An Introduction To Modular Forms
An Introduction To Modular Forms
Henri Cohen
arXiv:1809.10907v1 [math.NT] 28 Sep 2018
Abstract
In this course we introduce the main notions relative to the classical theory of mod-
ular forms. A complete treatise in a similar style can be found in the author’s book
joint with F. Strömberg [1].
1 Functional Equations
Henri Cohen
Institut de Mathématiques de Bordeaux, Université de Bordeaux, 351 Cours de la Libération,
33405 TALENCE Cedex, FRANCE, e-mail: [email protected]
1
2 Henri Cohen
absolutely convergent for all x ∈ R, where the Fourier coefficients a(n) are given by
the formula Z 1
a(n) = e−2π inx f (x) dx ,
0
which follows immediately from the orthonormality of the functions e2π imx (you
may of course replace the integral from 0 to 1 by an integral from z to z + 1 for any
z ∈ R).
An important consequence of this, easily proved, is the Poisson summation for-
mula: define the Fourier transform of f by
Z ∞
fb(x) = e−2π ixt f (t) dt .
−∞
We ignore all convergence questions, although of course they must be taken into
account in any computation.
Consider the function g(x) = ∑n∈Z f (x + n), which is exactly the averaging pro-
cedure mentioned above. Thus g(x + 1) = g(x), so g has a Fourier series, and an
easy computation shows the following (again omitting any convergence or regular-
ity assumptions):
∑ f (x + n) = ∑ fb(m)e2π imx .
n∈Z m∈Z
In particular
∑ f (n) = ∑ fb(m) .
n∈Z m∈Z
−aπ n2
T (a) = ∑e .
n∈Z
This is historically the first example of modularity, which we will see in more
detail below.
2
Exercise 1.4. Set S = ∑n≥1 e−(n/10) .
1. Compute
√ numerically S to 100 decimal digits, and show that it is apparently equal
to 5 π − 1/2. √
2. Show that in fact S is not exactly equal to 5 π − 1/2, and using the above corol-
lary give a precise estimate for the difference.
Exercise 1.5. 1. Show that the function f (x) = 1/ cosh(π x) is also invariant under
Fourier transform.
2. In a manner similar to the corollary, define
Note that if the function is only meromorphic, the region of convergence will be
limited by the closest pole. Consider for instance the function f (z) = 1/(e2π iz − 1) =
eπ iz /(2i sin(π z)). If we set y = ℑ(z) we have |e2π iz | = e−2π y , so if y > 0 we have the
Fourier expansion f (z) = − ∑n≥0 e2π inz , while if y < 0 we have the different Fourier
expansion f (z) = ∑n≤−1 e2π inz .
2 Elliptic Functions
The preceding section was devoted to periodic functions. We now assume that our
functions are defined on some subset of C and assume that they are doubly peri-
odic: this can be stated either by saying that there exist two R-linearly independent
complex numbers ω1 and ω2 such that f (z + ωi ) = f (z) for all z and i = 1, 2, or
equivalently by saying that there exists a lattice Λ in C (here Zω1 + Zω2 ) such that
for any λ ∈ Λ we have f (z + λ ) = f (z).
Note in passing that if ω1 /ω2 ∈ Q this is equivalent to (single) periodicity, and if
ω1 /ω2 ∈ R \ Q the set of periods would be dense so the only “doubly periodic” (at
least continuous) functions would essentially reduce to functions of one variable.
For a similar reason there do not exist nonconstant continuous functions which are
triply periodic.
In the case of simply periodic functions considered above there already existed
some natural functions such as e2π inx . In the doubly-periodic case no such function
exists (at least on an elementary level), so we have to construct them, and for this
we use the standard averaging procedure seen and used above. Here the group is the
lattice Λ , so we consider functions of the type f (z) = ∑ω ∈Λ φ (z + ω ). For this to
converge φ (z) must tend to 0 sufficiently fast as |z| tends to infinity, and since this is
a double sum (Λ is a two-dimensional lattice), it is easy to see by comparison with
an integral (assuming |φ (z)| is regularly decreasing) that |φ (z)| should decrease at
least like 1/|z|α for α > 2. Thus a first reasonable definition is to set
1 1
f (z) = ∑ = ∑
(z + ω )3 (m,n)∈Z (z + m ω 1 + nω2 )3
.
ω ∈Λ 2
However, this is not quite the basic elliptic function that we need. We can inte-
grate term by term, as long as we choose constants of integration such that the inte-
grated series continues to converge. To avoid stupid multiplicative constants, we in-
tegrate −2 f (z): all antiderivatives of −2/(z+ ω )3 are of the form 1/(z+ ω )2 +C(ω )
for some constant C(ω ), hence to preserve convergence we will choose C(0) = 0 and
C(ω ) = −1/ω 2 for ω 6= 0: indeed, |1/(z + ω )2 − 1/ω 2 | is asymptotic to 2|z|/|ω 3 |
as |ω | → ∞, so we are again in the domain of normal convergence. We will thus
define:
1 1 1
℘(z) = 2 + ∑ − ,
z ω ∈Λ \{0}
(z + ω )2 ω 2
D(ω ) depending on ω but not on z. Note a slightly subtle point here: we use the fact
that C \ Λ is connected. Do you see why?
Now as before it is clear that ℘(z) is an even function: thus, setting z = −ω /2 we
have ℘(ω /2) = ℘(−ω /2) + D(ω ) = ℘(ω /2) + D(ω ), so D(ω ) = 0 hence ℘(z +
ω ) = ℘(z) and ℘ is indeed an elliptic function. There is a mistake in this reasoning:
do you see it?
Since ℘ has poles on Λ , we cannot reason as we do when ω /2 ∈ Λ . Fortu-
nately, this does not matter: since ωi /2 ∈/ Λ for i = 1, 2, we have shown at least that
D(ωi ) = 0 hence that ℘(z + ωi ) = ℘(z) for i = 1, 2, so ℘ is doubly periodic (so
indeed D(ω ) = 0 for all ω ∈ Λ ).
The theory of elliptic functions is incredibly rich, and whole treatises have been
written about them. Since this course is mainly about modular forms, we will simply
summarize the main properties, and emphasize those that are relevant to us. All are
proved using manipulation of power series and complex analysis, and all the proofs
are quite straightforward. For instance:
Proposition 2.1. Let f be a nonzero elliptic function with period lattice Λ as above,
and denote by P = Pa a “fundamental parallelogram” Pa = {z = a + xω1 + yω2 , 0 ≤
x < 1, 0 ≤ y < 1}, where a is chosen so that the boundary of Pa does not contain
any zeros or poles of f (see Figure 1).
1. The number of zeros of f in P is equal to the number of poles (counted with
multiplicity), and this number is called the order of f .
2. The sum of the residues of f at the poles in P is equal to 0.
3. The sum of the zeros and poles of f in P belongs to Λ .
4. If f is nonconstant its order is at least 2.
Proof. For (1), (2), and (3), simply integrate f (z), f ′ (z)/ f (z), and z f ′ (z)/ f (z) along
the boundary of P and use the residue theorem. For (4), we first note that by (2) f
cannot have order 1 since it would have a simple pole with residue 0. But it also
cannot have order 0: this would mean that f has no pole, so it is an entire function,
and since it is doubly-periodic its values are those taken in P which is compact, so
6 Henri Cohen
f is bounded. By a famous theorem of Liouville (of which this is the no less most
famous application) it implies that f is constant, contradicting the assumption of
(4). ⊓
⊔
ω1 ω1 + a ω1 + ω2 + a
Ca
ω2
ω2 + a
a
Note that clearly ℘ has order 2, and the last result shows that we cannot find an
elliptic function of order 1. Note however the following:
Exercise 2.2. 1. By integrating term by term the series defining −℘(z) show that if
we define the Weierstrass zeta function
1 1 1 z
ζ (z) = + ∑ − + 2 ,
z ω ∈Λ \{0} z + ω ω ω
ω1 η2 − ω2 η1 = ±2π i ,
The main properties of ℘ that we want to mention are as follows: First, for z
sufficiently small and ω 6= 0 we can expand
1 1
= ∑ (−1)k (k + 1)zk k+2 ,
(z + ω )2 k≥0 ω
so
1
℘(z) = + ∑ (−1)k (k + 1)zk Gk+2 (Λ ) ,
z2 k≥1
where we have set
1
Gk (Λ ) = ∑ ω k
,
ω ∈Λ \{0}
1
℘(z) = + ∑ (2k + 1)z2k G2k+2 (Λ ) .
z2 k≥1
Second, one can show that all elliptic functions are simply rational functions in
℘(z) and ℘′ (z), so we need not look any further in our construction.
Third, and this is probably one of the most important properties of ℘(z), it satis-
fies a differential equation of order 1: the proof is as follows. Using the above Taylor
expansion of ℘(z), it is immediate to check that
2
F(z) = ℘′(z) − (4℘(z)3 − g2 (Λ )℘(z) − g3(Λ ))
2 are on the curve, hence if we draw the line through these two points (the tangent
to the curve if they are equal), it is immediate to see from Proposition 2.1 (3) that
the third point of intersection corresponds to the parameter −(z1 + z2 ), and can of
course be computed as a rational function of the coordinates of P1 and P2 . It follows
that ℘(z) (and ℘′ (z)) possesses an addition formula expressing ℘(z1 + z2 ) in terms
of the ℘(zi ) and ℘′ (zi ).
Exercise 2.3. Find this addition formula. You will have to distinguish the cases z1 =
z2 , z1 = −z2 , and z1 6= ±z2 .
An interesting corollary of the differential equation for ℘(z), which we will prove
in a different way below, is a recursion for the Eisenstein series G2k (Λ ):
Proposition 2.4. We have the recursion for k ≥ 4:
Proof. Taking the derivative of the differential equation and dividing by 2℘′ we
obtain ℘′′ (z) = 6℘(z)2 − g2 (Λ )/2. If we set by convention G0 (Λ ) = −1 and
G2 (Λ ) = 0, and for notational simplicity omit Λ which is fixed, we have ℘(z) =
∑k≥−1 (2k + 1)z2k G2k+2 , so on the one hand
G2k (Λ ) will tend to ∑n∈Z\{0} n−2k = 2ζ (2k), where ζ is the Riemann zeta function.
If follows that for all k ≥ 2, ζ (2k) is a polynomial in ζ (4) and ζ (6) with rational
coefficients. Of course this is a weak but nontrivial result, since we know that ζ (2k)
is a rational multiple of π 2k .
