Determination of normalized extremal quasimodular forms of depth 1 with integral Fourier coefficients
Tomoaki Nakaya
Faculty of Mathematics, Kyushu University, 744, Motooka, Nishi-ku, Fukuoka, 819-0395, Japan
t-nakaya@math.kyushu-u.ac.jp
Abstract.
The main purpose of this paper is to determine all normalized extremal quasimodular forms of depth 1 whose Fourier coefficients are integers.
By changing the local parameter at infinity from to the reciprocal of the elliptic modular -function, we prove that all normalized extremal quasimodular forms of depth 1 have a hypergeometric series expression and that integrality is not affected by this change of parameters.
Furthermore, by transforming these hypergeometric series expressions into a certain manageable form related to the Atkin(-like) polynomials and using the lemmas that appeared in the study of -adic hypergeometric series by Dwork and Zudilin, the integrality problem can be reduced to the fact that a polynomial vanishes modulo a prime power, which we prove.
We also prove that all extremal quasimodular forms of depth 1 with appropriate weight-dependent leading coefficients have integral Fourier coefficients by focusing on the hypergeometric expression of them.
A quasimodular form of weight on the full modular group is given as
Here is the standard Eisenstein series on of weight defined by
where is a variable in the complex upper half plane , and is -th Bernoulli number, e.g., . For the intrinsic definition of the quasimodular forms on a non-cocompact discrete subgroup of , we refer to [41, §5.3].
We denote the vector space of modular forms and cusp forms of weight on by and , respectively. It is well known that for even , but is not modular and is quasimodular.
Any quasimodular form of weight can be uniquely written as
with and we call it depth of .
Hence is the quasimodular form with minimum weight and depth on .
Let denote the vector space of quasimodular forms of weight and depth on . In particular, is equal to . (In the following, we often omit the reference to the group .)
From the fact , the generating function of the dimension of is given by the following (See [10] for the explicit formula of .):
In particular, since there is no quasimodular form in which the depth exceeds half the weight,
holds, so by taking the limit in the above equation, we have
We denote by the differential operator . It is well known that the Eisenstein series and satisfy the following differential relation (equation) by Ramanujan ([5, Thm. 0.21]. See also [24, §3] and [25, §4.1] for Halphen’s contribution.):
(1)
Therefore, the ring is closed under the derivation .
For , we denote .
The notion of extremal quasimodular forms was introduced and studied by Kaneko and Koike in [19]. They defined it as follows using the vanishing order of the quasimodular form .
Definition 1.
Let and . We call extremal if .
If, moreover, , is said to be normalized. We denote by the normalized extremal quasimodular form of weight and depth on (if exists).
We set
Pellarin gave an upper bound on in [30, Thm. 2.3], which turns out to be for the depth .
In a recent paper [10] Grabner constructed inductively the normalized extremal quasimodular forms of weight and depth . More precisely, he constructed the quasimodular form such that
inductively by using the Serre derivative.
From Definition 1 of the extremal quasimodular forms, it is necessary to confirm that , which follows from , and hence .
Moreover, these extremal quasimodular forms exist uniquely for .
This is because if the two forms satisfy , then the vanishing order of is given by , which contradicts the definition of .
In contrast to these cases of depth , the existence and the uniqueness of for the depth is still open.
Let be the set of weights such that the normalized extremal quasimodular form of weight and depth on has integral Fourier coefficients.
In [15], based on Grabner’s results, Kaminaka and Kato completely determined the sets , and showed that the set is a subset of a finite set of cardinality 22.
Our main theorem asserts that their finite set coincides with .
Theorem 1.
We have
(2)
Note that there is no quasimodular form of weight and depth , since there is no modular form of weight 2 on .
The minimum weight for the normalized extremal quasimodular forms with non-integral Fourier coefficients is 26, and the Fourier coefficients belong to :
Here are some examples of the normalized extremal quasimodular forms of lower weight and depth 1. Obviously, the following forms have integral Fourier coefficients:
where is the “discriminant” cusp form of weight 12 on :
Using this infinite product expression of or the differential relation (1), we see that .
The first few non-trivial examples of are given below.
In [15, Rem. 1.2], Grabner pointed out to Kato that the integrality of these Fourier coefficients can be proved by using a classical congruence formula between Ramanujan’s tau-function , defined by , and (see also Section 5 for the integrality of ).
Since holds (see Proposition 2), we also see that .
However, it is probably difficult to express the Fourier coefficients of, say, in terms of known number-theoretic functions and to show congruence formulas that they satisfy.
The key idea that avoids this difficulty is to use the reciprocal of the elliptic modular function instead of as the local parameter at infinity (or equivalently, at the cusp of ).
As will be shown later, Lemma 1 guarantees that the integrality of the coefficients is equivalent when expanded for each local parameter. Furthermore, this idea also provides a unified method of proof for all weights by using generalized hypergeometric series.
Observing the examples of above, we find that its Fourier coefficients are positive, except for . More generally, Kaneko and Koike conjectured in [19] that the Fourier coefficients of would be positive if and .
For this positivity conjecture, Grabner showed in [11] that the conjecture is true if and , by proving the asymptotic formula for the Fourier coefficients of .
Thus, from Theorem 1, we can conclude that the Fourier coefficients of are positive integers if and only if their weight belongs to .
The paper is organized as follows.
Sections 2 to 4 are related to the main theorem. In Section 2 we derive that the normalized extremal quasimodular forms of depth 1 have a hypergeometric expression, and rewrite them using the Atkin-like polynomials and their adjoint polynomials .
Then the main theorem can be reduced to the fact that a formal power series vanishes modulo a prime power, since the polynomial part arising from the “factorization” of a formal power series vanishes.
To “factorize” the formal power series, we use some results on -adic hypergeometric series by Dwork and Zudilin. In Section 3, we specialize their results and prove several propositions.
In Section 4 we prove the main theorem of this paper by combining the results of the previous sections.
