Previous set of notes: Notes 2. Next set of notes: Notes 4.
On the real line, the quintessential examples of a periodic function are the (normalised) sine and cosine functions ,
, which are
-periodic in the sense that
What about periodic functions on the complex plane? We can start with singly periodic functions which obey a periodicity relationship
for all
in the domain and some period
; such functions can also be viewed as functions on the “additive cylinder”
(or equivalently
). We can rescale
as before. For holomorphic functions, we have the following characterisations:
Proposition 1 (Description of singly periodic holomorphic functions)In both cases, the coefficients
- (i) Every
-periodic entire function
has an absolutely convergent expansion
where
is the nome
, and the
are complex coefficients such that
Conversely, every doubly infinite sequence
of coefficients obeying (2) gives rise to a
-periodic entire function
via the formula (1).
- (ii) Every bounded
-periodic holomorphic function
on the upper half-plane
has an expansion
where the
are complex coefficients such that
Conversely, every infinite sequence
obeying (4) gives rise to a
-periodic holomorphic function
which is bounded away from the real axis (i.e., bounded on
for every
).
can be recovered from
by the Fourier inversion formula
for any
in
(in case (i)) or
(in case (ii)).
Proof: If is
-periodic, then it can be expressed as
for some function
on the “multiplicative cylinder”
, since the fibres of the map
are cosets of the integers
, on which
is constant by hypothesis. As the map
is a covering map from
to
, we see that
will be holomorphic if and only if
is. Thus
must have a Laurent series expansion
with coefficients
obeying (2), which gives (1), and the inversion formula (5) follows from the usual contour integration formula for Laurent series coefficients. The converse direction to (i) also follows by reversing the above arguments.
For part (ii), we observe that the map is also a covering map from
to the punctured disk
, so we can argue as before except that now
is a bounded holomorphic function on the punctured disk. By the Riemann singularity removal theorem (Exercise 35 of 246A Notes 3)
extends to be holomorphic on all of
, and thus has a Taylor expansion
for some coefficients
obeying (4). The argument now proceeds as with part (i).
The additive cylinder and the multiplicative cylinder
can both be identified (on the level of smooth manifolds, at least) with the geometric cylinder
, but we will not use this identification here.
Now let us turn attention to doubly periodic functions of a complex variable , that is to say functions
that obey two periodicity relations
Within the world of holomorphic functions, the collection of doubly periodic functions is boring:
Proposition 2 Letbe an entire doubly periodic function (with periods
linearly independent over
). Then
is constant.
In the language of Riemann surfaces, this proposition asserts that the torus is a non-hyperbolic Riemann surface; it cannot be holomorphically mapped non-trivially into a bounded subset of the complex plane.
Proof: The fundamental domain (up to boundary) enclosed by is compact, hence
is bounded on this domain, hence bounded on all of
by double periodicity. The claim now follows from Liouville’s theorem. (One could alternatively have argued here using the compactness of the torus
.
To obtain more interesting examples of doubly periodic functions, one must therefore turn to the world of meromorphic functions – or equivalently, holomorphic functions into the Riemann sphere . As it turns out, a particularly fundamental example of such a function is the Weierstrass elliptic function
— 1. Doubly periodic functions —
Throughout this section we fix two complex numbers that are linearly independent over
, which then generate a lattice
.
We now study the doubly periodic meromorphic functions with respect to these periods that are not identically zero. We first observe some constraints on the poles of these functions. Of course, by periodicity, the poles will themselves be periodic, and thus the set of poles forms a finite union of disjoint cosets of the lattice
. Similarly, the zeroes form a finite union of disjoint cosets
. Using the residue theorem, we can obtain some further constraints:
Lemma 3 (Consequences of residue theorem) Letbe a doubly periodic meromorphic function (not identically zero) with periods
, poles at
, and zeroes at
.
- (i) The sum of residues at each
(i.e., we sum one residue per coset) is equal to zero.
- (ii) The number of poles
(counting multiplicity, but only counting once per coset) is equal to the number of zeroes
(again counting multiplicity, and once per coset).
- (iii) The sum of the poles
(counting multiplicity, and working in the group
) is equal to the sum of the zeroes
.
Proof: For (i), we first apply a translation so that none of the pole cosets intersects the fundamental parallelogram boundary
; this of course does not affect the sum of residues. Then, by the residue theorem, the sum in (i) is equal to the expression
For part (iii), we again translate so that none of the pole or zero cosets intersects , noting from part (ii) that any such translation affects the sum of poles and sum of zeroes by the same amount. By the residue theorem, it now suffices to show that
This lemma severely limits the possible number of behaviors for the zeroes and poles of a meromorphic function. To formalise this, we introduce some general notation:
Definition 4 (Divisors)
- (i) A divisor on the torus
is a formal integer linear combination
, where
ranges over a finite collection of points in the torus
(i.e., a finite collection of cosets
), and
are integers, with the obvious additive group structure; equivalently, the space
of divisors is the free abelian group with generators
for
(with the convention
).
- (ii) The number
is the degree
of a divisor
, the point
is the sum
of
, and each
is the order
of the divisor at
(with the convention that the order is
if
does not appear in the sum). A divisor is non-negative (or effective) if
for all
. We write
if
is non-negative (i.e., the order of
is greater than or equal to that of
at every point
, and
if
and
.
- (iii) Given a meromorphic function
(or equivalently, a doubly periodic function
) that is not identically zero, the principal divisor
is the divisor
, where
ranges over the zeroes and poles of
, and
is the order of the zero (if
is a zero) or negative the order of the pole (if
is a pole).
- (iv) Given a divisor
, we define
to be the space of all meromorphic functions
that are either zero, or are such that
. That is to say,
consists of those meromorphic functions that have at most a pole of order
at
if
is positive, or at least zero of order
if
is negative.
A divisor can be viewed as an abstraction of the concept of a set of zeroes and poles (counting multiplicity). Observe that principal divisors obey the laws ,
when
are meromorphic and non-zero. In particular, the space
of principal divisors is a subgroup of the space
of all divisors. By Lemma 3(ii), all principal divisors have degree zero, and from Lemma 3(iii), all principal divisors have sum zero as well. Later on we shall establish the converse claim that every divisor of degree and sum zero is a principal divisor; see Exercise 7.