To finish this section on elliptic functions and make the transition to modular
forms, we write explicitly Λ = Λ (ω1 , ω2 ) and by abuse of notation G2k (ω1 , ω2 ) :=
G2k (Λ (ω1 , ω2 )), and we consider the dependence of G2k on ω1 and ω2 . We note two
evident facts: first, G2k (ω1 , ω2 ) is homogeneous of degree −2k: for any nonzero
complex number λ we have G2k (λ ω1 , λ ω2 ) = λ −2k G2k (ω1 , ω2 ). In particular,
G2k (ω1 , ω2 ) = ω2−2k G2k (ω1 /ω2 , 1). Second, a general Z-basis of Λ is given by
(ω1′ , ω2′ ) = (aω1 + bω2 , cω1 + d ω2 ) with a, b, c, d integers such that ad − bc = ±1.
If we choose an oriented basis such that ℑ(ω1 /ω2 ) > 0 we in fact have ad − bc = 1.
Thus, G2k (aω1 + bω2 , cω1 + d ω2 ) = G2k (ω1 , ω2 ), and using homogeneity this
can be written
−2k a ω1 + b ω2 −2k ω1
(cω1 + d ω2 ) G2k , 1 = ω2 G2k ,1 .
c ω1 + d ω2 ω2
3.1 Definitions
The reason for the factor (ad − bc)k/2 is that λ γ has the same action on H as γ , so
this makes the formula homogeneous. For instance, F is weakly modular of weight
k if and only if F|k γ = F for all γ ∈ Γ .
We will also use the universal modular form convention of writing q for e2π iτ , so
that a Fourier expansion is of the type F(τ ) = ∑n≥0 a(n)qn . We use the additional
convention that if α is any complex number, qα will mean e2π iτα .
Exercise 3.4. Let F(τ ) = ∑n≥0 a(n)qn ∈ Mk (Γ ), and let γ = CA DB be a matrix in
M2+ (Z), i.e., A, B, C, and D are integers and ∆ = det(γ ) = AD − BC > 0. Set g =
gcd(A,C), let u and v be such that uA + vC = g, set b = uB + vD, and finally let
ζ∆ = e2π i/∆ . Prove the matrix identity
AB A/g −v g b
= ,
CD C/g u 0 ∆ /g
gk/2 2
F|k γ (τ ) = k ∑ ζ∆nbg a(n)qg /∆ ,
∆ n≥0
The first fundamental result in the theory of modular forms is that these spaces
are finite-dimensional. The proof uses exactly the same method that we have used to
prove the basic results on elliptic functions. We first note that there is a “fundamental
domain” (which replaces the fundamental parallelogram) for the action of Γ on H ,
given by
F = {τ ∈ H , −1/2 ≤ ℜ(τ ) < 1/2, |τ | ≥ 1} .
The proof that this is a fundamental domain, in other words that any τ ∈ H has
a unique image by Γ belonging to F is not very difficult and will be omitted. We
then integrate F ′ (z)/F(z) along the boundary of F, and using modularity we obtain
the following result:
Theorem 3.5 Let F ∈ Mk (Γ ) be a nonzero modular form. For any τ0 ∈ H , denote
by vτ0 (F) the valuation of F at τ0 , i.e., the unique integer v such that F(τ )/(τ − τ0 )v
is holomorphic and nonzero at τ0 , and if F(τ ) = G(e2π iτ ), define vi∞ (F) = v0 (G)
(i.e., the number of first vanishing Fourier coefficients of F). We have the formula
vτ (F) k
vi∞ (F) + ∑ e τ
=
12
,
τ ∈F
− 12 1
2
This theorem has many important consequences but, as already noted, the most
important is that it implies that Mk (Γ ) is finite dimensional. First, it trivially implies
that k ≥ 0, i.e., there are no modular forms of negative weight. In addition it easily
implies the following:
multiple of G4 G6 , G12 is a linear combination of G34 and G26 . Also, we see that ∆ is
a linear combination of G34 and G26 (we will see this more precisely below).
A basic result on the structure of the modular group Γ is the following:
Proposition 3.8.
Set T = 10 11 , which acts on H by the unit translation τ 7→ τ + 1,
and S = 01 −10 which acts on H by the symmetry-inversion τ 7→ −1/τ . Then Γ is
generated by S and T , with relations generated by S2 = −I and (ST )3 = −I (I the
identity matrix).
There are several (easy) proofs of this fundamental result, which we do not give.
Simply note that this proposition is essentially equivalent to the fact that the set F
described above is indeed a fundamental domain.
A consequence of this proposition is that to check whether some function F
has the modularity property, it is sufficient to check that F(τ + 1) = F(τ ) and
F(−1/τ ) = τ k F(τ ).
Exercise 3.9. (Bol’s identity). Let F be any continuous function defined on the
upper-half plance H , and define I0 (F, a) = F and for any integer m ≥ 1 and a ∈ H
set: Z τ
(τ − z)m−1
Im (F, a)(τ ) = F(z) dz .
a (m − 1)!
1. Show that Im (F, a)′ (τ ) = Im−1 (F, a)(τ ), so that Im (F, a) is an mth antiderivative
of F.
2. Let γ ∈ Γ , and assume that k ≥ 1 is an integer. Show that
where D = (1/2π i)d/d τ = qd/dq is the basic differential operator that we will
use (see Section 3.10).
4. Assume now that F is weakly modular of weight k ≥ 1 and holomorphic on H
(in particular if F ∈ Mk (Γ ), but |F| could be unbounded as ℑ(τ ) → ∞). Show
that
(Fa∗ |2−k |γ )(τ ) = Fa∗ (τ ) + Pk−2 (τ ) ,
where Pk−2 is the polynomial of degree less than or equal to k − 2 given by
Z a
(X − z)k−2
Pk−2 (X) = F(z) dz .
γ −1 (a) (k − 2)!
What this exercise shows is that the (k − 1)st derivative of some function which
behaves modularly in weight 2 − k behaves modularly in weight k, and conversely
that the (k − 1)st antiderivative of some function which behaves modularly in weight
k behaves modularly in weight k up to addition of a polynomial of degree at most
14 Henri Cohen
π 3 τ 3
F4∗ (τ ) = − + ∑ σ−3 (n)qn .
180 i n≥1
ζ (3) π3 τ
τ 2 F4∗ (−1/τ ) = F4∗ (τ ) + (1 − τ 2) − .
2 36 i
2. Equivalently, if we set
π 3 τ 3 π 3 τ ζ (3)
F4∗∗ (τ ) = − − + + ∑ σ−3 (n)qn
180 i 72 i 2 n≥1
Note that the appearance of ζ (3) comes from the fact that, up to a multiplicative
constant, the L-function associated to G4 is equal to ζ (s)ζ (s − 3), whose value at
s = 3 is equal to −ζ (3)/2.
It follows in particular from this exercise that if F(τ )R is any integrable function
which is invariant by the modular group Γ , the integral Γ \H F(τ )d µ makes sense
if it converges. Since
R F is a fundamental domain for the action of Γ on H , this can
also be written F F(τ )d µ . Thus it follows from the second part that we can define
An Introduction to Modular Forms 15
Z
dxdy
< f , g >= f (τ )g(τ )yk ,
Γ \H y2
Proof. Recall that Gk (τ ) = ∑(m,n)∈Z2 \{(0,0)}(mτ + n)−k . We split the sum according
to the GCD of m and n: we let d = gcd(m, n), so that m = dm1 and n = dn1 with
gcd(m1 , n1 ) = 1. It follows that
Gk (τ ) = 2 ∑ d −k Ek (τ ) = 2ζ (k)Ek (τ ) ,
d≥1
where Ek (τ ) = (1/2) ∑gcd(m,n)=1 (mτ + n)−k . We thus need to prove that < Ek , f >=
0.
the other hand, denote by Γ∞ the group generated by T , i.e., translations
On
1 b for b ∈ Z. This acts by left multiplication on Γ , and it is immediate to check
01
that a system of representatives for this action is given by matrices ( mu nv ), where
gcd(m, n) = 1 and u and v are chosen arbitrarily (but only once for each pair (m, n))
such that un − vm = 1. It follows that we can write
Ek (τ ) = ∑ (mτ + n)−k ,
γ ∈Γ∞ \Γ
where it is understood that γ = ( mu nv ) (the factor 1/2 has disappeared since γ and −γ
have the same action on H ).
Thus
Z
dxdy
< Ek , f > = ∑ (mτ + n)−k f (τ )yk
Γ \H γ ∈Γ \Γ y2
∞
Z
dxdy
= ∑ (mτ + n)−k f (τ )yk .
γ ∈Γ∞ \Γ Γ \H
y2
Now note that by modularity f (τ ) = (mτ + n)−k f (γ (τ )), and since ℑ(γ (τ )) =
ℑ(τ )/|mτ + n|2 it follows that
which trivially vanishes since the inner integral is simply the conjugate of the con-
stant term in the Fourier expansion of f , which is 0 since f ∈ Sk (Γ ).
The Fourier expansions of the Eisenstein series G2k (τ ) are easy to compute. The
result is the following:
(2π i)k
Gk (τ ) = 2ζ (k) + 2 ∑ σk−1 (n)qn ,
(k − 1)! n≥1
Since we know that when k is even 2ζ (k) = −(2π i)k Bk /k!, where Bk is the k-th
Bernoulli number defined by
t Bk
= ∑ tk ,
et − 1 k≥0 k!
2k
Ek (τ ) = 1 − ∑ σk−1 (n)qn .
Bk n≥1
This is the normalization of Eisenstein series that we will use. For instance
An Introduction to Modular Forms 17
E4 (τ ) = 1 + 240 ∑ σ3 (n)qn ,
n≥1
E6 (τ ) = 1 − 504 ∑ σ5 (n)qn ,
n≥1
E8 (τ ) = 1 + 480 ∑ σ7 (n)qn .
n≥1
In particular, the relations given above which follow from the dimension formula
become much simpler and are obtained simply by looking at the first terms in the
Fourier expansion:
It is quite difficult (but not impossible) to prove this directly, i.e., without using at
least indirectly the theory of modular forms.
This type of reasoning is one of the reasons for which the theory of modular
forms is so important (and lots of fun!): if you have a modular form F, you can
usually express it in terms of a completely explicit basis of the space to which it be-
longs since spaces of modular forms are finite-dimensional (in the present example,
the space is one-dimensional), and deduce highly nontrivial relations for the Fourier
coefficients. We will see a further example of this below for the number rk (n) of
representations of an integer n as a sum of k squares.
nk qn
∑ σk (n)qn = ∑ 1 − qn ,
n≥1 n≥1
Note that in this exercise we only compute F(k) for k ≡ 1 (mod 4). It is also
possible but more difficult to compute F(k) for k ≡ 3 (mod 4). For instance we
have:
Γ (1/4)8 1
F(3) = − .