In Section 5 we explicitly construct extremal quasimodular forms of depth 1 with integral Fourier coefficients by focusing on their hypergeometric expression.
In Section 6 we summarize some previous work on the normalized extremal quasimodular forms of depth 2 to 4 and present some supplementary results, in particular from a hypergeometric point of view. We also present some previous work in which the group is replaced by a low-level congruence subgroup or Fricke group for (extremal) quasimodular forms of depth 1.
In Appendices we give: (i) A table of the integral Fourier coefficients of the normalized extremal quasimodular forms of depth 1. (ii) Explicit formulas for the coefficients of a certain formal power series that appears in the proof of the main theorem.
2. as a formal power series of
Let be non-negative integers and with . The generalized hypergeometric series is defined by
where denotes the Pochhammer symbol. This series is clearly invariant under the interchange of each of the parameters , and the same is true for .
When , the series satisfies the differential equation
where and are some constants that depend on the parameters and . Using the Euler operator , the above differential equation can be rewritten as follows:
Here we collect some hypergeometric series identities that we will use in later discussions.
(3)
(4)
(5)
(6)
See [2, pp. 85-86] for the proof and historical background of the equations (5) and (6).
Proposition 1.
For sufficiently large , we have
(7)
(8)
(9)
(10)
The hypergeometric expressions of the Eisenstein series and are classical and well known, for the proof, see [36]. In particular, is one of the solutions of a hypergeometric differential equation at , and for the solution at , we refer to [1]. To obtain the second equation of (7), we use Clausen’s formula (5).
See [41, §5.4] for a more general explanation of the fact that a (holomorphic or meromorphic) modular form satisfies a linear differential equation with a modular function as a variable.
In contrast to and , the hypergeometric expression (9) of is less well known. To the best of the author’s knowledge, the expression can be found in the author’s Ph.D. thesis [27, Ch. 3] and [31, Thm. 5]. There is also an equivalent expression in [26, Ch. 2.6], although it looks a little different.
For the convenience of the reader, we will briefly review the proof in [27] here; by calculating the logarithmic derivative of , we have
In the fourth equality we used the fact that , which can be calculated with (1).
Equation (10) is obtained by transforming Equation (9) using Orr’s formula as as follows.
Alternatively, the expressions (9) and (10) can be obtained by setting in equations (21) and (58), since .
By setting , Proposition 1 immediately implies the following theorem, which is a generalization of Theorem 5 in [36] for .
Theorem 2.
Put , and then we have the ring isomorphism
(11)
Moreover, since is closed under the derivation , the ring is closed under the derivation . In other words, the isomorphism holds as a graded differential algebra over .
Based on the ring isomorphism (11), we can consider the problem of the integrality of the Fourier coefficients of normalized extremal quasimodular forms as an equivalence problem in .
Example 1.
The formula in equivalent to the formula in can be calculated as follows.
Hence we obtain . The equations corresponding to the remaining equations in (1) can be calculated in a similar way.
We introduce the Serre derivative (or Ramanujan–Serre derivative) defined by
From this definition, it is clear that the Leibniz rule is satisfied. According to convention, we use the following symbols for the iterated Serre derivative:
It is well known that for even and , which is a special case of Proposition 3.3 in [19].
By rewriting (1) using the Serre derivative, we obtain the following useful consequences.
(12)
Note that the Serre derivative is not necessarily depth-preserving, as can be seen from the first equation in (12). Therefore, it is necessary to confirm the depth of the quasimodular form created by repeatedly applying a differential operator including the Serre derivative as in (17).
Now we consider the following differential equation, called the Kaneko–Zagier equation or, in a more general context, a second-order modular linear differential equation:
(13)
This differential equation (with replaced by ) first appeared in the study of the -invariants of supersingular elliptic curves in [21] and was characterized in [17, §5]. Although we consider quasimodular form solutions of (13) in this paper, it should be emphasized that the Kaneko–Zagier equation (13) has modular form solutions of level ([17]), and ([16]) and mixed mock modular form solutions ([12]) for appropriate .
It is easy to rewrite the differential equation (13) as , where
(14)
We also define the differential operators ([10]) and its “adjoint” as follows:
(15)
Then we have the following identity for the composition of the differential operators.
(16)
We will not give the proof, since it is a simple calculation by using Ramanujan’s identity (1). However, this identity is important for investigating the inductive structure of the solutions of the differential equation .
Indeed, if a function satisfies , we see that satisfies .
Therefore, the identity (16) gives not only the structure of the quasimodular solution of (13), but also that of a modular solution, a logarithmic solution, and a formal -series solution.
The differential operator (16) can be considered as a third-order differential operator, so that the given second-order differential operators and are in the left and right factors, respectively.
From this point of view, the factors and its adjoint are uniquely and independently determined from the Fourier coefficients of the extremal quasimodular forms, up to a constant multiple.
We define the sequence of power series by the following differential recursions:
(17)
(18)
The proportionality constants appearing in the definition are chosen so that the leading coefficients of and are as follows. Since , we have for any with the aid of (16). Then the Frobenius ansatz (see [10, §2.3]) gives
By applying the differential operators and act on this power series, we have
Note that can also be expressed as .
From this construction and the property of the Serre derivative , it is clear that for and for even integer . Therefore, from the same discussion in Section 1, the power series is nothing but the normalized extremal quasimodular form , we have for .
By changing the variables , the Kaneko–Zagier equation (13), which is equivalent to , is transformed into
or equivalently (Recall that .)
where . This differential equation is a hypergeometric differential equation.
Since we now assume , the solution space of the above differential equation is spanned by the power series solution
(19)
and a logarithmic solution. Therefore, since the extremal quasimodular forms contain no logarithmic terms, we obtain the hypergeometric expression of for .
Proposition 2.
The normalized extremal quasimodular forms of even weight and depth on have the following hypergeometric expressions.
where
(20)
(21)
Proof.
The hypergeometric expression of can be obtained by setting in (19) and multiplying by .