Remark 5 One can define divisors on other Riemann surfaces, such as the complex plane. Observe from the fundamental theorem of algebra that if one has two non-zero polynomials
, then
if and only if
divides
as a polynomial. This may give some hint as to the origin of the terminology “divisor”. The machinery of divisors turns out to have a rich algebraic and topological structure when applied to more general Riemann surfaces than tori, for instance enabling one to associate an abelian variety (the Jacobian variety) to every algebraic curve; see these 246C notes for further discussion.
It is easy to see that is always a vector space. All non-zero meromorphic functions
belong to at least one of the
, namely
, so to classify all the meromorphic functions on
, it would suffice to understand what all the spaces
are.
Liouville’s theorem (in the form of Proposition 2) tells us that all elements of – that is to say, the holomorphic functions on
– are constant; thus
is one-dimensional. If
is a negative divisor, the elements of
are thus constant and have at least one zero, thus in these cases
is trivial.
Now we gradually work our way up to higher degree divisors . A basic fact, proven from elementary linear algebra, is that every time one adds a pole to
, the dimension of the space
only goes up by at most one:
Lemma 6 For any divisorand any
,
is a subspace of
of codimension at most one. In particular,
is finite-dimensional for any
.
Proof: It is clear that is a subspace of
. If
has order
at
, then there is a linear functional
that assigns to each meromorphic function
the
coefficient of the Laurent expansion of
at
(note from periodicity that the exact choice of coset representative
is not relevant. A little thought reveals that the kernel of
is precisely
, and the first claim follows. The second claim follows from iterating the first claim, noting that any divisor
can be obtained from a suitable negative divisor by the addition of finitely many poles
.
Now consider the space for some point
. Lemma 6 tells us that the dimension of this space is either one or two, since
was one-dimensional. The space
consists of functions
that possibly have a simple pole at most at
, and no other poles. But Lemma 3(i) tells us that the residue at
has to vanish, and so
is in fact in
and thus is constant. (One could also argue here using the other two parts of Lemma 2; how?) So
is no larger than
, and is thus also one-dimensional.
Now let us study the space – the space of meromorphic functions that have at most a double pole at
and no other poles. Again, Lemma 6 tells us that this space is one or two dimensional. To figure out which, we can normalise
to be the origin coset
. The question is now whether there is a doubly periodic meromorphic function that has a double pole at each point of
. A naive candidate for such a function would be the infinite series
Now we show that is doubly periodic, thus
and
for
. We just prove the first identity, as the second is analogous. From (6) we have
By construction, lies in
, and is clearly non-constant. Thus
is two-dimensional, being spanned by the constant function
and
. By translation, we see that
is two-dimensional for any other point
as well.
From (6) it is also clear that the function is even:
. In particular, for any
avoiding the half-lattice
(so that
and
occupy different locations in the torus
), the function
has a zero at both
and
. By Lemma 3(ii) there are no other zeroes of this function (and this claim is also consistent with Lemma 3(iii)); thus the divisor
of this function is given by
Exercise 7 (Classification of principal divisors)
- (i) Let
be four points
such that
. Show that the divisor
is a principal divisor. (Hint: if
are all distinct, use a function such as
If some of the
coincide, use some transformed version of the Weierstrass elliptic function
instead.)
- (ii) Show that every divisor of degree zero and sum zero is a principal divisor.
- (iii) Two divisors are said to be equivalent if their difference is a principal divisor. Show that two divisors are equivalent if and only if they have the same degree and same sum.
- (iv) Show that the quotient group
(known as the divisor class group or Picard group) is isomorphic (as a group) to
, and that the subgroup
arising from degree zero divisors (also known as the Jacobian variety of
) is isomorphic to
.
Now let us study the space , where we again normalise
for sake of discussion. Lemma 6 tells us that this space is two or three dimensional, being spanned by
,
, and possibly one other function. Note that the derivative
of the meromorphic function
is also doubly periodic with a triple pole at
, so it lies in
and is not a linear combination of
or
(as these have a lower order singularity at
). Thus
is three-dimensional, being spanned by
. A formal term-by-term differentiation of (6) gives (7). To justify (7), observe that the arguments that demonstrated the meromorphicity of the right-hand side of (6) also show the meromorphicity of (7). From Fubini’s theorem, the fundamental theorem of calculus, and (6) we see that
Turning now to , we could differentiate
yet again to generate a doubly periodic function
with a fourth order pole at the origin, but we can also work with the square
of the Weierstrass function. From Lemma 6 we conclude that
is four-dimensional and is spanned by
. In a similar fashion,
is a five-dimensional space spanned by
.
Something interesting happens though at . Lemma 6 tells us that this space is the span of
, and possibly one other function, which will have a pole of order six at the origin. Here we have two natural candidates for such a function: the cube
of the Weierstrass function, and the square
of its derivative. Both have a pole of order exactly six and lie in
, and so
must be a linear combination of
. But since
are even and
are odd,
must in fact just be a linear combination of
. To work out the precise combination, we see by repeating the derivation of (7) that
Exercise 8 Derive (8) directly from Proposition 2 by showing that the difference between the two sides is doubly periodic and holomorphic after removing singularities.
Exercise 9 (Classification of doubly periodic meromorphic functions)
- (i) For any
, show that
has dimension
, and every element of this space is a polynomial combination of
.
- (ii) Show that every doubly periodic meromorphic function is a rational function of
.
We have an alternate form of (8):
Exercise 10 Define the roots,
,
.
- (i) Show that
are distinct, and that
for all
. (Hint: use (10).) Conclude in particular that
,
, and
.
- (ii) Show that the modular discriminant
is equal to
, and is in particular non-zero.
If we now define the elliptic curve
Lemma 11 The mapdefined by (11) is a bijection between
and
.
Among other things, this lemma implies that the elliptic curve is topologically equivalent (i.e., homeomorphic to) a torus, which is not an entirely obvious fact (though if one squints hard enough, the real analogue of an elliptic curve does resemble a distorted slice of a torus embedded in
).
Proof: Clearly is the only point that maps to
, and (from (10)) the half-periods are the only points that map to
. It remains to show that all the other points
arise via
from exactly one element of
. The function
has exactly two zeroes by Lemma 3(ii), which lie at
for some
as
is even; since
,
is not equal to
, hence
is not a half-period. As
is odd, the map (11) must therefore map
to the two points
of the elliptic curve
that lie above
, and the claim follows.