80(2π )6 240
We have mentioned at the beginning of this course that one of the ways to obtain
functions satisfying functional equations is to use averaging over a suitable group or
set: we have seen this for periodic functions in the form of the Poisson summation
formula, and for doubly-periodic functions in the construction of the Weierstrass
℘-function. We can do the same for modular forms, but we must be careful in two
different ways. First, we do not want invariance by Γ , but we want an automorphy
factor (cτ + d)k . This is easily dealt with by noting that (d/d τ )(γ (τ )) = (cτ + d)−2 :
indeed, if φ is some function on H we can define
′
′ A B
questions, by using the chain rule ( f ◦ g) =
Exercise 3.16. Ignoring all convergence
′
( f ◦ g)g show that for all δ = C D ∈ Γ we have
But the second important way in which we must be careful is that the above
contruction rarely converges. There are, however, examples where it does converge:
1
F(τ ) =
∑ (aτ + b)m (cτ + d)k−m
.
γ = a b ∈Γ
c d
Show that if 2 ≤ m ≤ k − 2 and m 6= k/2 this series converges normally on any com-
pact subset of H (i.e., it is majorized by a convergent series with positive terms),
so defines a modular form in Mk (Γ ).
Note that the series converges also for m = k/2, but this is more difficult.
One of the essential reasons for non-convergence of the function F is the trivial
observation that for a given pair of coprime integers (c, d) there are infinitely many
elements γ ∈ Γ having (c, d) as their second row. Thus in general it seems more
reasonable to define
An Introduction to Modular Forms 19
where γc,d is any fixed matrix in Γ with second row equal to (c, d). However, we
need this to make sense: if γc,d = ac db ∈ Γ is one such matrix, it is clear that the
general matrix having second row equal to (c, d) is T n ac db = a+nc b+nd , and as
c d
usual T = 10 11 is translation by 1: τ 7→ τ +1. Thus, an essential necessary condition
for our series to make any kind of sense is that the function φ be periodic of period 1.
The simplest such function is of course the constant function 1:
Exercise 3.18. (See the proof of Proposition 3.12.) Show that
1 e2π inγc,d (τ )
Pk (n; τ ) = ∑
2 gcd(c,d)=1 (cτ + d)k
,
where we note that we can choose any matrix γc,d with bottom row (c, d) since the
function e2π inτ is 1-periodic, so that Pk (n; τ ) ∈ Mk (Γ ).
Exercise 3.19. Assume that k ≥ 4 is even.
1. Show that if n < 0 the series defining Pk diverges (wildly in fact).
2. Note that Pk (0; τ ) = Ek (τ ), so that limτ →i∞ Pk (0; τ ) = 1. Show that if n > 0 the
series converges normally and that we have limτ →i∞ Pk (n; τ ) = 0. Thus in fact
Pk (n; τ ) ∈ Sk (Γ ) if n > 0.
3. By using the same unfolding method as in Proposition 3.12, show that if f =
∑n≥0 a(n)qn ∈ Mk (Γ ) and n > 0 we have
(k − 2)!
< Pk (n), f >= a(n) .
(4π n)k−1
It is easy to show that in fact the Pk (n) generate Sk (Γ ). We can also compute
their Fourier expansions as we have done for Ek , but they involve Bessel functions
and Kloosterman sums.
Recall that by definition ∆ is the generator of the 1-dimensional space S12 (Γ ) whose
Fourier coefficient of q1 is normalized to be equal to 1. By simple computation, we
20 Henri Cohen
with no apparent formula for the coefficients. The nth coefficient is denoted τ (n)
(no confusion with τ ∈ H ), and called Ramanujan’s tau function, and ∆ itself is
called Ramanujan’s Delta function.
Of course, using ∆ = (E43 − E62 )/1728 and expanding the powers, one can give
a complicated but explicit formula for τ (n) in terms of the functions σ3 and σ5 , but
this is far from being the best way to compute them. In fact, the following exercise
already gives a much better method.
3. Deduce in particular the congruences τ (n) ≡ nσ5 (n) ≡ nσ1 (n) (mod 5) and
τ (n) ≡ nσ3 (n) (mod 7).
Although there are much faster methods, this is already a very reasonable way to
compute τ (n).
The cusp form ∆ is one of the most important functions in the theory of modular
forms. Its first main property, which is not at all apparent from its definition, is that
it has a product expansion:
Theorem 3.21 We have
∆ (τ ) = q ∏ (1 − qn)24 .
n≥1
Proof. We are not going to give a complete proof, but sketch a method which is one
of the most natural to obtain the result.
We start backwards, from the product R(τ ) on the right-hand side. The logarithm
transforms products into sums, but in the case of functions f , the logarithmic deriva-
tive f ′ / f (more precisely D( f )/ f , where D = qd/dq) also does this, and it is also
more convenient. We have
nqn
D(R)/R = 1 − 24 ∑ n
= 1 − 24 ∑ σ1 (n)qn
n≥1 1 − q n≥1
as is easily seen by expanding 1/(1 − qn) as a geometric series. This is exactly the
case k = 2 of the Eisenstein series Ek , which we have excluded from our discussion
for convergence reasons, so we come back to our series G2k (we will divide by the
An Introduction to Modular Forms 21
Note the essential fact that there is now a nonanalytic term 3/(π ℑ(τ )). We will of
course set the following definition:
Proof of the theorem. We can now prove the theorem on the product expansion
of ∆ : noting that (d/d τ )γ (τ ) = 1/(cτ + d)2 , the above formulas imply that if we set
S = R(γ (τ )) we have
22 Henri Cohen
D(S) D(R)
= (γ (τ ))(d/d τ )(γ (τ ))
S R
12 c
= (cτ + d)−2E2 (γ (τ )) = E2 (τ ) +
2 π i cτ + d
D(R) D(cτ + d)
= (τ ) + 12 .
R cτ + d
By integrating and exponentiating, it follows that
Exercise 3.25. We have shown in passing that D(∆ ) = E2 ∆ . Expanding the Fourier
expansion of both sides, show that we have the recursion
m 1 1
∑ e 2π m − 1
=
24
−
8 π
.
m≥1
An Introduction to Modular Forms 23
Proof. (sketch): denote by L(q, u) the left-hand side. We have clearly L(q, u/q) =
−uL(q, u), and since one can write L(q, u) = ∑k∈Z ak (q)uk this implies the recursion
ak (q) = −qk ak−1 (q), so ak (q) = (−1)k qk(k+1)/2 a0 (q), and separating k ≥ 0 and k <
0 this shows that
The slightly longer part is to show that a0 (q) = 1: this is done by setting u = i/q1/2
and u = 1/q1/2, which after a little computation implies that a(q4 ) = a(q), and from
there it is immediate to deduce that a(q) is a constant, and equal to 1. ⊓
⊔
To give the next corollaries, we need to define the Dedekind eta function η (τ ),
by
η (τ ) = q1/24 ∏ (1 − qn) ,
n≥1
c (where we always use the principal determination of the square root), and since
we see from the infinite product that η (i) 6= 0, replacing τ by i shows that in fact
c = 1. Thus η satisfies the two basic modular equations
η (γ (τ )) = vη (γ )(cτ + d)1/2 η (τ )
giving the formula for η (τ ). For the second formula, divide the triple product iden-
tity by 1 − 1/u and make u → 1. ⊓
⊔
Exercise 3.29. 1. Show that 24∆ D(η ) = η D(∆ ), and using the explicit Fourier ex-
pansion of η , deduce the recursion
k 2 k(3k + 1)
∑ (−1) (75k + 25k + 2 − 2n)τ n −
2
=0.
k∈Z
Exercise 3.30. Define the q-Pochhammer symbol (q)n by (q)n = (1−q)(1−q2) · · · (1−
qn ).
1. Set f (a, q) = ∏n≥1 (1 − aqn ), and define coefficients cn (q) by setting f (a, q) =
∑n≥0 cn (q)an . Show that f (a, q) = (1 − aq) f (aq, q), deduce that cn (q)(1 − qn) =
−qn cn−1 (q) and finally the identity
An Introduction to Modular Forms 25
2. Write in terms of the Dedekind eta function the identities obtained by specializ-
ing to a = 1, a = −1, a = −1/q, a = q1/2, and a = −q1/2.
3. Similarly, prove the identity
and once again write in terms of the Dedekind eta function the identities obtained
by specializing to the same five values of a.
4. By multiplying two of the above identities and using the triple product identity,
prove the identity
2
1 qn
= ∑
∏n≥1 (1 − qn) n≥0 (q)2n
.
Note that this last series is the generating function of the partition function p(n),
so if one wants to make a table of p(n) up to n = 10000, say, using the left-hand
side would require 10000 terms, while using the right-hand side only requires 100.
Since its introduction, the Ramanujan tau function τ (n) has fascinated number the-
orists. For instance there is a conjecture due to D. H. Lehmer that τ (n) 6= 0, and an
even stronger conjecture (which would imply the former) that for every prime p we
have p ∤ τ (p) (on probabilistic grounds, the latter conjecture is probably false).
To test these conjectures as well as others, it is an interesting computational chal-
lenge to compute τ (n) for large n (because of Ramanujan’s first two conjectures, i.e.,
Mordell’s theorem that we will prove in Section 4 below, it is sufficient to compute
τ (p) for p prime).
We can have two distinct goals. The first is to compute a table of τ (n) for n ≤ B,
where B is some (large) bound. The second is to compute individual values of τ (n),
equivalently of τ (p) for p prime.
Consider first the construction of a table. The use of the first recursion given
in the above exercise needs O(n1/2 ) operations per value of τ (n), hence O(B3/2 )
operations in all to have a table for n ≤ B.
However, it is well known that the Fast Fourier Transform (FFT) allows one to
compute products of power series in essentially linear time. Thus, using Corollary
3.28, we can directly write the power series expansion of η 3 , and use the FFT to
compute its eighth power η 24 = ∆ . This will require O(B log(B)) operations, so is
much faster than the preceding method; it is essentially optimal since one needs
O(B) time simply to write the result.
26 Henri Cohen
See [1] Exercise 12.13 of Chapter 12 for details. Using this formula and a cluster, it
should be reasonable to compute τ (p) for p of the order of 1016 .
Exercise 3.32. 1. Noting that Theorem 3.5 is valid more generally for modular
functions (with vτ ( f ) = −r < 0 if f has a pole of order r at τ ) and using the
specific properties of j(τ ), compute vτ ( f ) for the functions j(τ ), j(τ ) − 1728,
and D( j)(τ ), at the points ρ = e2π i/3 , i, i∞, and τ0 for τ0 distinct from these three
special points.
2. Set f = f (a, b, c) = D( j)a /( jb ( j − 1728)c ). Show that f is a modular form if and
only if 2c ≤ a, 3b ≤ 2a, and b + c ≥ a, and give similar conditions for f to be a
cusp form.
3. Show that E4 = f (2, 1, 1), E6 = f (3, 2, 1), and ∆ = f (6, 4, 3), so that for instance
D( j) = −E14 = −E42 E6 /∆ .
