On finiteness theorems for sums of generalized polygonal numbers
Abstract.
In this paper, we consider mixed sums of generalized polygonal numbers. Specifically, we obtain a finiteness condition for universality of such sums; this means that it suffices to check representability of a finite subset of the positive integers in order to conclude that the sum of generalized polygonal numbers represents every positive integer. The sub-class of sums of generalized polygonal numbers which we consider is those sums of -gonal numbers for which and we obtain a bound on the asymptotic growth of a constant such that it suffices to check the representability condition for .
Key words and phrases:
Shifted lattices, sums of polygonal numbers, theta series2020 Mathematics Subject Classification:
11E25, 11F27, 11F301. Introduction
A milestone in the arithmetic theory of quadratic forms is the Conway–Schneeberger -theorem [5], which states that a positive-definite integral quadratic form represents every positive integer if and only if it represents every positive integer up to . Besides its elegant statement, it provides a satisfactory classification of universal positive-definite integral quadratic forms, which are quadratic forms representing every positive integer. A theorem of this type is usually called a finiteness theorem.
Motivated by this theorem, it is natural to generalize the finiteness result to quadratic polynomials. In particular, we are interested in sums of generalized polygonal numbers. For an integer and , we define the -th generalized -gonal number to be
A sum of generalized polygonal numbers is a function of the form
with integer parameters and . If for some integer , the sum is called a sum of -gonal numbers. We say that a positive integer is represented by a sum , if there exist such that . Such a tuple is called a representation of by . We denote the number of representations of by . If every positive integer is represented by , we say that is universal.
A number of studies are devoted to prove finiteness theorems for sums of -gonal numbers. Indeed, by constructing Bhargava’s escalator tree [1], one can prove that for any integer , there exists a minimal constant such that every sum of -gonal numbers is universal if and only if it represents every positive integers up to [14, Lemma 2.1]. For small values of the parameter , the constant has been determined. To name a few, Bosma and the first author [3] showed that . The next case is a consequence of Conway–Schneeberger’s -theorem. For , Ju [9] showed that . For , Kamaraj, the first author, and Tomiyasu provided an upper bound for in [12]. For , Ju and Oh [10] proved that . In a recent paper, the first author and Liu [14, Theorem 1.1(1)] studied the growth of as . More precisely, they proved that for any real number , we have
(1.1) |
where the implied constant is effective. Later, Kim and Park [15] improved the upper bound by showing that the growth of is at most linear, giving lower and upper bounds for that are both linear, although the constant or proportionality for the upper bound is not explicitly known.
Generally, for a class of sums of generalized polygonal numbers, one can construct the corresponding escalator tree and try to prove a finiteness theorem with a constant depending on the class . For example, the classes considered above are those consisting of sums of -gonal numbers for a fixed integer . The most general class one could consider is the class containing all sums of generalized polygonal numbers with arbitrary parameters . However, such a uniform constant does not exist because by (1.1) we have an arbitrarily large lower bound .
It is thus natural to consider subclasses of . For any integer , we consider the class consisting of sums of generalized polygonal numbers with parameters such that . The reason we consider rather than an a priori more natural choice will become apparent later on. For these classes, we have the following finiteness theorems.
Theorem 1.1.
For any integer , there exists a minimal constant such that every sum of generalized polygonal numbers in is universal if and only if it represents every positive integer up to . Moreover, for any real number , we have
where the implied constant is ineffective.
Remark 1.2.
With the current technology, the ineffectiveness is inevitable because there are finitely many ternary sums appearing in the escalator tree that are highly likely to be universal sums but we have no idea to prove the universality of all of them. The technical difficulty was extensively summarized in [20]. There are also potentially universal quaternary sums in the escalator tree. They are easier to deal with but we do not work out the details for these because the ineffectiveness would remain due to the lack of classification of universal ternary sums. Furthermore, it is worth noting that an additional difficulty when dealing with quaternary sums is the existence of infinitely many universal quaternary sums in the escalator tree. For example, assuming the truth of various generalized Riemann hypotheses, the first author proved that every positive integer except is represented by in [13, Theorem 1.8]. Therefore any child of in the escalator tree is universal. For details, see the construction of the escalator tree below.
Remark 1.3.
Since is exactly the class of sums of triangular numbers, we know that . The class is the class of mixed sums of triangular numbers and squares. Because ternary sums in were classified in [31, 7, 19], the constant is effective in principle. The second author is going to make it explicit in a forthcoming paper. The constant is ineffective for any integer .