Similar expression of can be obtained by applying the Serre derivative to as follows ():
∎
Remark 1.
When , the Kaneko–Zagier equation (13) becomes the equation , which has 1 and as independent solutions.
In general, for , the two-dimensional solution space of the differential equation (13) has already been discussed in [17, §5], and its basis is given by and
where and are Atkin-like polynomials that appear later in this section, and the numbers and are defined by (27) and (28). The solution can also be obtained by the recurrence formula (17), where the initial value is replaced by .
It is well-known that the space has a basis for with . This basis is characterized by the fact that the leading terms of its Fourier expansion are .
We can construct a basis for with the same property using the normalized extremal quasimodular forms of depth .
Proposition 3.
For any even integer , a basis of the space is given by the following set: where
the notation is that in Theorem 2 and Proposition 2.
Proof.
We give a proof only for the case of , the remaining cases being similar. We put .
For any quasimodular form , we can determine the coefficient such that . Since , the form must be , and hence is spanned by . For the linear independence property, by comparing the Fourier coefficients, we have .
∎
Example 2.
Case of weight : The following equation holds.
where
Lemma 1.
The following claims hold.
(1)
.
(2)
.
(3)
.
(4)
.
(5)
and .
Proof.
(1)
It is clear that the following facts imply the claim:
(2)
Since , holds, and we have .
(3)
Using claim (2), we know that the coefficients satisfy and when the function is expressed in two ways as . This gives us the assertion.
(4)
It is clear from the facts that and .
(5)
By combining (7), (29) and (30), we have and hence and .
By Proposition 2, holds, so if , then and hence (note (3)). And conversely, if , we have and therefore holds.
The case of can be proved similarly.
∎
Remark 2.
In the above proof we used that , but a stronger claim holds. The hypergeometric series satisfies the following quadratic transformation of Gauss:
(22)
By putting in the above equation, we have
where is the -th Catalan number. (The second equality holds for .) Thus we have or equivalently .
Actually, it is known the stronger claim holds. See [13] for more details and related results.
We classify each element of the set (2) according to the remainder of modulo 12 as follows:
Here the symbol means that for each element in the list the number is added. From (4) of Lemma 1, it suffices to show that the function has integral Fourier coefficients when the weight is an element of the set .
Put and . The first few coefficients are given by
Consider the condition that these coefficients are integers.
In such a case, the denominator must divide the numerator of the above equations, and so and are integers if . Similarly, if , and are integers.
In these lists of , we exclude the case of , because it corresponds to the trivial case .
From (5) of Lemma 1, in order to prove the main theorem, we should show that the power series and are actually have integral coefficients for these lists of . That is, we prove that
•
, and then for ,
•
, and then for .
To prove these assertions, we will now rewrite the formal power series and into a more manageable form.
Although the symbols are slightly different, in their paper [19], Kaneko and Koike expressed the normalized extremal quasimodular forms by using the monic polynomials and as follows:
(23)
(24)
(25)
(26)
where the normalizing factor is given below for any except for .
(27)
(28)
Here, and are integers, since from the proof of (1) of Corollary 1.
In particular, the polynomial is equal to the original Atkin polynomial treated in [21].
We will refer to the polynomials as the Atkin-like polynomials.
Note that we have the polynomial corresponding to the quasimodular part as “”, which is different between us and [19].
The reason we do not follow the symbols of [21] is to write the best rational function approximation in a unified way, which we will discuss shortly afterwards.
For the convenience of the reader, we provide below a comparison table of these symbols. In the table, polynomials in the same column are equal.
Table 1. Comparison table of the Atkin-like polynomials and its adjoint polynomials
We also note that the polynomials and are the denominator and numerator, respectively, of the best rational-function approximation111If it is not the best approximation, it contradicts the extremality of . As a side note, for example, to get the rational function by Mathematica for certain , we just have to enter PadeApproximant, where and . of the following power series.
where is the power series with as the variable (recall Proposition 1), and then .
More generally, for some suitable polynomial , the orthogonal polynomial that appears when approximating the function has already been considered in [3] by Basha, Getz and Nover. However, they are treated in the context of a generalization of the Atkin orthogonal polynomials, and not from the point of view of extremal quasimodular forms.
Here and throughout this paper, the symbols , , and are defined as the following positive integer and formal power series, respectively:
(29)
(30)
(31)
Note that by using Proposition 1, and hold. Therefore, and hold with the aid of Lemma 1.
For a given polynomial , we denote its reciprocal polynomial as , i.e.,
Comparing equations (23) to (26) with Proposition 2 and using the hypergeometric expression of Eisenstein series in Proposition 1, we have
(32)
(33)
(34)
(35)
As we will see later in Section 4, the polynomials and are not necessarily elements of .
In such a case, multiply both sides of the above equation by an appropriate factor so that the right-hand side of (32) to (35) has integral coefficients.
Under this normalization, if all the coefficients of the formal power series on the right-hand side are congruent to modulo , then we can conclude that the corresponding formal power series and have integral coefficients. Here, for and , note (1) in Lemma 1.
Thus, in the next section we will investigate in detail the congruence formulas of the formal power series and modulo prime powers.
The above equations (32) to (35) are essentially the relations satisfied by a hypergeometric series , and can be interpreted in the following two ways.
(1)
Focusing on the left-hand side, from the definitions of the series and , the term depending on is a hypergeometric series with a certain rational parameter shifted by an integer.
Thus, the equations (32) to (35) are concrete expressions of certain contiguous relations (or three term relations) of a hypergeometric series .
For more general results of this kind, see [8] by Ebisu.
(2)
Focusing on the right-hand side, the polynomials of a certain degree or less that are multiplied by the series and are chosen so that the vanishing order for on the left-hand side is as large as possible.
Therefore, such polynomials are Hermite–Padé approximations for and , and the left-hand side is the corresponding remainder.
For the Hermite–Padé approximation of generalized hypergeometric series , see [28] by Nesterenko.