But the importance of this theorem lies in algebraic number theory. We give the
following theorem without explaining the necessary notions:
Theorem 3.35 Let τ be a CM point, and D = b2 − 4ac its discriminant, where we
choose gcd(a, b, c) = 1. Then Q( j(τ )) is the ring class field of discriminant
√ D, and in
particular if D is the discriminant of a quadratic field K = Q( D), then K( j(τ )) is
28 Henri Cohen
the Hilbert class field of K. In particular, the degree of the minimal polynomial of the
algebraic integer j(τ ) is equal to the class number h(D) of the order of discriminant
D.
Examples:
√
j((1 + i 3)/2) = 0 = 1728 − 3(24)2
j(i) = 1728 = 123 = 1728 − 4(0)2
√
j((1 + i 7)/2) = −3375 = (−15)3 = 1728 − 7(27)2
√
j(i 2) = 8000 = 203 = 1728 + 8(28)2
√
j((1 + i 11)/2) = −32768 = (−32)3 = 1728 − 11(56)2
√
j((1 + i 163)/2) = −262537412640768000 = (−640320)3
= 1728 − 163(40133016)2
√
j(i 3) = 54000 = 2(30)3 = 1728 + 12(66)2
j(2i) = 287496 = (66)3 = 1728 + 8(189)2
√
j((1 + 3i 3)/2) = −12288000 = −3(160)3 = 1728 − 3(2024)2
√
√ −191025 − 85995 5
j((1 + i 15)/2) =
2
√ √ !3 √ !2
1 − 5 75 + 27 5 273 + 105 5
= = 1728 − 3
2 2 2
Note that we give the results in the above form since it can be shown that the
functions j1/3 and ( j − 1728)1/2 also have interesting arithmetic properties.
The example with D = −163 is particularly spectacular:
Exercise 3.37. 1. Using once again the example of 163, compute heuristically a
few terms of the Fourier expansion of j assuming that it is of the form 1/q +
∑n≥0 c(n)q√n with c(n) reasonably small integers using the following method. Set
√
q = −e−π 163 , and let J = (−640320)3 be the exact value of j((−1 + i 163)/2).
By computing J −1/q, one notices that the result is very close to 744, so we guess
that c(0) = 744. We then compute (J − 1/q − c(0))/q and note that once again
the result is close to an integer, giving c(1), and so on. Go as far as you can with
this method.
An Introduction to Modular Forms 29
2. Do the same for 67 instead of 163. You will find the same Fourier coefficients
(but you can go less far).
3. On the
√
other hand, do the same for 58, starting with J equal to the integer close
to eπ 58 . You will find a different Fourier expansion: it corresponds in fact to
another modular function, this time defined on a subgroup of Γ , called a Haupt-
modul. √
4. Try to find other rational numbers D such that eπ D is close to an integer, and do
the same exercise for them (an example where D is not integral is 89/3).
If we differentiate the modular equation f ((aτ + b)/(cτ + d)) = (cτ + d)k f (τ ) with
a b ∈ Γ using the operator D = (1/(2π i))d/d τ (which gives simpler formulas
c d
than d/d τ since D(qn ) = nqn), we easily obtain
aτ + b k c
D( f ) = (cτ + d)k+2 D( f )(τ ) + f (τ ) .
cτ + d 2 π i cτ + d
Thus the derivative of a weakly modular form of weight k looks like one of weight
k + 2, except that there is an extra term. This term vanishes if k = 0, so the derivative
of a modular function of weight 0 is indeed modular of weight 2 (we have seen above
the example of j(τ ) which satisfies D( j) = −E14 /∆ ).
If k > 0 and we really want a true weakly modular form of weight k + 2 there are
two ways to do this. The first one is called the Serre derivative:
Exercise 3.38. Using Proposition 3.23, show that if f is weakly modular of weight
k then D( f ) − (k/12)E2 f is weakly modular of weight k + 2. In particular, if f ∈
Mk (Γ ) then SDk ( f ) := D( f ) − (k/12)E2 f ∈ Mk+2 (Γ ).
k2 f2 D( f1 ) − k1 f1 D( f2 ) ∈ Sk1 +k2 +2 (Γ ) .
Lemma 3.40. Let a and b be nonnegative integers such that 4a + 6b = 12r + 2. The
constant term of the Fourier expansion of Fr (a, b) = E4a E6b /∆ r vanishes.
proving that the result is true for r by induction on b since we assumed it true for
r − 1. ⊓
⊔
E12r−k+2
= ∑ cki qi ,
∆r i≥−r
In addition we have ck0 6= 0, so that a(0) = ∑1≤n≤r (ck−n /ck0 )a(n) is a linear combi-
nation with rational coefficients of the a(n) for 1 ≤ n ≤ r.
Proof. First note that by Corollary 3.6 we have r ≥ (k − 2)/12 (with equality only
if k ≡ 2 (mod 12)), so the definition of the coefficients cki makes sense. Note also
that since the Fourier expansion of E12r−k+2 begins with 1 + O(q) and that of
∆ r by qr + O(qr+1 ), that of the quotient begins with q−r + O(q1−r ) (in particular
ck−r = 1). The proof of the first part is now immediate: the modular form f E12r−k+2
belongs to M12r+2 (Γ ), so by Corollary 3.7 is a linear combination of E4a E6b with
4a+6b = 12r +2. It follows from the lemma that the constant term of f E12r−k+2 /∆ r
vanishes, and this constant term is equal to ∑0≤n≤r ck−n a(n), proving the first part of
the theorem. The fact that ck0 6= 0 (which is of course essential) is a little more diffi-
cult and will be omitted, see [1] Theorem 9.5.1. ⊓
⊔
An Introduction to Modular Forms 31
This theorem has (at least) two consequences. First, a theoretical one: if one
can construct a modular form whose constant term is some interesting quantity and
whose Fourier coefficients a(n) are rational, this shows that the interesting quantity
is also rational. This is what allowed Siegel to show that the value at negative inte-
gers of Dedekind zeta functions of totally real number fields are rational, see Section
7.2. Second, a practical one: it allows to compute explicitly the constant coefficient
a(0) in terms of the a(n), giving interesting formulas, see again Section 7.2.
We now come to one of the most amazing and important discoveries on modular
forms due to S. Ramanujan, which has led to the modern development of the subject.
Recall that we set
∆ (τ ) = q ∏ (1 − qm )24 = ∑ τ (n)qn .
m≥1 n≥1
We have τ (2) = −24, τ (3) = 252, and τ (6) = −6048 = −24 · 252, so that τ (6) =
τ (2)τ (3). After some more experiments, Ramanujan conjectured that if m and n are
coprime we have τ (mn) = τ (m)τ (n). Thus, by decomposing an integer into products
of prime powers, assuming this conjecture, we are reduced to the study of τ (pk ) for
p prime.
Ramanujan then noticed that τ (4) = −1472 = (−24)2 − 211 = τ (2)2 − 211 , and
again after some experiments he conjectured that τ (p2 ) = τ (p)2 − p11 , and more
generally that τ (pk+1 ) = τ (p)τ (pk ) − p11 τ (pk−1 ). Thus uk = τ (pk ) satisfies a linear
recurrence relation
uk+1 − τ (p)uk + p11 uk−1 = 0 ,
and since u0 = 1 the sequence is entirely determined by the value of u1 = τ (p).
It is well-known that the behavior of a linear recurrent sequence is determined by
its characteristic polynomial. Here it is equal to X 2 − τ (p)X + p11 , and the third of
Ramanujan’s conjectures is that the discriminant of this equation is always negative,
or equivalently that |τ (p)| < p11/2 .
Note that if α p and β p are the roots of the characteristic polynomial (necessarily
distinct since we cannot have |τ (p)| = p11/2 ), then τ (pk ) = (α pk+1 − β pk+1 )/(α p −
β p ), and the last conjecture says that α p and β p are complex conjugate, and in par-
ticular of modulus equal to p11/2 .
These conjectures are all true. The first two (multiplicativity and recursion) were
proved by L. Mordell only one year after Ramanujan formulated them, and indeed
the proof is quite easy (in fact we will prove them below). The third conjecture
|τ (p)| < p11/2 is extremely hard, and was only proved by P. Deligne in 1970 using
the whole machinery developed by the school of A. Grothendieck to solve the Weil
conjectures .
32 Henri Cohen
The main idea of Mordell, which was generalized later by E. Hecke, is to intro-
duce certain linear operators (now called Hecke operators) on spaces of modular
forms, to prove that they satisfy the multiplicativity and recursion properties (this is
in general much easier than to prove this on numbers), and finally to use the fact that
S12 (Γ ) = C∆ is of dimension 1, so that necessarily ∆ is an eigenform of the Hecke
operators whose eigenvalues are exactly its Fourier coefficients.
Although there are more natural ways of introducing them, we will define
the Hecke operator T (n) on Mk (Γ ) directly by its action on Fourier expansions
T (n)(∑m≥0 a(m)qm ) = ∑m≥0 b(m)qm , where
Note that we can consider this definition as purely formal, apart from the presence of
the integer k this is totally unrelated to the possible fact that ∑m≥0 a(m)qm ∈ Mk (Γ ).
A simple but slightly tedious combinatorial argument shows that these operators
satisfy
T (n)T (m) = ∑ d k−1 T (nm/d 2 ) .
d|gcd(n,m)
The inner sum is a complete geometric sum which vanishes unless p | m, in which
case it is equal to p. Thus, changing m into pm we have G(τ ) = p ∑m≥0 a(pm)qm .
On the other hand, we have trivially ∑ p|m a(m/p)qm = ∑m≥0 a(m)q pm = F(pτ ).
Replacing both of these formulas in the formula for T (p)(F) we see that
1 τ+ j
T (p)(F)(τ ) = pk−1 F(pτ ) +
p 0≤∑
F .
j<p p
which implies (and is equivalent to) the first two conjectures of Ramanujan.
Denote by Pk (n) the characteristic polynomial of the linear map T (n) on Sk (Γ ).
A strong form of the so-called Maeda’s conjecture states that for n > 1 the polyno-
mial Pk (n) is irreducible. This has been tested up to very large weights.
Exercise 4.2. The above proof shows that the Hecke operators also preserve the
space of modular functions, so by Theorem 3.31 the image of j(τ ) will be a rational
function in j:
1. Show for instance that
2. Set J = j − 744, i.e., j with no term in q0 in its Fourier expansion. Deduce that
and observe that the coefficients that we obtain are exactly the Fourier coeffi-
cients of J.
3. Prove that T (n)( j) is a polynomial in j. Does the last observation generalize?
34 Henri Cohen
The case of ∆ is quite special, in that the modular form space to which it naturally
belongs, S12 (Γ ), is only 1-dimensional. As can easily be seen from the dimension
formula, this occurs (for cusp forms) only for k = 12, 16, 18, 20, 22, and 26 (there
are no nonzero cusp forms in weight 14 and the space is of dimension 2 in weight
24), and thus the evident cusp forms ∆ Ek−12 for these values of k (setting E0 = 1)
are generators of the space Sk (Γ ), so are eigenforms of the Hecke operators and
share exactly the same properties as ∆ , with p11 replaced by pk−1 .
