s inequality (with the “wrong” definition of the genus ) from Cech cohomology as follows. For any natural number
, let
denote the sheaf on
, defined by setting the sections on any open set
of
to be the rational functions with no poles at
, except possibly for a pole of order up to
at
in the case that
contains
. Then
is the space of global sections on this sheaf, that is to say the zeroth Cech cohomology
. On the other hand, we also have the skyscraper sheaf
on
, defined by setting the sections on
to be
if
contains
and
otherwise. For
, we have an obvious short exact sequence of sheaves
which upon taking Cech cohomology (and using the snake lemma) gives the long exact sequence
One can compute that and
, so the long exact sequence becomes
The first part of this long exact sequence already recovers the trivial bounds (3). But it also shows that the dimensions of the spaces are non-increasing, and decrease by one precisely when the left inequality in (3) is sharp. If one then makes the non-standard definition
, then this decrease can occur at most
times, and Riemann’s inequality then follows. (However, it is not immediately obvious with this definition that
is finite; this requires some additional effort, e.g. invoking the Riemann-Hurwitz formula.) To relate this definition of
to the more usual notions of genus requires the use of Serre duality, which will not be discussed here.
The upper bound in Theorem 2 (or more precisely, a generalisation of this bound) can be converted into a comparable lower bound, namely
Theorem 5 Let
be an absolutely irreducible quasiprojective curve of bounded degree, defined over
, and let
be a perfect square. Then
where the implied constant can depend on
but is uniform in
in the limit
(holding
and
fixed).
Proof: The upper bound follows from Theorem 2, after first removing all the singularities from , normalising
to be projective, and noting that the genus does not depend on
. (Note that removing singularities only deletes
points at most.) So the remaining task is to establish the matching lower bound. Unfortunately, the trace formula (1) is not enough by itself to convert upper bounds to lower bounds, due to the case in which all the
are positive, although strangely enough it can be used for the reverse task of converting lower bounds to upper bounds. Instead, we use a Galois-theoretic argument, though for (somewhat idiosyncratic) reasons, I will disguise the Galois theory by writing everything in a geometric language rather than an algebraic one.
The basic idea is to embed (perhaps with a bounded number of points removed) as a component of a larger (non-reducible) curve
for which (a) the number of “rational points” on this larger curve
can be counted almost exactly; and (b) an upper bound similar to Theorem 2 exists for the number of “rational points” on each of the components of
, so that a lower bound can then be established by subtraction.
An easy instance of this trick arises in the case of hyperelliptic curves, which we will write affinely (deleting the point at infinity, by abuse of notation) as for some polynomial
defined over
that is not a perfect square. We can embed this in the reducible curve
where is an arbitrarily chosen quadratic non-residue of
. The curve
is the union of two curves, one of which is a dilate of the other, so they both have the same genus
, and so they each have at most
points (in the regime when
is bounded and
is a perfect square), thanks to Theorem 2. On the other hand, if
is non-zero, then exactly one of
and
is a quadratic residue, and so for all but a bounded number of
, there are exactly two values of
with
, and so
. One can then subtract the upper bound for one of the components from this estimate to obtain a matching lower bound for the other component, and in particular
.
In general we can proceed as follows. Let be the degree of
, so
. After deleting
points from
, we can view
as a degree
cover of the affine line
with
many points removed, with projection map
. Once we do so, we can pass to the lifted curve
, defined as the collection of distinct
-tuples
in
that lie over the same point in
, thus
. This is a degree
cover of (most of)
, and is defined over
. It also carries a free action of the permutation group
: if
, we define
(the inverse is there to make the action a left action; one could also work with right actions instead if desired).
The curve need not be absolutely irreducible, so we break it into absolutely irreducible components. The permutation group
acts transitively on these components, so if we pick one of these components, say
, and let
be the stabiliser, then
is a subgroup of
(in particular,
), and one can index the components of
as
, as
ranges over a set of coset representatives of
. Each fibre of
over a generic point of the base
is an orbit of
; since the map
projects
to the connected curve
, we conclude that the projection of this orbit must have cardinality
. In other words, the permutation group
is a transitive group.
Even though is defined over
, the individual components
of
need not be. In particular, the Frobenius image
of
might be another component of
, say
.