Unfortunately, the case treated in the papers [8] and [28] is that the corresponding hypergeometric differential equation has no logarithmic solution, which does not include our case (recall Remark 1).
3. Congruence formulas for and modulo prime powers
First, note that in the prime factorization of the normalizing factor , which appears in the proof of the main theorem in Section 4, the exponents of all prime factors above 11 are 1. On the other hand, under the proper normalization described at the end of the previous section, the exponents of the prime factors 2, 3, 5, and 7 are at most 8, 5, 2, and 2, respectively.
Therefore, to prove the main theorem, we need to calculate the specific congruence formulas for and modulo and the appropriate prime . (Of course, results for the largest exponents are valid for smaller ones.)
More specifically, in this section we will prove that and modulo these prime powers give some polynomials or rational functions.
Throughout this section, for the formal power series and , the symbol means that all the power series coefficients of their difference are congruent to 0 modulo .
If the prime is greater than or equal to 5, the following proposition corresponds to the special case of Lemma 2, but here we give a simpler alternative proof using Lucas’ theorem.
Proposition 4.
For any prime , we have
Thus, in particular, for and .
Proof.
We first recall Lucas’ theorem for the binomial coefficient modulo a prime ; for some nonnegative integers and such that ,
(36)
holds under the usual convention , and if .
For some generalizations of Lucas’ theorem, we refer to the extensive historical survey [22] by Meštrović.
To calculate for and , we classify by the value of .
From Lucas’ theorem, it is easy to see that if , the following congruence holds:
Similarly, if , since and , so
Now we assume that , and thus the binomial coefficient vanishes.
Also, if and , and vanish, respectively. Therefore, we have
The claim for is obtained from a similar calculation, noting that if , then . If then holds, we see that and hence for .
∎
Remark 3.
For , we can prove that . Since , and so holds. By comparing (7) and (30), we have and obtain the desired result.
Lemma 2.
Let be a prime number. Let
and denotes the -th derivative . Then the following congruence formulas hold for any :
Proof.
These assertions correspond to the special cases of Lemma 3.4 (i) and (ii) in [7, p. 45], respectively. The lemma requires the following three assumptions:
a)
.
b)
for all .
c)
for all .
Note that since we now consider to be the formal power series, a domain in the original assumption c) can simply be the ring of -adic integers .
First, the assumption a) clearly holds.
Following [7, p. 30], for a given prime number and some , the symbol denotes the unique element such that .
Put and then
holds. Therefore, this corresponds to the case of in Corollary 2 of [7, p. 36], it can be seen that assumptions b) and c) hold.
∎
We rewrite the result of this lemma for into the following statement for and by changing the variables.
Proposition 5.
The following congruence formulas are valid.
(37)
(38)
(39)
(40)
Proof.
Assuming that and in Lemma 2, the direct calculation gives
and so
On the other hand, for from Proposition 4, and holds, thus we have
Hence, with the help of Lemma 2, we obtain the congruence formula (37) for :
(41)
Similarly, we have
and so
Note that changing the variables , since , we have
(42)
By combining (41) and (42), we obtain the congruence formula (38) for :
Next, assuming that and in Lemma 2, the direct calculation gives
and so
Unlike the case of , is not congruent modulo , so we first calculate the following congruence relation. Noting that the congruence ,
Replacing with in the above equation, we have
and hence
The calculation for is the same as for the case of . By substituting
into the corresponding part, we obtain
This completes the proof of the proposition.
∎
Unfortunately, since and , Lemma 2 cannot be applied directly to and .
Therefore, we consider the following formal power series to which Lemma 2 can be applied:
(43)
(44)
Lemma 3.
Suppose that the sequence of integers defined by (44) and let . Then for all nonnegative integers , one has
(45)
Although it looks a slightly different, this lemma corresponds (partially) to the case of of Lemmas 11 and 12 in [42] by Zudilin. We will not prove the above lemma again here, but will explain below how to rewrite the calculation in [42].
From the equation that appeared in the proof of Lemma 12 in [42], for all nonnegative integers such that and , we have
By putting and in this equation, we obtain
Since this congruence formula holds even when , we obtain the first assertion in (45).
Next, by combining [42, Eq. (34)] and the definition of a constant [42, Eq. (1)] depending on , we have
for all nonnegative integers such that .
The second assertion in (45) can be obtained by rewriting as in this equation. Consequently, Lemma 3 guarantees assumptions b) and c) of Lemma 3.4 in [7, p. 45], and hence the following lemma holds.
Lemma 4.
Let and be the formal power series and polynomials defined by (43) respectively. For , the following congruence formulas hold for any :
Here, denotes the -th derivative .
Proposition 6.
The following congruence formulas are valid.
(46)
(47)
(48)
(49)
Proof.
By putting and using the quadratic transformation (22), we have
and then
To explicitly calculate the difference between these formal power series, we focus on the following congruence formula. By changing the variable , we have
and then
On the other hand, from the power series expansion of calculated above,
Since we know that as mentioned in Remark 3, holds for , that is, the second term of the above equation is congruent to modulo . Therefore,
Hence we have and so .
By combining these congruences, we obtain the congruence formula (46) for :
Note that the last equality uses the congruence relation .
To obtain the formula (47), we calculate the change of the variable with the following:
Hence we have
Since equations (48) and (49) can be obtained by the similar calculation with and in Lemma 4, we omit the proof.
∎
From Propositions 5 and 6, we obtain the following corollary.
Corollary 1.
The following formal infinite product expressions hold.
Question 1.
Recall that and . What are the counterparts of the above congruence formulas in the (quasi) modular form side?
We still cannot answer this vague question, but we can prove the following assertion as a closely related result. (Recall that .)
Proposition 7.
Let be the Möbius function and be the Atkin inner product defined as
Then the formal power series can be expressed formally as follows:
(50)
(51)
Here are some examples of exponents .
Proof.