When the dimension is greater than 1, we must work slightly more. From the for-
mulas given above it is clear that the T (n) form a commutative algebra of operators
on the finite dimensional vector space Sk (Γ ). In addition, we have seen above that
there is a natural scalar product on Sk (Γ ). One can show the not completely trivial
fact that T (n) is Hermitian for this scalar product, hence in particular is diagonaliz-
able. It follows by an easy and classical result of linear algebra that these operators
are simultaneously diagonalizable, i.e., there exists a basis Fi of forms in Sk (Γ ) such
that T (n)Fi = λi (n)Fi for all n and i. Identifying Fourier coefficients as we have done
above for ∆ shows that if Fi = ∑n≥1 ai (n)qn we have ai (n) = λi (n)ai (0). This im-
plies first that ai (0) 6= 0, otherwise Fi would be identically zero, so that by dividing
by ai (0) we can always normalize the eigenforms so that ai (0) = 1, and second, as
for ∆ , that ai (n) = λi (n), i.e., the eigenvalues are exactly the Fourier coefficients. In
addition, since the T (n) are Hermitian, these eigenvalues are real for any embedding
into C, hence are totally real, in other words their minimal polynomial has only real
roots. Finally, using Theorem 3.5, it is immediate to show that the field generated
by the ai (n) is finite-dimensional over Q, i.e., is a number field.
Exercise 5.1. Consider the space S = S24 (Γ ), which is the smallest weight where
the dimension is greater than 1, here 2. By the structure theorem given above, it is
generated for instance by ∆ 2 and ∆ E43 . Compute the matrix of the operator T (2)
on this basis of S, diagonalize this matrix, so find the eigenfunctions of T (2) on S
(the prime number 144169 should occur). Check that these eigenfunctions are also
eigenfunctions of T (3).
Thus, let F = ∑n≥1 a(n)qn be a normalized eigenfunction for all the Hecke oper-
ators in Sk (Γ ) (for instance F = ∆ with k = 12), and consider the Dirichlet series
a(n)
L(F, s) = ∑ s
,
n≥1 n
for the moment formally, although we will show below that it converges for
ℜ(s) sufficiently large. The multiplicativity property of the coefficients (a(nm) =
a(n)a(m) if gcd(n, m) = 1, coming from that of the T (n)) is equivalent to the fact
that we have an Euler product (a product over primes)
An Introduction to Modular Forms 35
a(p j )
L(F, s) = ∏ L p (F, s) with L p (F, s) = ∑ js
,
p∈P j≥0 p
(multiply both sides by the denominator to check this). We have thus proved the
following theorem:
Theorem 5.2 Let F = ∑n≥1 a(n)qn ∈ Sk (Γ ) be an eigenfunction of all Hecke oper-
ators. We have an Euler product
a(n) 1
L(F, s) = ∑ n s
=∏
1 − a(p)p −s + pk−1 p−2s
.
n≥1 p∈P
Note that we have not really used the fact that F is a cusp form: the above theorem
is still valid if F = Fk is the normalized Eisenstein series
Bk Bk
Fk (τ ) = − Ek (τ ) = − + ∑ σk−1 (n)qn ,
2k 2k n≥1
with σa (p) = pa + 1.
2. Show that mn
σa (m)σa (n) = ∑ d a σa ,
d|gcd(m,n)
d2
Everything that we have done up to now is purely formal, i.e., we do not need to
assume convergence. However in the sequel we will need to prove some analytic
36 Henri Cohen
results, and for this we need to prove convergence for certain values of s. We begin
with the following easy bound, due to Hecke:
Proof. The trick is to consider the function g(τ ) = |F(τ )ℑ(τ )k/2 |: since we have
seen that ℑ(γ (τ )) = ℑ(τ )/|cτ + d|2 , it follows that g(τ ) is invariant under Γ . It
follows that supτ ∈H g(τ ) = supτ ∈F g(τ ), where F is the fundamental domain used
above. Now because of the Fourier expansion and the fact that F is a cusp form,
|F(τ )| = O(e−2π ℑ(τ ) ) as ℑ(τ ) → ∞, so g(τ ) tends to 0 also. It immediately follows
that g is bounded on F, hence on H , so that there exists a constant c1 > 0 such that
|F(τ )| ≤ c1 ℑ(τ )−k/2 for all τ .
We can now easily prove Hecke’s bound: from the Fourier series section we know
that for any y > 0
Z 1
a(n) = e2π ny F(x + iy)e−2π inx dx ,
0
so that |a(n)| ≤ c1 e2π ny y−k/2 , and choosing y = 1/n proves the proposition with
c = e2 π c1 . ⊓
⊔
Corollary 5.5. The L-function of a cusp form of weight k converges absolutely (and
uniformly on compact subsets) for ℜ(s) > k/2 + 1.
Remark 5.6. Deligne’s deep result mentioned above on the third Ramanujan conjec-
ture implies that we have the following optimal bound: there exists c > 0 such that
|a(n)| ≤ cσ0 (n)n(k−1)/2 , and in particular |a(n)| = O(n(k−1)/2+ε ) for all ε > 0. This
implies that the L-function of a cusp form converges absolutely and uniformly on
compact subsets in fact also for ℜ(s) > (k + 1)/2.
Exercise 5.7. . Define for all s ∈ C the function σs (n) by σs (n) = ∑d|n d s if n ∈ Z>0 ,
σs (0) = ζ (−s)/2 (and σs (n) = 0 otherwise). Set
1. Compute S(s1 , s2 ; n) exactly in terms of σs1 +s2 +1 (n) for (s1 , s2 ) = (3, 3) and
(3, 5), and also for (s1 , s2 ) = (1, 1), (1, 3), (1, 5), and (1, 7) by using properties
of the function E2 .
2. Using Hecke’s bound for cusp forms, show that if s1 and s2 are odd positive
integers the ratio S(s1 , s2 ; n)/σs1 +s2 +1 (n) tends to a limit L(s1 , s2 ) as n → ∞, and
compute this limit in terms of Bernoulli numbers. In addition, give an estimate
for the error term |S(s1 , s2 ; n)/σs1 +s2 +1 (n) − L(s1 , s2 )|.
3. Using the values of the Riemann zeta function at even positive integers in terms
of Bernoulli numbers, show that if s1 and s2 are odd positive integers we have
An Introduction to Modular Forms 37
Try to prove it for s1 = s2 = 2, and then for general s1 , s2 . If you succeed, give
also an estimate for the error term analogous to the one obtained above.
Corollary 5.9. The function L(F, s) is a holomorphic function which can be analyti-
cally continued to the whole of C. In addition, if we set Λ (F, s) = (2π )−sΓ (s)L(F, s)
we have the functional equation Λ (F, k − s) = i−k Λ (F, s).
Note that in our case k is even, so that i−k = (−1)k/2 , but we prefer writing the
constant as above so as to be able to use a similar result in odd weight, which occur
in more general situations.
Proof. Indeed, splitting the integral at 1, changing t into 1/t in one of the integrals,
and using modularity shows immediately that
Z ∞
(2π )−sΓ (s)L(F, s) = F(it)(t s−1 + ikt k−1−s ) dt .
1
Since the integral converges absolutely and uniformly for all s (recall that F(it)
tends exponentially fast to 0 when t → ∞), this immediately implies the corollary.
⊓
⊔
As an aside, note that the integral formula used in the above proof is a very
efficient numerical method to compute L(F, s), since the series obtained on the right
by term by term integration is exponentially convergent. For instance:
Exercise 5.10. Let F(τ ) = ∑n≥1 a(n)qn be the Fourier expansion of a cusp form of
weight k on Γ . Using the above formula, show that the value of L(F, k/2) at the
center of the “critical strip” 0 ≤ ℜ(s) ≤ k is given by the following exponentially
convergent series
38 Henri Cohen
a(n) −2π n
L(F, k/2) = (1 + (−1)k/2) ∑ k/2
e Pk/2 (2π n) ,
n≥1 n
Note in particular that if k ≡ 2 (mod 4) we have L(F, k/2) = 0. Prove this directly.
Exercise 5.11. 1. Prove that if F is not necessarily a cusp form we have |a(n)| ≤
cnk−1 for some c > 0.
2. Generalize the proposition and the integral formulas so that they are also valid
form non-cusp forms; you will have to add polar parts of the type 1/s and 1/(s −
k).
3. Show that L(F, s) still extends to the whole of C with functional equation, but
that it has a pole, simple, at s = k, and compute its residue. In passing, show that
L(F, 0) = −a(0).
Λ (F, j)/ω(−1) j ∈ K ,
Exercise 5.13. (see also Exercise 3.9). For F ∈ Sk (Γ ) define the period polynomial
P(F, X) by
Z i∞
P(F; X) = (X − τ )k−2 F(τ ) d τ .
0
1. For γ ∈ Γ show that
Z γ −1 (i∞)
P(F; X)|2−k = (X − τ )k−2 F(τ ) d τ .
γ −1 (0)
3. Show that
k−2
k−2
P(F; X) = − ∑ (−i)k−1− j Λ (F, k − 1 − j)X j .
j=0 j
If we replace the expression (cτ + d)k by |cτ + d|2s for some complex number s, we
can also obtain functions which are invariant by Γ , although they are nonanalytic.
More precisely:
ys
G(s)(τ ) = ∑ |cτ + d|2s
and
(c,d)∈Z2 \{(0,0)}
1 ys
E(s)(τ ) = ∑ ℑ(γ (τ ))s = ∑ .
γ ∈Γ∞ \Γ
2 gcd(c,d)=1 |cτ + d|2s
This is again an averaging procedure, and it follows that G(s) and E(s) are in-
variant under Γ . In addition, as in the case of the holomorphic Eisenstein series Gk
and Ek , it is clear that G(s) = ζ (2s)E(s). One can also easily compute their Fourier
expansion, and the result is as follows:
Proposition 5.15. Set Λ (s) = π −s/2Γ (s/2)ζ (s). We have the Fourier expansion
40 Henri Cohen
σ2s−1 (n)
Λ (2s)E(s) = Λ (2s)ys + Λ (2−2s)y1−s +4y1/2 ∑ s−1/2
Ks−1/2 (2π ny) cos(2π nx) .
n≥1 n
In the above, Kν (x) is a K-Bessel function which we do not define here. The
main properties that we need is that it tends to 0 exponentially (more precisely
Kν (x) ∼ (π /(2x))1/2e−x as x → ∞) and that K−ν = Kν . It follows from the above
Fourier expansion that E(s) has an analytic continuation to the whole complex
plane, that it satisfies the functional equation E (1 − s) = E (s), where we set
E (s) = Λ (2s)E(s), and that E(s) has a unique pole, at s = 1, which is simple with
residue 3/π , independent of τ .
Exercise 5.16. Using the properties of the Riemann zeta function ζ (s), show this
last property, i.e., that E(s) has a unique pole, at s = 1, which is simple with residue
3/π , independent of τ .