Analogously to the Riemann sphere , the elliptic curve
can be given the structure of a Riemann surface, by prescribing the following charts:
- (i) When
is a point in
other than
or
, then locally
is the graph of a holomorphic branch
of the square root of
near
, and one can use
as a coordinate function in a sufficiently small neighbourhood of
.
- (ii) In the neighbourhood of
for some
, the function
has a simple zero at
and so has a local inverse
that maps a neighbourhood of
to a neighbourhood of
, and a point
sufficiently near
can be parameterised by
. One can then use
as a coordinate function in a neighbourhood of
.
- (iii) A neighbourhood of
consists of
and the points
in the remaining portion of
with
sufficiently large; then
is asymptotic to a square root of
, so in particular
and
should both go to zero as
goes to infinity in
. We rewrite the defining equation
of the curve in terms of
and
as
. The function
has a simple zero at zero and thus has a holomorphic local inverse
that maps
to
, and we have
in a neighbourhood of infinity. We can then use
as a coordinate function in a neighbourhood of
, with the convention that this coordinate function vanishes at infinity.
It is then a tedious but routine matter to check that has the structure of a Riemann surface. We then claim that the bijection
defined by (11) is holomorphic, and thus a complex diffeomorphism of Riemann surfaces. In the neighbourhood of any point
of the torus
other than the origin
,
maps to a neighbourhood of finite point
of
, including the three points
, the holomorphicity is a routine consequence of composing together the various local holomorphic functions and their inverses. In the neighbourhood of the origin
,
maps
for small
to a point of
with a Laurent expansion
While we have shown that all tori are complex diffeomorphic to elliptic curves, the converse statement that all elliptic curves are diffeomorphic to tori will have to wait until the next section for a proof, once we have set up the machinery of modular forms.
Exercise 12 (Group law on elliptic curves)
- (i) Let
be three distinct elements of the torus
that are not equal to the origin
. Show that
if and only if the three points
,
,
are collinear in
, in the sense that they lie on a common complex line
for some complex numbers
with
not both zero.
- (ii) What happens in (i) if (say)
and
agree? What about if
?
- (iii) Using (i), (ii), give a purely geometric definition of a group addition law on the elliptic curve
which is compatible with the group addition law on the torus
via (11). (We remark that the associativity property of this law is not obvious from a purely geometric perspective, and is related to the Cayley-Bacharach theorem in classical geometry; see this previous blog post.)
Exercise 13 (Addition law) Show that for anywith
, one has
Exercise 14 (Special case of Riemann-Roch)
- (i) Show that if two divisors
are equivalent (in the sense of Exercise 7(iii)), then the vector spaces
and
are isomorphic (in particular, they have the same dimension).
- (ii) If
is a divisor of some degree
, show that the dimension of the space
is zero if
, equal to
if
, equal to
if
and
has non-zero sum, and equal to
if
and
has zero sum. (Hint: use Exercise 7(iii) and part (i) to replace
with an equivalent divisor of a simple form.)
- (iii) Verify the identity
for any divisor
. This is a special case of the more general Riemann-Roch theorem, discussed in these 246C notes.
Exercise 15 (Elliptic integrals)
Remark 16 The integralis an example of an elliptic integral; many other elliptic integrals (such as the integral arising when computing the perimeter of an ellipse) can be transformed into this form (or into a closely related integral) by various elementary substitutions. Thus the Weierstrass elliptic function
can be employed to evaluate elliptic integrals, which may help explain the terminology “elliptic” that occurs throughout these notes. In 246C notes we will introduce the notion of a meromorphic
-form on a Riemann surface. The identity (12) can then be interpreted in this language as the differential form identity
, where
are the standard coordinates on the elliptic curve
; the meromorphic
-form is initially only defined on
outside of the four points
, but this identity in fact reveals that the form extends holomorphically to all of
; it is an example of what is known as an Abelian differential of the first kind.
Remark 17 The elliptic curve(for various choices of parameters
) can be defined in other fields than the complex numbers (though some technicalities arise in characteristic two and three due to the pathological behaviour of the discriminant in those cases). On the other hand, the Weierstrass elliptic function
is a transcendental function which only exists in complex analysis and does not have a direct analogue in other fields. So this connection between elliptic curves and tori is specific to the complex field. Nevertheless, many facts about elliptic curves that were initially discovered over the complex numbers through this complex-analytic link to tori, were then reproven by purely algebraic means, so that they could be extended without much difficulty to many other fields than the complex numbers, such as finite fields. (For instance, the role of the complex torus can be replaced by the Jacobian variety, which was briefly introduced in Exercise 7.) Elliptic curves over such fields are of major importance in number theory (and cryptography), but we will not discuss these topics further here.
— 2. Modular functions and modular forms —
In Exercise 32 of 246A Notes 5, it was shown that two tori and
are complex diffeomorphic if and only if one has
Let us write for the set of all tori
quotiented by the equivalence relation of complex diffeomorphism; this is the (classical, level one, noncompactified) modular curve. By the above discussion, this set can also be identified with the set of pairs
of linearly independent (over
) complex numbers quotiented by the equivalence relation given implicitly by (13). One can simplify this a little by observing that any pair
is equivalent to
for some
in the upper half-plane
, namely either
or
depending on the relative phases of
and
; this quantity
is known as the period ratio. From (13) (swapping the roles of
as necessary), we then see that two pairs
are equivalent if one has
If we use the relation to write
Exercise 18 Suppose thatis an element of
which is fixed by some element
of
which is not the identity or negative identity. Let
be the lattice
.
- (i) Show that
obeys a dilation invariance
for some complex number
which is not real.
- (ii) Show that the dilation
in part (i) must have magnitude one. (Hint: look at a non-zero element of
of minimal magnitude.)
- (iii) Show that there is no rotation invariance
with
. (Hint: again, work with a non-zero element of
of minimal magnitude, and use the fact that
is closed under addition and subtraction. It may help to think geometrically and draw plenty of pictures.)
- (iv) Show that
is equivalent to either the Gaussian lattice
or the Eisenstein lattice
, and conclude that the period ratio
is equivalent to either
or
.