As already mentioned in [21, §5], the moment-generating function of the Atkin inner product is given by
(52)
By transforming the left-hand side of the above equation using the hypergeometric expressions of the Eisenstein series in Proposition 1, we have
Therefore, we have by comparing the coefficients of on both sides of the above equation, and obtain the desired expression (50) of by using the Möbius inversion formula. Equation (51) is obtained by dividing both sides of the equation
by and then integrating with respect to .
∎
Remark 4.
Equation (52), and therefore equation (51), is equivalent to:
4. Proof of the main theorem
First, we summarize the normalizing factors and the polynomials and that appear in the main theorem.
From the definition of the normalizing factor in Section 1, we see that its prime factors do not exceed the weight of the corresponding normalized extremal quasimodular form.
Case of weight . Prime factorization of normalizing factors.
Atkin-like polynomials.
Adjoint polynomials.
Case of weight . Prime factorization of normalizing factors.
(Original) Atkin polynomials.
Adjoint polynomials.
Case of weight . Prime factorization of normalizing factors.
Atkin-like polynomials.
Adjoint polynomials.
Case of weight . Prime factorization of normalizing factors.
We prove only the non-trivial and highest weight case .
The remaining cases can be proved in a similar way.
Since holds from (5) of Lemma 1, we transform the formal power series according to (34) as follows:
Since the polynomials and belong to , we multiply the both sides of the above equation by , so that the right-hand side is the power series with integral coefficients.
Using Proposition 4, we obtain the following congruence formula for prime numbers :
By direct calculation using Mathematica, we can see that the polynomial part of the right-hand side of the above equation is congruent to 0 modulo .
To perform similar calculations for modulo , and , we use Propositions 5 and 6. For example, from (46) and (47), we have
and then the polynomial part of the right-hand side is congruent to 0 modulo .
The same assertion holds for the remaining cases.
Combining the results for each of these primes, we have the desired congruence
The proof of Theorem 1 in this paper is a “hypergeometric” proof. Can this theorem be proved using only the theory of modular forms?
5. Other choice of the leading coefficients
Why did we choose the leading coefficient as 1 in Definition 1 of the normalized extremal quasimodular form? Was it really a natural choice? Of course, the set of weights of with integral Fourier coefficients changes if we choose a number other than 1 as the leading coefficient.
In this section we show that has integral Fourier coefficients if we choose a constant based on the hypergeometric expressions of in Proposition 2.
The reason we often focus on (positive) integers is that we can expect them to count “something”, such as the dimension or degree or order of some mathematical objects, the number of curves or points with certain arithmetic or geometric properties, and so on.
Interestingly, Nebe proved the integrality of the Fourier coefficients of by using the properties of the automorphism group of the Leech lattice in the appendix of [30]222The author speculates that Nebe probably uses to mean in the proof of Theorem A.1. Hence the correct expression of for is given by .:
Therefore, the number 393120 is twice the number of lattice vectors of squared norm 4 in , and is exactly equal to this number.
It is known that the Eisenstein series is the theta series of the -lattice. Since , as with , the Fourier coefficients of are related to the number of lattice vectors in the -lattice.
Note that although and were obtained as derivatives of a certain modular form, the only such that can be obtained this way are for . This fact is easily seen by comparing the dimension formulas of and .
Inspired by these coincidences333The modular solutions of the Kaneko–Zagier equation (13) with small weights are also closely related to the theta series of the ADE-type root lattice. For more details, see [17, p. 158]. and , although we still do not know what the Fourier coefficients of these forms counting up, we have arrived at the following theorem.
Theorem 3.
Let be the normalizing factor defined by (27) and (28). Then, for any , the Fourier coefficients of the following extremal quasimodular forms are all integers:
Furthermore, since , the forms and have the same properties. (Note: Since the factors and are defined by the product of binomial coefficients (27), we substitute after dividing by and reducing.)
In the proof of this theorem, we use the following power series instead of as the local parameter at infinity, based on the method of Remark 2:
(53)
Proof.
Throughout this proof, we consider to be and .
From Proposition 2, the normalized extremal quasimodular forms are the product of the power of and a hypergeometric series. Furthermore, since holds from Remark 2, we have . Therefore, as already mentioned in Lemma 1 (3), it suffices to show that to prove the theorem, where the constant is an appropriate normalization constant.
First we calculate the case of . From the quadratic transformation formula (22) we have
(54)
It is convenient to use the following formulas to calculate the last equality:
A similar calculation for yields
(55)
To prove the assertions about and , we recall . Then, we can see that the coefficient of of the following formal power series is an integer:
Considering the divisors of the normalizing factors , Theorem 3 implies the following claim.
Corollary 2.
The denominators of the Fourier coefficients of the normalized extremal quasimodular forms are only divisible by prime numbers .
This claim was originally stated as in
[19].
When and , they are proved in [30, Thm. 3.3] and [23, Thm. 1.8, Thm. 1.9], respectively, and the original claim is proved in [10]. Note that our proof is mainly based on hypergeometric expressions of , which is different from their proof.
According to Theorem 3, there exists a constant such that . However, this constant is probably too large, and integrality holds for smaller numbers. For example, although holds, we can show that the stronger claim holds in the similar way as in the proof of Theorem 1. As a more general setting, we define the number as .
How do such numbers depend on weight and depth?
Remark 5.
In the study of the denominator of Atkin polynomials [21, §9], the formal power series and , which are defined as follows, play a central role:
where and for ,
The initial power series are given by the following.
Although the definition for the power series is not given in [21], the details are omitted here because it can be derived from the same consideration as . In fact, these power series and the power series and defined by (20) and (21) have the following correspondences:
These correspondences can be proved by using a contiguous relation of a certain hypergeometric series. In [21, p. 120], it is stated without proof that “the power series has integral coefficients and is divisible by for all ”, which is essentially equivalent to a weaker version of our Theorem 3. However, it should be emphasized that essentially the equivalent forms appeared in [21] before the introduction of the normalized extremal quasimodular forms in [19].