There are many reasons for introducing these nonholomorphic Eisenstein series,
but for us the main reason is that they are fundamental in unfolding methods. Recall
that using unfolding, in Proposition 3.12 we showed that Ek (or Gk ) was orthogonal
to any cusp form. In the present case, we obtain a different kind of result called a
Rankin–Selberg convolution. Let f and g be in Mk (Γ ), one of them being a cusp
form. Since E(s) is invariant by Γ the scalar product < E(s) f , g > makes sense, and
the following proposition gives its value:
Proposition 5.17. Let f (τ ) = ∑n≥0 a(n)qn and g(τ ) = ∑n≥0 b(n)qn be in Mk (Γ ),
with at least one being a cusp form. For ℜ(s) > 1 we have
Γ (s + k − 1) a(n)b(n)
< E(s) f , g >=
(4π )s+k−1 ∑ s+k−1
.
n≥1 n
Proof. We essentially copy the proof of Proposition 3.12 so we skip the details:
setting temporarily F(τ ) = f (τ )g(τ )yk which is invariant by Γ , we have
Z
< E(s) f , g > = ∑ ℑ(γ (τ ))s F(γ (τ )) d µ
Γ \H γ ∈Γ \Γ
∞
= ∑ ℑ(τ )s F(τ ) d µ
Γ∞ \H
Z ∞ Z 1
s+k−2
= y F(x + iy) dx dy .
0 0
The inner integral is equal to the constant term in the Fourier expansion of F,
hence is equal to ∑n≥1 a(n)b(n)e−4π ny (note that by assumption one of f and g is a
cusp form, so the term n = 0 vanishes), and the proposition follows. ⊓
⊔
Corollary 5.18. For ℜ(s) > k set
a(n)b(n)
R( f , g)(s) = ∑ ns
.
n≥1
An Introduction to Modular Forms 41
1. R( f , g)(s) has an analytic continuation to the whole complex plane and satisfies
the functional equation R(2k − 1 − s) = R(s) with
3 (4π )k
< f,g > .
π (k − 1)!
(k + ℓ − 2)! a(n)b(n)
< Ek f , g >= ∑ nk+ℓ−1 .
(4π )k+ℓ−1 n≥1
We have used as basic definition of (weak) modularity F|k γ = F for all γ ∈ Γ . But
there is no reason to restrict to Γ : we could very well ask the same modularity
condition for some group G of transformations of H different from Γ .
There are many types of such groups, and they have been classified: for us, we
will simply distinguish three types, with no justification. For any such group G we
42 Henri Cohen
can talk about a fundamental domain, similar to F that we have drawn above (I do
not want to give a rigorous definition here). We can distinguish essentially three
types of such domains, corresponding to three types of groups.
The first type is when the domain (more precisely its closure) is compact: we say
in that case that G is cocompact. It is equivalent to saying that it does not have any
“cusp” such as i∞ in the case of G. These groups are very important, but we will not
consider them here.
The second type is when the domain is not compact (i.e., it has cusps), but it has
finite volume for the measure d µ = dxdy/y2 on H defined in Exercise 3.11. Such a
group is said to have finite covolume, and the main example is G = Γ that we have
just considered, hence also evidently all the subgroups of Γ of finite index.
Exercise 6.1. Show that the covolume of the modular group Γ is finite and equal to
π /3.
The third type is when the volume is infinite: a typical example is the group
Γ∞ generated by integer translations, i.e., the set of matrices 10 1n . A fundamental
domain is then any vertical strip in H of width 1, which can trivially be shown
to have infinite volume. These groups are not important (at least for us) for the
following reason: they would have “too many” modular forms. For instance, in the
case of Γ∞ a “modular form” would simply be a holomorphic periodic function of
period 1, and we come back to the theory of Fourier series, much less interesting.
We will therefore restrict to groups of the second type, which are called Fuchsian
groups of the first kind. In fact, for this course we will even restrict to subgroups G
of Γ of finite index.
However, even with this restriction, it is still necessary to distinguish two types
of subgroups: the so-called congruence subgroups, and the others, of course called
non-congruence subgroups. The theory of modular forms on non-congruence sub-
groups is quite a difficult subject and active research is being done on them. One
annoying aspect is that they apparently do not have a theory of Hecke operators.
Thus will will restrict even more to congruence subgroups. We give the following
definitions:
Definition 6.2. Let N ≥ 1 be an integer.
1. We define
a b 10
Γ (N) = {γ = ∈Γ, γ ≡ (mod N)} ,
c d 01
a b 1∗
Γ1 (N) = {γ = ∈Γ, γ ≡ (mod N)} ,
c d 01
a b ∗∗
Γ0 (N) = {γ = ∈Γ, γ ≡ (mod N)} ,
c d 0∗
It is clear that Γ (N) ⊂ Γ1 (N) ⊂ Γ0 (N), and it is trivial to prove that Γ (N)
is normal in Γ (hence in any subgroup of Γ containing it), that Γ1 (N)/Γ (N) ≃
Z/NZ (with the map ac db 7→ b mod N), and that Γ1 (N) is normal in Γ0 (N) with
Γ0 (N)/Γ1 (N) ≃ (Z/NZ)∗ (with the map ac db 7→ d mod N).
If G is a congruence subgroup of level N we have Γ (N) ⊂ G, so (whatever the
definition) a modular form on G will in particular be on Γ (N). Because of the above
isomorphisms, it is not difficult to reduce the study of forms on Γ (N) to those on
Γ1 (N), and the latter to forms on Γ0 (N), except that we have to add a slight “twist”
to the modularity property. Thus for simplicity, we will restrict to modular forms on
Γ0 (N).
In view of the definition given for Γ , it is natural to say that F is weakly modular
of weight kon Γ0 (N) if for all γ ∈ Γ0 (N) we have F|k γ = F, where we recall that
if γ = ac db then F|k γ (τ ) = (cτ + d)−k F(τ ). To obtain a modular form, we need
also to require that F is holomorphic on H , plus some additional technical condi-
tion “at infinity”. In the case of the full modular group Γ , this condition was that
F(τ ) remains bounded as ℑ(τ ) → ∞. In the case of a subgroup, this condition is
not sufficient (it is easy to show that if we do not require an additional condition
the corresponding space will in general be infinite-dimensional). There are several
equivalent ways of giving the additional condition. One is the following: writing as
usual τ = x + iy, we require that there exists N such that in the strip −1/2 ≤ x ≤ 1/2,
we have |F(τ )| ≤ yN as y → ∞ and |F(τ )| ≤ y−N as y → 0 (since F is 1-periodic,
there is no loss of generality in restricting to the strip).
It is easily shown that if F is weakly modular and holomorphic, then the above
inequalities imply that |F(τ )| is in fact bounded as y → ∞ (but in general not as
y → 0), so the first condition is exactly the one that we gave in the case of the full
modular group.
Similarly, we can define a cusp form by asking that in the above strip |F(τ )| tends
to 0 as y → ∞ and as y → 0.
Now that we have a solid definition of modular form, we can try to proceed as
in the case of the full modular group. A number of things can easily be generalized.
It is always convenient to choose a system of representatives (γ j ) of right cosets for
Γ0 (N) in Γ , so that G
Γ = Γ0 (N)γ j .
j
44 Henri Cohen
For instance, if F is the fundamental domain of Γ seen above, one can choose D =
F
γ j (F) as fundamental domain for Γ0 (N). The theorem that we gave on valuations
generalizes immediately:
vτ (F) k
∑ eτ
= [Γ : Γ0 (N)] ,
12
τ ∈D
where D is D to which is added a finite number of “cusps” (we do not explain this;
it is not the topological closure), eτ = 2 (resp., 3) if τ is Γ -equivalent to i (resp., to
ρ ), and eτ = 1 otherwise, and we can then deduce the dimension of Mk (Γ0 (N)) and
Sk (Γ0 (N)) as we did for Γ :
Theorem 6.4 We have M0 (Γ0 (N)) = C (i.e., the only modular forms of weight 0 are
the constants) and S0 (Γ0 (N)) = {0}. For k ≥ 2 even, we have
where δk,2 is the Kronecker symbol (1 if k = 2, 0 otherwise) and the Ai are given as
follows:
k−1 1
12 ∏
A1 = N 1+ ,
p|N
p
k−1 k −3
A2,3 =
3
−
3 ∏ 1+ p if 9 ∤ N, 0 otherwise,
p|N
k−1 k −4
A2,4 =
4
−
4 ∏ 1+ p if 4 ∤ N, 0 otherwise,
p|N
1
A3 = ∑ φ (gcd(d, N/d)) .
2 d|N
We give a few examples of modular forms on subgroups. First note the following
easy lemma:
Lemma 6.5. If F ∈ Mk (Γ0 (N)) then for any m ∈ Z≥1 we have F(mτ ) ∈ Mk (Γ0 (mN)).
Proof. Trivial since when ac db ∈ Γ0 (mN) one can write (m(aτ + b)/(cτ + d)) =
(a(mτ ) + mb)/((c/m)τ + d). ⊓
⊔
Thus we can already construct many forms on subgroups, but in a sense they are
not very interesting, since they are “old” in a precise sense that we will define below.
An Introduction to Modular Forms 45
A second more interesting example is Eisenstein series: there are more general
Eisenstein series than those that we have seen for Γ , but we simply give the follow-
ing important example: using a similar proof to the above lemma we can construct
Eisenstein series of weight 2 as follows. Recall that E2 (τ ) = 1 − 24 ∑n≥1 σ1 (n)qn is
not quite modular, and that E2∗ (τ ) = E2 (τ ) − 3/(π ℑ(τ )) is weakly modular (but of
course non-holomorphic). Consider the function F(τ ) = NE2 (N τ ) − E2 (τ ), analo-
gous to the construction of the lemma with a correction term.
We have the evident but crucial fact that we also have F(τ ) = NE2∗ (N τ ) − E2∗ (τ )
(since ℑ(τ ) is multiplied by N), so F is also weakly modular on Γ0 (N), but since it
is holomorphic we have thus constructed a (nonzero) modular form of weight 2 on
Γ0 (N).
A third important example is provided by theta series. This would require a book
in itself, so we restrict to the simplest case. We have seen in Corollary 1.3 that the
2
function T (a) = ∑n∈Z e−aπ n satisfies T (1/a) = a1/2 T (a), which looks like (and is)
a modularity condition. This was for a > 0 real. Let us generalize and for τ ∈ H
set
2 2
θ (τ ) = ∑ qn = ∑ e2π in τ ,
n∈Z n∈Z
so that for instance we simply have T (a) = θ (ia/2). The proof of the functional
equation for T that we gave using Poisson summation is still valid in this more
general case and shows that
Exercise 6.6. 1. Using the dimension formulas, show that 2E2 (2τ )− E2 (τ ) together
with 4E2 (4τ ) − E2 (τ ) form a basis of M2 (Γ0 (4)).