Remark 19 The conformal mapon the complex numbers preserves the Gaussian integers
and thus descends to a conformal map from the Gaussian torus
to itself; similarly the conformal map
preserves the Eisenstein integers and thus descends to a conformal map from the Eisenstein torus
to itself. These rare examples of complex tori equipped with additional conformal automorphisms are examples of tori (or elliptic curves) endowed with complex multiplication. There are additional examples of elliptic curves endowed with conformal endomorphisms that are still considered to have complex multiplication, and have a particularly nice algebraic number theory structure, but we will not pursue this topic further here.
Remark 20 The fact that the action ofon lattices contains fixed points is somewhat annoying, as it prevents one from immediately viewing the modular curve as a Riemann surface. However by passing to a suitable finite index subgroup of
, one can remove these fixed points, leading to a theory that is cleaner in some respects. For instance, one can work with the congruence group
, which roughly speaking amounts to decorating the lattices
(or their tori
) with an additional “
-marking” that eliminates the fixed points. This leads to a modification of the theory which is for instance well suited for studying theta functions; the role of the
-invariant in the discussion below is then played by the modular lambda function
, which also gives a uniformisation of the twice-punctured complex plane
. However we will not develop this parallel theory further here.
If we let be the elements
of
not equivalent to
or
, and
the equivalence class of tori not equivalent to the Gaussian torus
or the Eisenstein torus
, then
can be viewed as the quotient
of the Riemann surface
by the free and proper action of
, so it has the structure of a Riemann surface;
can thus be thought of as the Riemann surface
with two additional points added. Later on we will also add a third point
(known as the cusp) to the Riemann surface to compactify it to
.
A function on the modular curve
can be thought of, equivalently, as a function
that is
-invariant in the sense that
for all
and
, or equivalently that one has the identity
We define a modular function to be a meromorphic function on
that obeys the condition (15), and which also has at most polynomial growth at the cusp
in the sense that one has a bound of the form
Exercise 21
- (i) Let
be two elements of
with
. Show that it is possible to transform the quadruplet
to the quadruplet
after a finite number of applications of the moves
and
({Hint: use the principle of infinite descent, applying the moves in a suitable order to decrease the lengths of
and
when the dot product
is not too small, taking advantage of the Lagrange identity
to determine when this procedure terminates. It may help to think geometrically and draw plenty of pictures.) Conclude that the two matrices (17) generate all of
.
- (ii) Show that a function
obeys (15) if and only if it obeys both (18) and (19).
Exercise 22 (Standard fundamental domain) Define the standard fundamental domainfor
to be the set
- (i) Show that every lattice
is equivalent (up to dilations) to a lattice
with
, with
unique except when it lies on the boundary of
, in which case the lack of uniqueness comes either from the pair
for some
, or from the pair
for some
. (Hint: arrange
so that
is a non-zero element of
of minimal magnitude.)
- (ii) Show that
can be identified with the fundamental domain
after identifying
with
for
, and
with
for
. Show also that the set
is then formed the same way, but first deleting the points
from
.
We will give some examples of modular functions (beyond the trivial example of constant functions) shortly, but let us first observe that when one differentiates a modular function one gets a more general class of function, known as a modular form. In more detail, observe from (14) that the derivative of the Möbius transformation is
, and hence by the chain rule and (15) the derivative of a modular function
would obey the variant law
Exercise 23 Letbe a natural number. Show that a function
obeys (20) if and only if it is
-periodic in the sense of (18) and obeys the law
for all
.
Exercise 24 (Lattice interpretation of modular forms) Letbe a modular form of weight
. Show that there is a unique function
from lattices
to complex numbers such that
for all
, and such that one has the homogeneity relation
for any lattice
and non-zero complex number
.
Observe that the product of a modular form of weight and a modular form of weight
is a modular form of weight
, and that the ratio of two modular forms of weight
will be a modular function (if the denominator is not identically zero). Also, the space of modular forms of a given weight is a vector space, as is the space of modular functions. This suggests a way to generate non-trivial modular functions, by first locating some modular forms and then taking suitable rational combinations of these forms.
Somewhat analogously to how we used Lemma 3 to investigate the spaces for divisors
on a torus, we will investigate the space of modular forms via the following basic formula:
Theorem 25 (Valence formula) Letbe a modular form of weight
, not identically zero. Then we have
where
is the order of vanishing of
at
,
is the order of vanishing of
(i.e.,
viewed as a function of the nome
) at
, and
ranges over the zeroes of
that are not equivalent to
, with just one zero counted per equivalence class. (This will be a finite sum.)
Informally, this formula asserts that the point only “deserves” to be counted in
with multiplicity
due to its order
stabiliser, while the point
only “deserves” to be counted in
with multiplicity
due to its order
stabiliser. (The cusp
has an infinite stabiliser, but this is compensated for by taking the order with respect to the nome variable
rather than the period ratio variable
.) The general philosophy of weighting points by the reciprocal of the order of their stabiliser occurs throughout mathematics; see this blog post for more discussion.
Proof: Firstly, from Exercise 22, we can place all the zeroes in the fundamental domain
. When parameterised in terms of the nome
, this domain is compact, hence has only finitely many zeros, so the sum in (22) is finite.
As in the proof of Lemma 3(ii), we use the residue theorem. For simplicity, let us first suppose that there are no zeroes on the boundary of the fundamental domain except possibly at the cusp
. Then for
large enough, we have from the residue theorem that
Suppose now that there is a zero on the right edge of
, and hence also on the left edge
by periodicity, for some
. One can account for this zero by perturbing the contour
to make a little detour to the right of
(e.g., by a circular arc), and a matching detour to the right of
. One can then verify that the same argument as before continues to work, with this boundary zero being counted exactly once. Similarly, if there is a zero on the left arc
for some
, and hence also at
by modularity, one can make a detour slightly above
and slightly below
(with the two detours being related by the transform
to ensure cancellation), and again we can argue as before. If instead there is a zero at
, one makes an (approximately) semicircular detour above
; in this case the detour does not cancel out, but instead contributes a factor of
in the limit as the radius of the detour goes to zero. Finally, if there is a zero at
(and hence also at
), one makes detours by two arcs of angle approximately
at these two points; these two (approximate) sixth-circles end up contributing a factor of
in the limit, giving the claim.
Exercise 26 (Quick applications of the valence formula)
- (i) Let
be a modular form of weight
, not identically zero. Show that
is equal to
or an even number that is at least
.