6. Further directions
Up to the previous section we have discussed the extremal quasimodular forms of depth on . It is therefore natural to consider a generalization that changes these factors.
The case of depth on has already been considered in [19, 10, 15], but we note that in these papers there is no perspective of using a generalized hypergeometric series.
Therefore, we focus mainly on these hypergeometric aspects and present in this section some complementary results of previous studies without detailed proofs.
To describe the case of depth , we introduce the Rankin–Cohen brackets ([41, §5.2], see also [40, 4]), which is defined for integers and functions on by
We then define the -th order differential operator as follows.
In particular, when , this operator gives the Serre derivative and the differential operator defined by (14), respectively. Also, by specializing Proposition 3.3 in [19], we can see that .
The following identity holds for the integer :
(56)
The proof of this identity by direct calculation is long and tedious, but by using (56) repeatedly, we can easily rewrite the differential operator into a form using the Serre derivative for .
As will be described later, some extremal quasimodular forms are annihilated by the differential operator . Therefore, such rewriting is useful to derive an inductive structure of extremal quasimodular forms according to Grabner’s method.
Incidentally, the identity (56) does not seem to hold for , for example, we have the following identities for and :
In [19] Kaneko and Koike showed that (Recall the Kaneko–Zagier equation (13).) and , and conjectured that and . Hence, for , we also call the differential equation the Kaneko–Zagier equation.
Recently, Grabner [10] gave an affirmative answer for their conjecture, more precisely, he gave a concrete form of the differential equation satisfied by the balanced quasimodular forms.
By specializing his results, we can find the concrete form of the differential equation satisfied by the normalized extremal quasimodular forms for any even integer and and .
Of course, for and , there is a fifth-order differential operator such that .
The specific form of is not described here, but is of the form shown below:
where .
6.1. Case of depth
Depth 2. The Kaneko–Zagier equation is transformed into
or equivalently
where and . Since this differential equation is a hypergeometric differential equation, a similar statement as in Proposition 2 holds for depth .
Proposition 8.
The normalized extremal quasimodular forms of even weight and depth on have the following hypergeometric expressions.
(57)
(58)
From Grabner’s result [10], the relation holds, so the latter expression can be obtained from the former after a little calculation.
We omit the proof because it is similar to Proposition 2.
It is known from [15] that , and the Fourier coefficients of the corresponding normalized extremal quasimodular forms are given below:
The positivity of the Fourier coefficients of can also be seen from the fact that .
In these cases, it is easier to show the integrality by calculating the Fourier coefficients directly, but here we introduce a proof using a hypergeometric series as in the case of depth 1. We define
and then and hold. Since
the integrality of the former is obvious, while that of the latter is not.
Noticing that the exponents of the prime factors of are all , we can prove the latter integrality using the Lucas’ theorem in (36). For we set and use Lucas’ theorem only once, but for , we have to use it several times. For example, the calculation for and is performed as follows.
The following assertions are the depth counterparts of Proposition 3. The proof is the same, so it is omitted.
Proposition 9.
For each even integer , a basis of the space is given by the following set:
Depth 3. The Kaneko–Zagier equation is transformed into
where and .
This differential equation is not a hypergeometric differential equation. However, noting the “accidental relation” , we obtain the following expression of using a hypergeometric series:
It is known from [15] that , more specifically, we have
The positivity of the Fourier coefficients of can also be seen from the fact that .
In considering the meaning of integrality of Fourier coefficients, it is worth noting that the -th Fourier coefficient of gives the number of simply ramified coverings of genus and degree of an elliptic curve over . See [6, 20] for more details on this claim.
Proposition 10.
For any even integer , a basis of the space is given by the following set:
Note that the Fourier expansion of each element of the basis is slightly different from that of depths 1 and 2. More specifically, there are two forms and whose Fourier expansion is . Therefore, the form such that is expressed as a linear combination of , and .
Depth 4. The Kaneko–Zagier equation is transformed into
where and .
This differential equation is not a hypergeometric differential equation.
It is known from [15] that , that is, there are no normalized extremal quasimodular forms with integral Fourier coefficients. However, we note that holds as for depths 2 and 3.
Remark 6.
The fourth- and fifth-order differential equations that appear for depths and , respectively, are “rigid” in the sense of Katz (see [9, Rem. 17]).
Proposition 11.
For any even integer , a basis of the space is given by the following set:
The same (but slightly more complicated) phenomenon as in the case of depth occurs for the Fourier expansion of each element of these bases. For instance, there is no element in the set for which , but there are two elements for which . However, there is only one element for which , and there are two elements for which .
For a given even integer and depth , let as before, and express the normalized extremal quasimodular form as follows:
where the number is chosen so that the polynomial of degree is the monic polynomial. We call this polynomial the generalized Atkin polynomial.
Although it is possible to calculate the polynomial separately for a given and , as already mentioned in Section 1, the existence and uniqueness of for the depth are still open, it is not yet clear whether the polynomial determined by the above equation always exists for .
The following conjecture is suggested by numerical experiments .
Conjecture 1.
Let be a prime number and put . Then we have for .
Example 3.
The case of , that is, the case of :
For the prime , these polynomials satisfy the above conjecture.
For the depths greater than 5, almost nothing is known. For example, we do not even know what differential equation the extremal quasimodular form satisfies.
Unlike the case of depth , as already pointed out in [19], any extremal modular form of weight and depth cannot satisfy the differential equation .
Of course, it is possible to calculate specific examples separately, e.g., the normalized extremal quasimodular form of weight and depth
is the solution of the following modular linear differential equation:
As can be expected from this example, as the weights or depths increases, the coefficients of the differential equation become more complex and the calculations rapidly become unmanageable.
From the observation of some specific examples, it seems that the coefficients cannot be a simple polynomial with as a variable.
We also note that the denominator 11 of the Fourier coefficients of is greater than the weight 10. (cf. Corollary 2)
At the end of Section 2, we pointed out that the normalized extremal quasimodular forms of depth 1 are essentially the remainder of some Hermite–Padé approximation.