2. Using the Fourier expansion of E2 , deduce an explicit formula for the Fourier
expansion of θ 4 , and hence that r4 (n), the number of representations of n as
a sum of 4 squares (in Z, all permutations counted) is given for n ≥ 1 by the
formula
46 Henri Cohen
Remark 6.7. Using more general methods one can give “closed” formulas for rk (n)
for k = 1, 2, 3, 4, 5, 6, 7, 8, and 10, see e.g., [1].
We can introduce the same Hecke operators as before, but to have a reasonable
definition we must add a coprimality condition: we define T (n)(∑m≥0 a(m)qm ) =
∑m≥0 b(m)qm , with
b(m) = ∑ d k−1a(mn/d 2) .
d|gcd(m,n)
gcd(d,N)=1
that they preserve modularity, so in particular the T (n) form a commutative algebra
of operators on Sk (Γ0 (N)). And this is where the difficulties specific to subgroups of
Γ begin: in the case of Γ we stated (without proof nor definition) that the T (n) were
Hermitian with respect to the Petersson scalar product, and deduced the existence of
eigenforms for all Hecke operators. Unfortunately here the same proof shows that
the T (n) are Hermitian when n is coprime to N, but not otherwise.
It follows that there exist common eigenforms for the T (n), but only for n co-
prime to N, which creates difficulties.
An analogous problem occurs for Dirichlet characters: if χ is a Dirichlet charac-
ter modulo N, it may in fact come by natural extension from a character modulo M
for some divisor M | N, M < N. The characters which have nice properties, in par-
ticular with respect to the functional equation of their L-functions, are the primitive
characters, for which such an M does not exist.
A similar but slightly more complicated thing can be done for modular forms.
It is clear that if M | N and F ∈ Mk (Γ0 (M)), then of course F ∈ Mk (Γ0 (N)). More
generally, by Lemma 6.5, for any d | N/M we have F(d τ ) ∈ Mk (Γ0 (N)). Thus we
want to exclude such “oldforms”. However it is not sufficient to say that a newform
An Introduction to Modular Forms 47
where B(d) is the operator sending F(τ ) to F(d τ ). Note that the sums in the above
formula are direct sums.
Exercise 6.8. The above formula shows that
2. Using Theorem 6.4, deduce a direct formula for the dimension of the new space.
Proposition 6.9. Let F ∈ Sk (Γ0 (N)) and WN = N0 −1
0 .
1. We have F|kWN ∈ Sk (Γ0 (N)), where
2. If F is an eigenform (in the new space) then F|kWN = ±F for a suitable sign ±.
Proof. (1): this simply follows from the fact that WN normalizes Γ0 (N): WN−1Γ0 (N)WN =
Γ0 (N) as can easily be checked, and the same result would be true for any other nor-
malizing operator such as the Atkin–Lehner operators which we will not define. The
operator WN is called the Fricke involution.
48 Henri Cohen
(2): It is easy to show that WN commutes with all Hecke operators T (n) when
gcd(n, N) = 1, so by what we have mentioned above, if F is an eigenform in the new
space it is automatically an eigenform for WN , and since WN acts as an involution,
its eigenvalues are ±1. ⊓
⊔
The eigenforms can again be normalized with a(1) = 1, and their L-function has
an Euler product, of a slightly more general shape:
1 1
L(F, s) = ∏ −s + pk−1 p−2s ∏ 1 − a(p)p−s
.
p∤N
1 − a(p)p p|N
Proposition 5.8 is of course still valid, but is not the correct normalization to obtain
a functional equation. We replace it by
Z ∞
N s/2 (2π )−sΓ (s)L(F, s) = F(it/N 1/2 )t s−1 dt ,
0
which of course is trivial from the proposition by replacing t by t/N 1/2 . Indeed,
thanks to the above proposition we split the integral at t = 1, and using the action of
WN we deduce the following proposition:
Proposition 6.10. Let F ∈ Sknew (Γ0 (N)) be an eigenform for all Hecke operators,
and write F|kWN = ε F for some ε = ±1. The L-function L(F, s) extends to a holo-
morphic function in C, and if we set Λ (F, s) = N s/2 (2π )−sΓ (s)L(F, s) we have the
functional equation
Λ (F, k − s) = ε i−k Λ (F, s) .
Proof. Indeed, the trivial change of variable t into 1/t proves the formula
Z ∞
N s/2 (2π )−sΓ (s)L(F, s) = F(it/N 1/2 )(t s−1 + ε ik t k−1−s ) dt ,
1
Once again, we leave to the reader to check that if F(τ ) = ∑n≥1 a(n)qn we have
Consider again the problem of sums of squares, in other words of the powers of
θ (τ ). We needed to raise it to a power which is a multiple of 4 so as to have a pure
modularity property as we defined it above. But consider the function
θ 2 (τ ). The
same proof that we mentioned for θ 4 shows that for any γ = ac db ∈ Γ0 (4) we have
An Introduction to Modular Forms 49
−4
θ 2 (γ (τ )) = (cτ + d)θ 2(τ ) ,
d
where −4 d is the Legendre–Kronecker character (in this specific case equal to
(−1) (d−1)/2 since d is odd, being coprime
to c). Thus it satisfies a modularity prop-
erty, except that it is “twisted” by −4d . Note that the equation makes sense since if
we change γ into −γ (which does not change γ (τ )), then
(cτ + d) is changed into
−(cτ + d), and −4 d is changed into −4
−d = − −4
d . It is thus essential that the
multiplier that we put in front of (cτ + d)k , here −4d , has the same parity as k.
We mentioned above that the study of modular forms on Γ1 (N) could be reduced
to those on Γ0 (N) “with a twist”. Indeed, more precisely it is trivial to show that
M
Mk (Γ1 (N)) = Mk (Γ0 (N), χ ) ,
χ (−1)=(−1)k
where χ ranges through all Dirichlet characters modulo N of the specified parity,
and where Mk (Γ0 (N), χ ) is defined as the space of functions F satisfying
fχ = ∑ χ (d) f |k Md .
0≤d<N, gcd(d,N)=1
These spaces are just as nice as the spaces Mk (Γ0 (N)) and share exactly the same
properties. They have finite dimension (which we do not give), there are Eisen-
stein series, Hecke operators, newforms, Euler products, L-functions, etc... An ex-
cellent rule of thumb is simply to replace any formula containing d k−1 (or pk−1 )
by χ (d)d k−1 (or χ (p)pk−1 ). In fact, in the Euler product of the L-function of an
eigenform we do not need to distinguish p ∤ N and p | N since we have
1
L(F, s) = ∏ 1 − a(p)p−s + χ (p)pk−1−2s ,
p∈P
The space M1 (Γ0 (4), χ−4 ) has dimension 1, generated by the single Eisenstein
series
(−4) n (D) D k−1
1 + 4 ∑ σ0 (n)q , where σk−1 (n) = ∑ d
n≥1 d|n
d
according to our rule of thumb (which does not tell us the constant 4). Comparing
(−4)
constant coefficients, we deduce that r2 (n) = 4σ0 (n), where as usual r2 (n) is the
number of representations of n as a sum of two squares. This formula was in essence
discovered by Fermat.
For r6 (n) we must work slightly more: θ 6 ∈ M3 (Γ0 (4), χ−4 ), and this space has
dimension 2, generated by two Eisenstein series. The first is the natural “rule of
thumb” one (which again does not give us the constant)
(−4)
F1 = 1 − 4 ∑ σ2 (n)qn ,
n≥1
where
(D,∗) D
σk−1 =∑ d k−1 ,
d|n
n/d
(D)
a sort of dual to σk−1 (these are my notation). Since θ 6 = 1 + 12q + · · ·, comparing
the Fourier coefficients of 1 and q shows that θ 6 = F1 + 16F2, so we deduce that
(−4) (−4,∗) −4 −4
r6 (n) = −4σ2 (n) + 16σ2 (n) = ∑ 16 −4 d2 .
d|n
n/d d
The explicit dimension formulas alluded to above are valid for k ∈ Z except for
k = 1; in addition, thanks to the theorems mentioned below, we also have explicit
dimension formulas for k ∈ 1/2 + Z. Thus, the theory of modular forms of weight 1
is very special, and their general construction more difficult.
This is also reflected in the construction of Galois representations attached to
modular eigenforms, which is an important and deep subject that we will not men-
tion in this course, except to say the following: in weight k ≥ 2 these representations
are ℓ-adic (or modulo ℓ), i.e., with values in GL2 (Qℓ ) (or GL2 (Fℓ )), while in weight
1 they are complex representations, i.e., with values in GL2 (C). The construction
in weight 2 is quite old, and comes directly from the construction of the so-called
Tate module T (ℓ) attached to an Abelian variety (more precisely the Jacobian of a
modular curve), while the construction in higher weight, due to Deligne, is much
An Introduction to Modular Forms 51
deeper since it implies the third Ramanujan conjecture |τ (p)| < p11/2 . Finally, the
case of weight 1 is due to Deligne–Serre, in fact using the construction for k ≥ 2
and congruences.
first example that we have met is of course the Ramanujan delta function
∆ (τ ) = η (τ )24 . Other examples are for instance η (τ )η (23τ ) ∈ S1 (Γ0 (23), χ−23 ),
η (τ )2 η (11τ )2 ∈ S2 (Γ0 (11)), and η (2τ )30 /η (τ )12 ∈ S9 (Γ0 (8), χ−4 ).
• Closely related to eta quotients are q-identities involving the q-Pochhammer
symbol (q)n and generalizing those seen in Exercise 3.30, many of which give
modular forms not related to the eta function.
• A much deeper construction comes from algebraic geometry: by the modular-
ity theorem of Wiles et al., to any elliptic curve defined over Q is associated a
modular form in S2 (Γ0 (N)) which is a normalized Hecke eigenform, where N
is the so-called conductor of the curve. For instance the eta quotient of level 11
just seen above is the modular form associated to the isogeny class of the elliptic
curve of conductor 11 with equation y2 + y = x3 − x2 − 10x − 20.
In this brief section, we will describe modular forms of a more general kind than
those seen up to now.
Coming back again to the function θ , the formulas seen above suggest that θ itself
must be considered a modular form, of weight 1/2. We have already mentioned that
2 −4
θ (γ (τ )) = (cτ + d)θ 2(τ ) .
d
But what about θ itself? For this, we must be very careful about the determination
of the square root:
Notation: z1/2 will always denote the principal determination of the square root,
i.e., such that −π /2 < Arg(z1/2 ) ≤ π /2. For instance (2i)1/2 = 1 + i, (−1)1/2 = i.