- (ii) (Liouville theorem for
) If
is a modular form of weight zero, show that it is constant. (Hint: apply the valence theorem to various shifts
of
by constant.)
- (iii) For
, show that the vector space of modular forms of weight
is at most one dimensional. (Hint: in these cases, there are a very limited number of solutions to the equation
with
natural numbers.)
- (iv) Show that there are no cusp forms of weight
when
or
, and for
the space of cusp forms of weight
is at most one dimensional.
- (v) Show that for any
, the space of cusp forms of weight
is a subspace of the space of modular forms of weight
of codimension at most one, and that both spaces are finite-dimensional.
A basic example of modular forms are provided by the Eisenstein series
that we have already encountered for even integers
Exercise 27 Give an alternate proof thatis a cusp form, not using the valence identity, by first establishing that
and
.
We can now create our first non-trivial modular function, the -invariant
Lemma 28 We haveand
.
Proof: Using the rotation symmetry we see that
, hence
which implies that
and hence
. Similarly, using the rotation symmetry
we have
, hence
. (One can also use the valence formulae to get the vanishing
).
Being modular, we can think of as a map from
to
. We have the following fundamental fact:
Proposition 29 The mapis a bijection.
Proof: Note that for any ,
if and only if
is a zero of
. It thus suffices to show that for every
, the zeroes of the function
in
consist of precisely one orbit of
. This function is a modular form of weight
that does not vanish at infinity (since
does not vanish while
does). By the valence formula, we thus have
-
has a simple zero at precisely one
-orbit, not equivalent to
or
.
-
has a double zero at
(and equivalent points), and no other zeroes.
-
has a triple zero at
(and equivalent points), and no other zeroes.
Note that this proof also shows that has a double zero at
and
has a triple zero at
, but that
has a simple zero for any
not equivalent to
or
.
We can now give the entire modular curve the structure of a Riemann surface by declaring
to be the coordinate function. This is compatible with the existing Riemann surface structure on
since
was already holomorphic on this portion of the curve. Any modular function
can then factor as
for some meromorphic function
that is initially defined on the punctured complex plane
; but from meromorphicity of
on
and at infinity we see that
blows up at an at most polynomial rate as one approaches
,
, or
, and so
is in fact a meromorphic function on the entire Riemann sphere and is thus a rational function (Exercise 19 of 246A Notes 4). We conclude
Proposition 30 Every modular function is a rational function of the-invariant
.
Conversely, it is clear that every rational function of is modular, thus giving a satisfactory description of the modular functions.
Exercise 31 Show that every modular function is the ratio of two modular forms of equal weight (with the denominator not identically zero).
Exercise 32 (All elliptic curves are tori) Letbe two complex numbers with
. Show that there is a lattice
such that
and
, so in particular the elliptic curve
is complex diffeomorphic to a torus
.
Remark 33 By applying some elementary algebraic geometry transformations one can show that any (smooth, irreducible) cubic plane curvegenerated by a polynomial
of degree
is a Riemann surface complex diffeomorphic to a torus
after adding some finite number of points at infinity; also, some degree
curves such as
can also be placed in this form. However we will not detail the required transformations here.
A famous application of the theory of the -invariant is to give a short Riemann surface-based proof of the the little Picard theorem (first proven in Theorem 55 of 246A Notes 4):
Theorem 34 (Little Picard theorem) Letbe entire and non-constant. Then
omits at most one point of
.
Proof: Suppose for contradiction that omits at least two points of
. By applying a linear transformation, we may assume that
omits the points
and
. Then
is a holomorphic function from
to
. Since the domain
is simply connected,
lifts to a holomorphic function from
to
. Since
is complex diffeomorphic to a disk, this lift must be constant by Liouville’s theorem, hence
is constant as required. (This is essentially Picard’s original proof of this theorem.)
The great Picard theorem can also be proven by a more sophisticated version of these methods, but it requires some study of the possible behavior of elements of ; see Exercise 37 below.
All modular forms are -periodic, and hence by Proposition 1 should have a Fourier expansion, which is also a Laurent expansion in the nome. As it turns out, the Fourier coefficients often have a highly number-theoretic interpretation. This can be illustrated with the Eisenstein series
; here we follow the treatment in Stein-Shakarchi. To compute the Fourier coefficients we first need a computation:
Exercise 35 Letand
, and let
be the nome. Establish the identity
in two different ways:
- (i) By applying the Poisson summation formula (Proposition 3(v) of Notes 2).
- (ii) By first establishing the identity
by applying Proposition 1 to the difference of the two sides, and differentiating in
. (It is also possible to establish (27) from differentiating and then manipulating the identities in Exercises 25 or 27 of Notes 1.)
From (25), (26) (and symmetry) one has
Remark 36 If one expands out a few more terms in the above expansions, one can calculateThe various coefficients in here have several remarkable algebraic properties. For instance, applying this expansion at
for a natural number
, so that
, one obtains the approximation
Now for certain values of
, most famously
, the torus
admits a complex multiplication that allows for computation of the
-invariant by algebraic means (think of this as a more advanced version of Lemma 28; it is closely related to the fact that the ring of algebraic integers in
admit unique factorisation, see these previous notes for some related discussion). For instance, one can eventually establish that
which eventually leads to the famous approximation
(first observed by Hermite, but also attributed to Ramanujan via an April Fools’ joke of Martin Gardner) which is accurate to twelve decimal places. The remaining coefficients have a remarkable interpretation as dimensions of components of a certain representation of the monster group known as the moonshine module, a phenomenon known as monstrous moonshine. For instance, the smallest irreducible representation of the monster group has dimension
, precisely one less than the
coefficient of
. The Fourier coefficients
of the (normalised) modular discriminant,
form a sequence known as the Ramanujan
function and obeys many remarkable properties. For instance, there is the totally non-obvious fact that this function is multiplicative in the sense that
whenever
are coprime; see Exercise 43.
Exercise 37 (Great Picard theorem)
- (i) Show that every fractional linear transformation
on
with
,
is either of finite order (elliptic case), conjugate to a translation
for some
after conjugating by another fractional linear transformation (parabolic case), or conjugate to a dilation
for some
after conjugating by another fractional linear transformation (hyperbolic case). (Hint: study the eigenvalues and eigenvectors of
, based on the value of the trace
and in particular whether the magnitude of the trace is less than two, equal to two, or greater than two. Note that the trace also has to be an integer.)