Now let be a generic point of
defined over
, then the fibre of
above
is a free
-orbit and thus has cardinality
orbit. Let
be a point in this fibre. Applying Frobenius, we see that
lies in the fibre of
above
, and hence
for a unique . Of course, if
does not lie over a point
defined over
, then the equation (13) cannot hold. Since there are
points in
with
points removed that are defined over
, and each of these give rise to
points in
, we conclude that
On the other hand, a modification of Theorem 2 shows the upper bound
for each . Strictly speaking, Theorem 2 is only directly applicable in the case when
is the identity. However, even when
is not the identity, one can “twist” the proof of Theorem 2 to establish the above estimate, basically by replacing the curve
in (2) by a twisted variant
which is still isomorphic to ; we leave the details to the interested reader. If we subtract
instances of (15) from (14), we obtain that
for each , which on combining with (15) gives
for each .
Next, as acts transtively on
, we see that there are
values of
such that
. We conclude that
But as has degree
above
, and
has degree
, we see that for generic
there are exactly
points
in
that lie above
. We conclude that
and the claim follows.
Finally, if one uses the trace formula (1), then Theorem 5 implies the bounds
for all , and hence Theorem 1. Indeed, if we let
be of maximal magnitude, thus
for all
, and an easy application of the Dirichlet approximation theorem then shows that
for infinitely many even , which when combined with Theorem 5 implies that
for infinitely many even . Taking
roots and then sending
to infinity (cf. the “tensor power trick“) we obtain (16) as required.
12 comments
Comments feed for this article
2 May, 2014 at 8:17 pm
Che!
I think you have typo “elliptic curves have genus 1”
[Corrected, thanks – T.]
9 May, 2014 at 8:39 pm
Anonymous
I still see “elliptic curves have genus 1” in the text. Why is it a typo?
[This is the corrected form of the text; previously I had written “elliptic curves had genus
“. -T.]
2 May, 2014 at 8:32 pm
Andrew S
I think there is a small typo near the beginning. Is the number of rational points
or
or
? A nice post, nonetheless. Though, if you think the definition
is wrong, then I wonder what you must think of the definition via the Hilbert polynomial!
[Corrected, thanks – T.]
3 May, 2014 at 5:42 am
Lior Silberman
There’s an arrow missing in the second version of the long exact sequence in cohomology (after the k). Also, I think “non-reducible” was supposed to be “non-irreducible”.
[Corrected, thanks – T.]
3 May, 2014 at 5:48 am
Lior Silberman
Also, in the formula after (16) you use i both as a bound and as a free variable, with the value of the free instance fixed by the previous sentence.
[Corrected, thanks – T.]
3 May, 2014 at 10:25 pm
Brian Mingus
It would be awesome if you could include a note indicating why a particular proof is interesting and important or what about it is novel for those of us more lay readers who still enjoy following your technical blog.
15 May, 2014 at 12:32 am
O Ahmadi
There is a typo in the formula for the elliptic discriminant. The coefficient of b^2 should be 27.
[Corrected, thanks – T.]
5 July, 2014 at 7:09 am
Angel Fernandez
It is true the projection in some of the elliptic curves are integers. He has demonstrated the mathematical Enfer Diez; by an equation, we can deternine whether there is an elliptical curve.
See in: Universal Journal of Computational Mathematics Vol 1 (1);2013
Title: Equation for solve elliptical curves.
14 August, 2015 at 6:38 am
משפט הסה בהסתברות גבוהה (מיני-פוסט) | One and One
[…] עם תכונות התאפסות מסויימות, ואפשר לקרוא עליה כאן או כאן. למעשה, ההוכחה שלו לא הייתה חזקה מספיק כדי לתת את הקבוע […]
17 May, 2019 at 8:53 am
A function field analogue of Riemann zeta statistics | What's new
[…] can in turn be established by elementary methods such as Stepanov’s method, as discussed in this previous post). The non-squarefree case is more complicated (and can be established using the machinery of […]
28 May, 2019 at 5:13 am
Recent Progress on a Problem of Sárközy | George Shakan
[…] now proceed to the proof of Theorem 1, which adopts the Stepanov method of auxiliary […]
18 April, 2022 at 5:55 pm
Anonymous
Legend!!