In fact, with the help of the ring isomorphism in Theorem 2, the extremal quasimodular forms of any depth can be regarded as the remainder of some Hermite–Padé approximation (of the first kind) for a suitable function.
On the other hand, there is an approximation called the Hermite–Padé approximation of the second kind (or simultaneous Padé approximation) [28].
For the depth , these two approximations are essentially the same, but not for the depth . What are the properties of the quasimodular forms derived from the simultaneous Padé approximation?
6.2. Case of other groups
In this subsection we summarize previous works on Atkin orthogonal polynomials and extremal quasimodular forms of depth 1 on the congruence subgroup and the Fricke group of low-levels.
If the level is less than or equal to 4, the corresponding quasimodular Eisenstein series of weight 2 can be expressed using the hypergeometric series (see [27, Ch. 3]), so it is probably possible to treat the normalized extremal quasimodular forms in the same way as in this paper.
On the other hand, for the Fricke groups of levels 5 and 7, the local Heun function appears instead of the hypergeometric series .
Because of this difference, it may be difficult to obtain various concrete expressions for these levels as in this paper.
Note also that in the case of level 1, the Atkin polynomials and the normalized extremal quasimodular forms are directly connected as in (24), but this is not necessarily the case for level .
The next table summarizes the previous works by Tsutsumi, Kaneko, Koike, Sakai, and Shimizu. Note that some extremal quasimodular forms also appear in [33], and the Kaneko–Zagier equation for is treated in [18], but its quasimodular solution is the quasimodular form on .
Table 2. The previous works related to this paper (with levels).
In the case of level 1, three differential equations appeared in [19, Thm. 2.1], but since there are relations (17) and (18), it is essentially sufficient to consider only the differential equation (13). Extend this result to level and clarify the relationship between quasimodular solutions of some differential equations using Grabner’s method.
Although of little relevance to quasimodular forms, it is worth noting that a certain basis of the vector space of modular forms of weight on is expressed by a generalized hypergeometric series . For more details on this fact, see [14, 9, 39] and its references.
Acknowledgement
The author thanks to Professor Masanobu Kaneko for many valuable comments on reading an early draft of this paper.
Appendix A Appendices
A.1. Table of the integral Fourier coefficients of
Put and then the first few terms of the integral Fourier coefficients of are given in the following table.
Table 3. The first few integral Fourier coefficients
1
2
3
4
5
2
6
18
84
292
630
1512
8
66
732
4228
15630
48312
10
258
6564
66052
390630
1693512
12
56
1002
9296
57708
269040
14
128
4050
58880
525300
3338496
16
296
16602
377456
4846908
41943120
18
99
3510
64944
764874
6478758
20
183
10134
269832
4326546
47862918
22
339
29430
1127904
24615834
355679478
24
144
7944
235840
4451130
59405952
28
384
44664
2460160
79196970
1693028352
30
190
14460
608570
16463120
314562708
32
286
29988
1652834
56608952
1335336084
34
430
62220
4496090
195047840
5680752948
38
336
43587
3065648
136437750
4219436160
54
378
62532
6109740
401161950
19083824856
58
618
155412
21940620
2005126350
128986599096
68
581
147042
21956168
2203554570
160242315903
80
678
204756
37135249
4592036697
416237464122
114
855
341886
85507600
15092041050
2010698806050
118
1095
549246
169413760
36358101930
5819797557810
A.2. Explicit formulas for the coefficients of and
We give explicit formulas for the coefficients of the power series and that appear in the proof of Theorem 1. However, as we have already seen in equations (33) and (34), we multiply the power series and by the factor .
Our main theorem is equivalent to the fact that the following power series belong to .
. The coefficients of the power series and correspond to the sequences A001421 and A145493 in the Online Encyclopedia of Integer Sequences (OEIS) [29], respectively.
.
where
and these polynomials are irreducible over .
.
. The coefficients of the power series and correspond to the sequences A145492 and A145494 in OEIS, respectively.
where
and these polynomials are irreducible over .
References
[1]
N. Archinard.
Exceptional sets of hypergeometric series.
J. Number Theory, 101(2):244–269, 2003.
[2]
W. N. Bailey.
Generalized hypergeometric series.
Cambridge Tracts in Mathematics and Mathematical Physics, No. 32.
Stechert-Hafner, Inc., New York, 1964.
[3]
S. Basha, J. Getz, H. Nover, and E. Smith.
Systems of orthogonal polynomials arising from the modular
-function.
J. Math. Anal. Appl., 289(1):336–354, 2004.
[4]
P. B. Cohen, Y. Manin, and D. Zagier.
Automorphic pseudodifferential operators.
In Algebraic aspects of integrable systems, volume 26 of
Progr. Nonlinear Differential Equations Appl., pages 17–47.
Birkhäuser Boston, Boston, MA, 1997.
[5]
S. Cooper.
Ramanujan’s theta functions.
Springer, Cham, 2017.
[6]
R. Dijkgraaf.
Mirror symmetry and elliptic curves.
In The moduli space of curves (Texel Island, 1994),
volume 129 of Progr. Math., pages 149–163. Birkhäuser
Boston, Boston, MA, 1995.
[7]
B. Dwork.
-adic cycles.
Inst. Hautes Études Sci. Publ. Math., (37):27–115, 1969.
[8]
A. Ebisu.
Three term relations for the hypergeometric series.
Funkcial. Ekvac., 55(2):255–283, 2012.
[9]
C. Franc and G. Mason.
Hypergeometric series, modular linear differential equations and
vector-valued modular forms.
Ramanujan J., 41(1-3):233–267, 2016.
[10]
P. J. Grabner.
Quasimodular forms as solutions of modular differential equations.
Int. J. Number Theory, 16(10):2233–2274, 2020.
[11]
P. J. Grabner.
Asymptotic expansions for the coefficients of extremal quasimodular
forms and a conjecture of Kaneko and Koike.
Ramanujan J., 57(3):1021–1041, 2022.