1/2 1/2
Warning: we do not in general have (z1 z2 )1/2 = z1 z2 , but only up to sign. As a
second notation, when k is odd, z will always denote (z1/2 )k and not (zk )1/2 (for
k/2
where v(γ , τ ) = ±1 and may depend on γ and τ . A detailed study of Gauss sums
shows that v(γ , τ ) = −4c
d , the general Kronecker symbol, so that the modularity
equation for θ is, for any γ ∈ Γ0 (4):
An Introduction to Modular Forms 53
c −4 −1/2
θ (γ (τ )) = vθ (γ )(cτ + d)1/2 θ (τ ) with vθ (γ ) = .
d d
Note that there is something very subtle going on here: this complicated theta mul-
tiplier system vθ (γ ) must satisfy a complicated cocycle relation coming from the
trivial identity θ ((γ1 γ2 )(τ )) = θ (γ1 (γ2 (τ ))) which can be shown to be equivalent to
the general quadratic reciprocity law.
The following definition is due to G. Shimura:
and if the usual holomorphy and conditions at the cusps are satisfied (equivalently
if F 2 ∈ M2k (Γ0 (N), χ 2 χ−4 )).
Note that if k ∈ 1/2 + Z we have vθ (γ )4k = χ−4 , which explains the extra factor
χ−4 in the above definition.
Since vθ (γ ) is defined only for γ ∈ Γ0 (4) we need Γ0 (N) ⊂ Γ0 (4), in other words
4 | N. In addition, by definition vθ (γ )(cτ + d)1/2 = θ (γ (τ ))/θ (τ ) is invariant if we
change γ into −γ , so if k ∈ 1/2 + Z the same is true of vθ (γ )2k (cτ + d)k , hence it
follows that in the above definition we must have χ (−d) = χ (d), i.e., χ must be an
even character (χ (−1) = 1).
As usual, we denote by Mk (Γ0 (N), χ ) and Sk (Γ0 (N), χ ) the spaces of modular
and cusp forms. The theory is more difficult than the theory in integral weight, but
is now well developed. We mention a few items:
1. There is an explicit but more complicated dimension formula due to J. Oesterlé
and the author.
2. By a theorem of Serre–Stark, modular forms of weight 1/2 are simply linear
combinations of unary theta functions generalizing the function θ above.
3. One can easily construct Eisenstein series, but the computation of their Fourier
expansion, due to Shimura and the author, is more complicated.
4. As usual, if we can express θ m solely in terms of Eisenstein series, this leads
to explicit formulas for rm (n), the number of representation of n as a sum of m
squares. Thus, we obtain explicit formulas for r3 (n) (due to Gauss), r5 (n) (due to
Smith and Minkowski), and r7 (n), so if we complement the formulas in integral
weight, we have explicit formulas for rm (n) for 1 ≤ m ≤ 8 and m = 10.
5. The deeper part of the theory, which is specific to the half-integral weight case,
is the existence of Shimura lifts from Mk (Γ0 (N), χ ) to M2k−1 (Γ0 (N/2), χ 2 ), the
description of the Kohnen subspace Sk+ (Γ0 (N), χ ) which allows both the Shimura
lift to go down to level N/4, and also to define a suitable Atkin–Lehner type new
space, and the deep results of Waldspurger, which nicely complement the work
of Shimura on lifts.
54 Henri Cohen
We could try to find other types of interesting modularity properties than those
coming from θ . For instance, we have seen that the Dedekind eta function is a
modular form of weight 1/2 (not in Shimura’s sense), and more precisely it satisfies
the following modularity equation, now for any γ ∈ Γ :
η (γ (τ )) = vη (γ )(cτ + d)1/2η (τ ) ,
η 2 (τ + 1/2) η 5 (2τ )
θ (τ ) = = 2 .
η (2τ + 1) η (τ )η 2 (4τ )
Exercise 7.3. 1. Prove these relations in the following way: first show that
the right-
hand sides satisfy the same modularity equations as θ for T = 1 1 and W =
01 4
0 −1 , so in particular that they are weakly modular on Γ (4), and second show
4 0 0
that they are really modular forms, in other words that they are holomorphic on
H and at the cusps.
2. Using the definition of η , deduce two product expansions for θ (τ ).
We could also try to study modular forms of fractional or even real weight k
not integral or half-integral, but this would lead to functions with no interesting
arithmetical properties.
In a different direction, we can relax the condition of holomorphy (or meromor-
phy) and ask that the functions be eigenfunctions of the hyperbolic Laplace operator
2
∂ ∂2 ∂2
∆ = −y2 2
+ 2
= −4y2
∂ x ∂ y ∂ τ∂ τ
which can be shown to be invariant under Γ (more generally under SL2 (R)) to-
gether with suitable boundedness conditions. This leads to the important theory of
Maass forms. The case of the eigenvalue 0 reduces to ordinary modular forms since
∆ (F) = 0 is equivalent to F being a linear combination of a holomorphic and an-
tiholomorphic (i.e., conjugate to a holomorphic) function, each of which will be
modular or conjugate of modular.
The case of the eigenvalue 1/4 also leads to functions having nice arithmetical
properties, but all other eigenvalues give functions with (conjecturally) transcen-
dental coefficients, but these functions are useful in number theory for other rea-
sons which we cannot explain here. Note that a famous conjecture of Selberg as-
serts that for congruence subgroups there are no eigenvalues λ with 0 < λ < 1/4.
An Introduction to Modular Forms 55
For instance, for the full modular group, the smallest nonzero eigenvalue is λ =
91.1412 · · ·, which is quite large.
Exercise 7.4. Using the fact that ∆ is invariant under Γ show that ∆ (ℑ(γ (τ ))) =
s(1 − s)ℑ(γ (τ )) and deduce that the nonholomorphic Eisenstein series E(s) intro-
duced in Definition 5.14 is an eigenfunction of the hyperbolic Laplace operator with
eigenvalue s(1 − s) (note that it does not satisfy the necessary boundedness condi-
tions, so it is not a Maass form: the functions E(s) with ℜ(s) = 1/2 constitute what
is called the continuous spectrum, and the Maass forms the discrete spectrum of ∆
acting on Γ \H ).
The last generalization that we want to mention (there are much more!) is to several
variables. The natural idea is to consider holomorphic functions from H r to C,
now for some r > 1, satisfying suitable modularity properties. If we simply ask
that γ ∈ Γ (or some subgroup) acts component-wise, we will not obtain anything
interesting. The right way to do it, introduced by Hilbert–Blumenthal, is to consider
a totally real number field K of degree r, and denote by ΓK the group of matrices
γ = ac db ∈ SL2 (ZK ), where ZK is the ring of algebraic integers of K (we could
also consider the larger group GL2 (ZK ), which leads to a very similar
theory).
Such
ai bi
a γ has r embeddings γi into SL2 (R), which we will denote by γi = ci di , and the
correct definition is to ask that
Note that the restriction to totally real number fields is due to the fact that for γi to
preserve the upper-half plane it is necessary that γi ∈ SL2 (R). Note also that the γi
are not independent, they are conjugates of a single γ ∈ SL2 (ZK ).
A holomorphic function satisfying the above is called a Hilbert-Blumenthal mod-
ular form (of parallel weight k, one can also consider forms where the exponents
for the different embeddings are not equal), or more simply a Hilbert modular form
(note that there are no “conditions at infinity”, since one can prove that they are
unless K = Q).
automatically satisfied
Since T = 10 11 ∈ SL2 (ZK ) is equal to all its conjugates,
such modular forms
have Fourier expansions, but using the action of 10 α1 with α ∈ ZK it is easy to
show that these expansions are of a special type, involving the codifferent d−1 of K,
which is the fractional ideal of x ∈ K such that Tr(xZK ) ⊂ Z, where Tr denotes the
trace.
One can construct Eisenstein series, here called Hecke–Eisenstein series, and
compute their Fourier expansion. One of the important consequences of this com-
putation is that it gives an explicit formula for the value ζK (1 − k) of the Dedekind
zeta function of K at negative integers (hence by the functional equation of ζK , also
56 Henri Cohen
at positive even integers), and in particular it proves that these values are rational
numbers, a theorem due to C.-L. Siegel as an immediate consequence of Theorem
3.41. An example is as follows:
√
Proposition 7.5. Let K = Q( D) be a real quadratic field with D a fundamental
discriminant. Then:
1. We have
1 D − s2
ζK (−1) =
60 ∑√ σ1 4 ,
|s|< D
1 D − s2
120 ∑√
ζK (−3) = σ3 .
4
|s|< D
Note that this formula can be generalized to arbitrary D, and is due to Smith and
(much later) to Minkowski. There also exists a similar formula for r7 (D): when −D
(not D) is a fundamental discriminant
D
r7 (D) = −28 41 − 4 L(χ−D , −2) .
2
There exist three software packages which are able to compute with modular forms:
magma, Sage, and Pari/GP since the spring of 2018. We give here some basic
Pari/GP commands with little or no explanation (which is available by typing ?
or ??): we encourage the reader to read the tutorial tutorial-mf available with
the distribution and to practice with the package, since it is an excellent way to learn
about modular forms. All commands begin with the prefix mf, with the exception
of lfunmf which more properly belongs to the L-function package.
Creation of modular forms: mfDelta (Ramanujan Delta), mfTheta (ordinary
theta function), mfEk (normalized Eisenstein series Ek ), more generally mfeisenstein,
mffrometaquo (eta quotients), mffromqf (theta function of lattices with or
without spherical polynomial), mffromell (from elliptic curves over Q), etc...
Arithmetic operations: mfcoefs (Fourier coefficients at infinity), mflinear
(linear combination, so including addition/subtraction and scalar multiplication),
mfmul, mfdiv, mfpow (clear), etc...
Modular operations: mfbd, mftwist, mfhecke, mfatkin, mfderivE2,
mfbracket, etc...
Creation of modular form spaces: mfinit, mfdim (dimension of the space),
mfbasis (random basis of the space), mftobasis (decomposition of a form on
the mfbasis), mfeigenbasis (basis of normalized eigenforms).
Searching for modular forms with given Fourier coefficients:
mfeigensearch, mfsearch.
Expansion of F|k γ : mfslashexpansion.
Numerical functions: mfeval (evaluation at a point in H or at a cusp), mfcuspval
(valuation at a cusp), mfsymboleval (computation of integrals over paths in the
completed upper-half plane), mfpetersson (Petersson scalar product), lfunmf
(L-function associated to a modular form), etc...
Note that for now Pari/GP is the only package for which these last functions
(beginning with mfslashexpansion) are implemented.
58 Henri Cohen
The literature on modular forms is vast, so I will only mention the books which I
am familar with and that in my opinion will be very useful to the reader. Note that
the classic book [4] is absolutely remarkable, but may be difficult for a beginning
course.
In addition to the recent book [1] by F. Strömberg and the author (which of course
I strongly recommend !!!), I also highly recommend the paper [5], which is essen-
tially a small book. Perhaps the most classical reference is [3]. The more recent book
[2] is more advanced since its ultimate goal is to explain the modularity theorem of
Wiles et al.
References