- (ii) Let
be holomorphic. Show that there exists a holomorphic function
such that
for all
, as well as a fractional linear transformation
with
and
such that
for all
.
- (iii) If the transformation
in (ii) is in the elliptic case of (i), show that
is bounded in a neighbourhood of
, and hence has a removable singularity at the origin. (Hint:
will have some finite period and can thus be studied using Proposition 1 after applying a Möbius transform to map
to a disk.)
- (iv) If the transformation
in (ii) is in the hyperbolic case of (i), show that
is bounded in a neighbourhood of
, and hence has a removable singularity at the origin. (Hint: The standard branch of
maps
to an annulus, and is invariant with respect to the dilation action
. Use this to create a bounded
-periodic holomorphic function on
.)
- (v) If the transformation
in (ii) is in the parabolic case of (i), show that
exhibits at most polynomial growth as one approaches
, and hence has at most a pole at the origin. (Hint: If for instance
, then
is
-periodic and takes values in
, and one can now repeat the arguments of (iii). Also use the expansion (28).)
- (vi) Use the previous parts of this exercise to give another proof of the great Picard theorem (Theorem 56 of 245A Notes 4): the image of a holomorphic function in a punctured disk
with an essential singularity at
omits at most one value of
.
Exercise 38 (Dimension of space of modular forms)
- (i) If
is an even natural number, show that the dimension of the space of modular forms of weight
is equal to
except when
is equal to
mod
, in which case it is equal to
. (Hint: for
this follows from Exercise 26; to cover the larger ranges of
, use the modular discriminant
to show that the space of cusp forms of weight
is isomorphic to the space of modular forms of weight
.
- (ii) If
is an even natural number, show that a basis for the space of modular forms of weight
is provided by the powers
where
range over natural numbers (including zero) with
.
Thus far we have constructed modular forms and modular functions starting from Eisenstein series . There is another important, and seemingly quite different, way to generate modular forms coming from theta functions. Typically these functions are not quite modular in the sense given in these notes, but are close enough that after some manipulation one can transform theta functions into modular forms. The simplest example of a theta function is the Jacobi theta function
Exercise 39 Define the Dedekind eta functionby the formula
or in terms of the nome
![]()
where
is one of the
roots of
.
- (i) Establish the modified
-periodicity
and the modified modularity
using the standard branch of the square root. (Hint: a direct application of Poisson summation applied to
gives a sum that looks somewhat like
but with different numerical constants (in particular, one sees terms like
instead of
arising). Split the index of summation
into three components
,
,
based on the residue classes modulo
and rearrange each component separately.)
- (ii) Establish the identity
(Hint: show that both sides are cusp forms of weight
that vanish like
near the cusp.)
Remark 40 The relationship betweenand the
power of the eta function can be interpreted (after some additional effort) as a relation
between the modular discriminant
and the theta function
of a certain highly symmetric
-dimensional lattice
known as the Leech lattice, but we will not pursue this connection further here.
The function has a remarkable factorisation coming from Euler’s pentagonal number theorem
Theorem 41 (Jacobi triple product identity) For anyand
, one has
Observe that by replacing by
and
with
we have
Proof: Let us denote the left-hand side and right-hand side of (33) by and
respectively. For fixed
, both sides are clearly holomorphic in
, with
. Our strategy in showing that
and
agree (following Stein-Shakarchi) is to first observe that they have many of the same periodicity properties. We clearly have
-periodicity
Remark 42 Another equivalent form of (32) iswhere
is the partition function of
– the number of ways to represent
as the sum of positive integers (up to rearrangement). Among other things, this formula can be used to ascertain the asymptotic growth of
(which turns out to roughly be of the order of
, as famously established by Hardy and Ramanujan).
Theta functions can be used to encode various number-theoretic quantities involving quadratic forms, such as sums of squares. For instance, from (30) and collecting terms one obtains the formula
Exercise 43 (Hecke operators) Letbe a natural number.
Simultaneous eigenfunctions of the Hecke operators are known as Hecke eigenfunctions and are of major importance in number theory.
- (i) If
is a modular form of weight
, and
is the corresponding function on lattices given by Exercise 24, and
is a positive natural number, show that there is a unique modular form
of weight
whose corresponding function
on lattices is related to
by the formula
where the sum ranges over all sublattices
of
whose index
is equal to
. Show that
is a linear operator on the space of weight
modular forms that also maps the space of weight
cusp forms to itself; this operator is known as a Hecke operator.
- (ii) Give the more explicit formula
- (iii) Show that the Hecke operators all commute with each other, thus
whenever
is a modular form of weight
and
are positive natural numbers. Furthermore show that
if
are coprime.
- (iv) If
is a modular form of weight
with Fourier expansion
, show that for any coprime positive integers
that the
coefficient of
is equal to
.
- (v) Establish the multiplicativity
of the Ramanujan tau function (the Fourier coefficients of the modular discriminant). (Hint: use the one-dimensionality of the space of cusp forms of weight
to conclude that
is a simultaneous eigenfunction of the Hecke operators.)
55 comments
Comments feed for this article
2 February, 2021 at 9:18 pm
On the notes on applications
Would number theoretic and sphere packing applications be covered in the class and would the notes be posted?
[No, I am not planning to cover those topics in this course. -T]
3 February, 2021 at 2:12 am
Anonymous
The top-level tags of this post appear to have the wrong course tagging.
[Fixed, thanks – T.]
3 February, 2021 at 2:47 am
Anonymous
typo in Lemma 6: P should be in the torus, not in \Lambda
[Corrected, thanks – T.]
3 February, 2021 at 2:51 am
Anonymous
Lemma 3 is referred to as Lemma 2.
[Corrected – T]
3 February, 2021 at 6:58 am
Anonymous
I think there’s a typo in the first paragraph: \omega Z\C should be C/\omega Z
[Both forms are acceptable. In the case of quotients of actions by abelian groups it makes little difference, but when it comes to nonabelian groups such as ${\bf SL}_2({\bf Z})$ the convention matters a little bit, and I have elected to have the groups act on the left and define quotients accordingly. -T]
3 February, 2021 at 12:59 pm
Anonymous
The product representation of
function is closely related to the product representation of the generating function for the partition function
(and can be used to derive its convergent asymptotic series)
[A remark to this effect has been added, thanks – T.]