[12]
P. Guerzhoy.
A mixed mock modular solution of the Kaneko-Zagier equation.
Ramanujan J., 36(1-2):149–164, 2015.
[13]
N. Heninger, E. M. Rains, and N. J. A. Sloane.
On the integrality of th roots of generating functions.
J. Combin. Theory Ser. A, 113(8):1732–1745, 2006.
[14]
T. Ibukiyama.
Modular forms of rational weights and modular varieties.
Abh. Math. Sem. Univ. Hamburg, 70:315–339, 2000.
[15]
T. Kaminaka and F. Kato.
Extremal quasimodular forms of lower depth with integral Fourier
coefficients.
Kyushu J. Math., 75(2):351–364, 2021.
[16]
M. Kaneko.
On modular forms of weight satisfying a certain
differential equation.
In Number theory, volume 15 of Dev. Math., pages 97–102.
Springer, New York, 2006.
[17]
M. Kaneko and M. Koike.
On modular forms arising from a differential equation of
hypergeometric type.
volume 7, pages 145–164. 2003.
Rankin memorial issues.
[18]
M. Kaneko and M. Koike.
Quasimodular solutions of a differential equation of hypergeometric
type.
In Galois theory and modular forms, volume 11 of Dev. Math., pages 329–336. Kluwer Acad. Publ., Boston, MA, 2004.
[19]
M. Kaneko and M. Koike.
On extremal quasimodular forms.
Kyushu J. Math., 60(2):457–470, 2006.
[20]
M. Kaneko and D. Zagier.
A generalized Jacobi theta function and quasimodular forms.
In The moduli space of curves (Texel Island, 1994),
volume 129 of Progr. Math., pages 165–172. Birkhäuser
Boston, Boston, MA, 1995.
[21]
M. Kaneko and D. Zagier.
Supersingular -invariants, hypergeometric series, and Atkin’s
orthogonal polynomials.
In Computational perspectives on number theory (Chicago, IL,
1995), volume 7 of AMS/IP Stud. Adv. Math., pages 97–126. Amer.
Math. Soc., Providence, RI, 1998.
[22]
R. Meštrović.
Lucas’ theorem: its generalizations, extensions and applications
(1878–2014), 2014.
arXiv:1409.3820v1[math.NT].
[23]
A. Mono.
On a conjecture of Kaneko and Koike, 2020.
arXiv:2005.06882v1[math.NT].
[24]
J. A. C. Morales, H. Movasati, Y. Nikdelan, R. Roychowdhury, and M. A. C.
Torres.
Manifold ways to Darboux-Halphen system.
SIGMA Symmetry Integrability Geom. Methods Appl., 14:Paper No. 003, 14 pages, 2018.
[25]
H. Movasati.
Quasi-modular forms attached to elliptic curves, I.
Ann. Math. Blaise Pascal, 19(2):307–377, 2012.
[26]
H. Movasati.
Gauss-Manin connection in disguise: Calabi-Yau modular
forms, volume 13 of Surveys of Modern Mathematics.
International Press, Somerville, MA; Higher Education Press, Beijing,
2017.
Appendix C by Hossein Movasati and Khosro M. Shokri; Appendix D by
Carlos Matheus.
[27]
T. Nakaya.
A Study on Supersingular Polynomials, Modular Differential
Equations, and Hypergeometric Series.
PhD thesis, 2018.
https://doi.org/10.15017/1931729.
[28]
Y. V. Nesterenko.
Hermite-Padé approximants of generalized hypergeometric
functions.
Russian Acad. Sci. Sb. Math., 83(1):189–219, 1995.
[29]
OEIS Foundation Inc. (2023).
The On-Line Encyclopedia of Integer Sequences.
https://oeis.org.
[30]
F. Pellarin.
On extremal quasi-modular forms after Kaneko and Koike.
Kyushu J. Math., 74(2):401–413, 2020.
With an appendix by Gabriele Nebe.
[31]
F. Pellarin, T. Rivoal, and J.-A. Weil.
Some remarks on classical modular forms and hypergeometric series.
http://rivoal.perso.math.cnrs.fr/articles/modhyp.pdf, 2020.
[32]
Y. Sakai.
The Atkin orthogonal polynomials for the low-level Fricke groups
and their application.
Int. J. Number Theory, 7(6):1637–1661, 2011.
[33]
Y. Sakai.
The Atkin orthogonal polynomials for the Fricke groups of levels
5 and 7.
Int. J. Number Theory, 10(8):2243–2255, 2014.
[34]
Y. Sakai and K. Shimizu.
Modular differential equations with regular singularities at elliptic
points for the Hecke congruence subgroups of low-levels.
Math. J. Okayama Univ., 57:1–12, 2015.
[35]
Y. Sakai and H. Tsutsumi.
Extremal quasimodular forms for low-level congruence subgroups.
J. Number Theory, 132(9):1896–1909, 2012.
[36]
P. F. Stiller.
Classical automorphic forms and hypergeometric functions.
J. Number Theory, 28(2):219–232, 1988.
[37]
H. Tsutsumi.
The Atkin inner product for .
J. Math. Kyoto Univ., 40(4):751–773, 2000.
[38]
H. Tsutsumi.
The Atkin orthogonal polynomials for congruence subgroups of low
levels.
Ramanujan J., 14(2):223–247, 2007.
[39]
R. Vidunas.
Darboux evaluations for hypergeometric functions with the projective
monodromy .
Ramanujan J., 53(1):85–121, 2020.
[40]
D. Zagier.
Modular forms and differential operators.
volume 104, pages 57–75. 1994.
K. G. Ramanathan memorial issue.
[41]
D. Zagier.
Elliptic modular forms and their applications.
In The 1-2-3 of modular forms, Universitext, pages 1–103.
Springer, Berlin, 2008.
[42]
V. V. Zudilin.
Integrality of power expansions related to hypergeometric series.
Math. Notes, 71(5):604–616, 2002.