4 February, 2021 at 12:13 pm
Anonymous
The identity (8) can also be proved by observing (via Laurent expansion coefficients) that its LHS is entire – so by proposition 2 it is a constant.
4 February, 2021 at 12:22 pm
Anonymous
Correction: I meant that the difference(!) between the LHS and the RHS (not the LHS) of (8) is entire – so it must be a constant.
[An exercise to this effect has been added – T.]
4 February, 2021 at 5:15 pm
Lior Silberman
For a different approach to Exercise 17, note that if
fixes a point in
then it lies in the intersection of a discrete group and a compact group, hence had finite order.
Its eigenvalues must then be roots of unity which are also quadratic irrationalities. Since a primitive root of unity of order
has degree
, we see that the order must be one of 1,2,3,4,6. This computes the cojugacy classes of these elements.
Also, dividing by
which act trivially we see that
has order 1,2,3 in
.
5 February, 2021 at 9:31 am
Terence Tao
Thanks for this! One can also see that an element of
that has roots of unity as eigenvalues must have trace in the set
which also gives the eigenvalues and thus the conjugacy class in
. But I was not able to easily show that any two such elements of
with the same eigenvalues were conjugate in
(which is what one needs to specify the fixed points) without messing around with lattices as in the stated exercise. (Maybe this is just some consequence of the algebraic integer nature of the eigenvalues, but I didn’t immediately see an argument.)
11 February, 2021 at 6:24 am
Jarek Kuben
This argument can be finished using the shape of the standard fundamental domain. With notation as above, if
for
, then
and we can WLOG assume
by replacing
with
if necessary. This equation also implies that
is an eigenvector of
with eigenvalue
, and since
, it’s
. But if we also WLOG assume
, then it follows that
.
12 February, 2021 at 3:53 pm
Lior Silberman
Let
be an element of finite order (other than
) with characteristic polynomial
where
, and consider
as a module of over the commutative subring
.
As a ring we have
where
is a root of
, that is a primitive root of unity of order 3,4,6. It so happens that
is the ring of integers of the quadratic field
, and that these rings of integers are PIDs (in fact both these rings are Euclidean). By the classification of modules over a PID and since
is torsion-free we see that
is isomorphic to
as an
-module.
Now if
are two such elements then we get an isomorphism
intertwining the two
-module structures, in other words an element of
conjugating
.
12 February, 2021 at 6:35 pm
Terence Tao
Nice! So a little bit of algebraic number theory allows one to show the transitivity of the
action amongst all elements of finite order with fixed characteristic polynomial. (Unfortunately this argument is difficult to work into this complex analysis class as it requires a set of prerequisites that are basically disjoint from those already assumed for this course.)
12 February, 2021 at 7:07 pm
Lior Silberman
By the way you can also apply the same idea to the hyperbolic elements (those with
). The the ring
is an order in the corresponding real quadratic field
(a subring which is also a lattice), it is not hard to show
is one of the ideal classes of the order. We thus obtain the well-known bijection between closed geodesics on the modular surface and ideal classes in quadratic orders.
12 February, 2021 at 7:08 pm
Lior Silberman
(and I forgot to add to the bijection the conjugacy classes in the modular group of [primitive] hyperbolic elements)
3 June, 2023 at 12:07 pm
Jarek Kuben
It’s also possible to combine the lattice and algebraic number theory approaches in a way that perhaps gives more insight and shows that the two exceptional orbits are related to complex multiplication.
Indeed, let
be a fixed point of
, then
, thus
, where
. We claim that the map
given by
is a group isomorphism. Indeed, it’s easy to check that it’s a homomorphism, and it’s injective since
implies
,
, and then
, and thus
, and so
and
. It’s also surjective since for
the couple
is a
-basis of
with
, and so there exists
such that
, which implies
and
.
However for a lattice
it’s
and thus
, except for the case when
is a scalar multiple of an invertible ideal in some imaginary quadratic order
, in which case it’s
and
.
It follows that
has a stabilizer larger that
iff
is a scalar multiple of an invertible ideal in an imaginary quadratic order that has units other that
. However the only such orders are
and
, and they both have trivial class group, thus
is a scalar multiple of
or
, and so
belongs to the orbit of
or
.
Furthermore it’s
and
, and we can use the above map to explicitly compute these stabilizers. Indeed, it’s
and
, hence
and
.
5 February, 2021 at 5:43 am
Anonymous
Professor Tao, in the proof of Great Picard’s theorem, why is it that
lifts to a 1-periodic function?
[Oops, this is an issue. It can be resolved by analysing the possible shifts by elements of
that can occur, but this is a little tricky. For now I have deleted the proof of the theorem, will think more about whether there is a painless fix to it. UPDATE: now an exercise. -T]
5 February, 2021 at 11:01 am
Anonymous
In exercise 34 it seems that
should be 
[Corrected, thanks – T.]
5 February, 2021 at 12:56 pm
Anonymous
In the proof of theorem 39, it seems that there are some latex problems in some formulas.
[Fixed now – T.]
6 February, 2021 at 10:04 am
Anonymous
In remark 42 it seems (from the product representation of the generating function for
) that the summand
should also be included in the displayed sum.
[Corrected, thanks – T.]
7 February, 2021 at 10:14 am
Anonymous
In the first line of proposition 1, it seems that “periodic” should appear between “singly” and “holomorphic”.
[Corrected, thanks – T.]
9 February, 2021 at 4:01 am
Anonymous
In proposition 2, the periods should be linearly independent over
(not
)
[Corrected, thanks – T.]
9 February, 2021 at 4:19 am
Anonymous
In exercise 37, the statement of the great Picard theorem is missing.
[Added, thanks – T.]
13 February, 2021 at 2:10 pm
Anonymous
I don’t know if this is the best place to put it, but there is a typo remark 40 (descriminant)
16 February, 2021 at 7:41 am
Anonymous
Should Lemma 6 be codimension at most 1, i.e. 0 or 1?
[Corrected, thanks – T.]
16 February, 2021 at 8:11 pm
RJD
I admit, I’m a little puzzled in exercise 43. I am reading Serre’s “A course in Arithmetic” and looking at the Hecke operator section. He defines T_m(f) in the way I expect (i.e., using the notation above, T_m(f)(z)=G(z,1)). However, he derives that this is a modular form using the more explicit expression in part (ii). Is there some way of concluding that this is the unique modular form with lattice function G without using the expression directly?
18 February, 2021 at 3:42 pm
Terence Tao
Yes;
is clearly a function on
(so different generators give rise to the same value of
) and obeys the same homogeneity relation as
does (see Exercise 24), so modularity can be proven rather painlessly without needing the explicit expansion in (ii).
19 February, 2021 at 6:00 pm
RJD
Okay, so I have a dumb question. I agree that weak modularity follows fairly easily. What about holomorphicity at infinity? What am I missing here? It seems Serre does what he does to establish the holomorphicity at infinity, so I was wondering specifically about how that is taken care of.
22 February, 2021 at 6:18 pm
Terence Tao
There are only a bounded number of sublattices
of
of a fixed index (note that all lattices are isomorphic to each other as abelian groups, so once one has finiteness for say
one has finiteness for the rest), so if
is bounded at infinity then so is
. (Holomorphicity can be established by a similar argument.) Part (ii) comes from actually enumerating these bounded number of sublattices, but this isn’t necessary for the qualitative properties in (i).
24 February, 2021 at 11:44 pm
RJD
Thank you so much for your responses Professor Tao! One more question. You mention that F being bounded at infinity implies that G is bounded at infinity, and that holomorphicity can be established by a similar argument. What do you mean by bounded at infinity? Do you mean that the associated modular function given by Exercise 24 is bounded as $Im(\tau) \rightarrow \infty$?
[Yes – T.]
27 February, 2021 at 9:14 am
Course announcement: 246B, complex analysis | What's new
[…] Elliptic functions and their relatives; […]
27 February, 2021 at 9:37 am
246B, Notes 2: Some connections with the Fourier transform | What's new
[…] Previous set of notes: Notes 1. Next set of notes: Notes 3. […]
27 February, 2021 at 9:40 am
246B, Notes 4: The Riemann zeta function and the prime number theorem | What's new
[…] set of notes: Notes 3. Next set of notes: 246C Notes […]
12 February, 2022 at 12:13 pm
J.
Throughout the notes, you have used both
(as in the “additive unit circle”) and something like
. Are there any reasons for the preferences?
14 February, 2022 at 8:45 am
Terence Tao
For abelian actions the two notations are interchangeable. For non-abelian actions like
one has to use the left quotient if one wants the group
to act on the left, so I set up my notation for Riemann surface structures on quotient spaces using left quotients (and I have tried to emphasise the analogy between the torus
and the modular curve
in these notes). But I am not viewing the unit circle
as part of a larger class of possibly non-abelian quotient spaces, so here I am happy to use the traditional right quotient notation.
13 April, 2022 at 2:50 pm
J
If one has a group
and
being a normal subgroup of
, then one has the quotient group
.
In this set of notes, it seems to be the case that
is just a set and
being any group acting on
; and you talk about the “quotient”
or
. Is there a group structure hiding somewhere for the set
? Also, is there a “normal subgroup” lurking somewhere?
19 April, 2022 at 8:54 am
Terence Tao
One can quotient out any set by any equivalence relation to obtain a quotient space. In particular, any group acting on a space gives rise to a quotient space through the orbit equivalence relation. For instance, one can quotient any group
by any subgroup
to obtain a quotient space
(of left-cosets) or
(of right cosets). The normality of
is only required if one wants this quotient space to also be a group.
16 April, 2022 at 3:22 pm
J
In the proof of Lemma 3(iii), the path
is not closed. What modification of the argument principle is used?
19 April, 2022 at 9:18 am
Terence Tao
While the path
is not closed, the image
is. Apply a change of variables
to conclude.
16 April, 2022 at 4:19 pm
J
Consider the example of
and
where
. According to the notes,
consists of only constant functions and thus
by the definition.
But I got
because the order of zero is
at
. So
and thus
.
16 April, 2022 at 4:20 pm
J
What is going wrong in the calculation above?
[The function
is not elliptic (doubly periodic) and so does not give a well-defined holomorphic function on the torus. -T.]
17 April, 2022 at 6:37 am
J
… To figure out which, we can normalise
however this series turns out to not be absolutely convergent.
I am confused with the reasoning here.
The goal was to study the dimension of
, which is at least
since every constant function
(or equivalently, a doubly periodic function
) satisfies
. So question is now whether
– there is a meromorphic function
that has a double pole at
;
– or equivalently, a doubly periodic (meromorphic) function
that has a double pole at some point of
.
(Why should it be that “The question is now whether there is a doubly periodic meromorphic function that has a double pole at each point of
.”? Isn’t it that any single point in
is a representative of the coset?)
What is really the equivalent way to say “poles” for meromorphic function
and meromorphic function
?
19 April, 2022 at 9:22 am
Terence Tao
If a doubly periodic function
has a pole at any point
, it will also have a pole (of the same order) at every other point in the coset
by periodicity, and so when viewed as a function
, it can be said to have a pole at the point
(of the specified order).
17 April, 2022 at 9:26 am
J
Minor typo in (6):
is meant to be
[Corrected, thanks – T.]
17 April, 2022 at 12:08 pm
J
In proving the double periodicity of the function
, it is said that
… The series on the right is absolutely convergent, and on every coset of
Cosets of
is of the form
for some
. What do you mean by “coset of
“?
[A set of the form
for some complex number
. -T.]
18 April, 2022 at 4:46 pm
J
In (11):
–>
.
… we have
in a neighbourhood of infinity.
–>
… we have
in a neighbourhood of infinity.
[Corrected, thanks – T.]
20 April, 2022 at 12:56 pm
J
… so to demonstrate absolute convergence it suffices to show that
But a simple volume packing argument (considering the areas of the translates
Can you please elaborate on what the “volume packing argument” is?
22 April, 2022 at 8:56 am
Terence Tao
As
ranges over the lattice points in the disk
, the translates
of the fundamental domain each have area
and are disjoint (up to null sets) and occupy (a subset of) a disk of radius
, which has area
. So the total number of such lattice points is at most
.
(The same argument also gives a comparable lower bound.)
20 April, 2022 at 1:55 pm
J
If