Remark 1.2 leads to the following conditional result.
Theorem 1.4.
Assuming that the classification of universal ternary sums has been completed, the implied constants in Theorem 1.1 can be made effective.
In fact, a conjectural list of universal ternary sums can be determined via a computer search, where sums involving are excluded because and represent the same set of integers. Although it seems impossible to prove the universality of all of them, there are ternary sums among them that are confirmed to be universal, by a non-exhaustive search of relevant literature and by algebraic methods. The confirmed cases are listed in Table LABEL:tbl::confirmedcases and the complete data of the conjectural list can be found in the thesis of the second author.
: | [17] | : | [31] | : | [32] | : | [32] | : | [32] |
: | [32] | : | [11] | : | [17] | : | [31] | ||
: | [32] | : | [17] | : | [31] | ||||
: | [32] | : | [17] | : | [31] | : | [32] | : | [32] |
: | [32] | : | [11] | : | [32] | : | [32] | : | [32] |
: | [31] | : | [31] | : | [32] | : | [11] | : | [32] |
: | [7] | : | [7] | : | [31] | : | [31] | : | [7] |
: | [19] | : | [31] | : | [32] | : | [11] | : | [32] |
: | [32] | : | [6] | : | [11] | : | [32] | ||
: | [32] | : | [32] | : | [11] | ||||
: | [32] | : | [32] | : | [32] | : | [32] | ||
: | [32] | : | [11] | : | [32] | : | [11] | : | [32] |
: | [11] | : | [11] | : | [11] | ||||
: | [32] | : | [32] | : | [32] | ||||
: | [32] | : | [11] | : | [32] | : | [11] | ||
: | [17] | : | [31] | : | [32] | : | [11] | : | [17] |
: | [7] | : | [17] | : | [7] | : | [32] | : | [31] |
: | [32] | : | [31] | : | [11] | ||||
: | [32] | : | [32] | : | [11] | ||||
: | [32] | : | [32] | : | [32] | : | [31] | ||
: | [9] | : | [32] | : | [32] | ||||
: | [32] | ||||||||
: | [32] | ||||||||
: | [9] | : | [31] | : | [31] | ||||
: | [31] | : | [31] | ||||||
: | [31] | : | [31] | ||||||
: | [8] | : | [32] | ||||||
: | [32] | : | [6] | ||||||
: | [32] | : | [32] | : | [18] | : | [18] | : | [18] |
: | [18] | : | [6] | ||||||
: | [11] | ||||||||
: | [32] | : | [32] | ||||||
: | [32] | : | [32] | ||||||
: | [6] | : | [32] | : | [6] | : | [18] | ||
: | [32] | ||||||||
: | [6] | : | [6] | : | [32] | : | [18] | ||
: | [18] | : | [6] | : | [9] | ||||
: | [32] | : | [9] | ||||||
Theorem 1.5.
Excluding sums involving generalized hexagonal numbers, there are at most ternary universal sums. Among them, there are ternary sums that are confirmed to be universal, given in Table LABEL:tbl::confirmedcases.
We can avoid the existence of universal ternary and quaternary sums to obtain effective results, as well as better growth by restricting to subclasses of . For any integers such that , we define the subclass as the class of sums of generalized polygonal numbers with parameters such that and . For these classes, we can establish effective finiteness theorems.
Theorem 1.6.
For any integers such that , there exists a minimal constant such that every sum of generalized polygonal numbers in is universal if and only if it represents every positive integer up to . For any real number , we have the following upper bounds for the constant :
-
(1)
If , then we have
-
(2)
If , then we have
For each case, the implied constant is effective.
The rest of the paper is organized as follows. We begin by constructing the escalator tree and proving elementary properties of the escalator tree in Section 2. To obtain the existence and the growth of the constants and , we have to study the representations of integers by nodes. In Section 3, we convert the problem into the study of the number of representations by a shifted lattice, which splits as a sum of the Fourier coefficients of an Eisenstein series and a cusp form. In Section 4 and Section 5, we give estimates on the Eisenstein part and the cuspidal part, respectively. Finally, we prove the main theorem.
2. Elementary Properties of the Escalator Tree
We first construct the escalator tree for a given class of sums of generalized polygonal numbers. The escalator tree for universal sums of generalized polygonal numbers in is a rooted tree constructed inductively as follows. We define the root to be with depth and then inductively construct the nodes of depth from the nodes of depth as follows. If a node of depth is universal, then it is a leaf of the tree. If a sum of generalized polygonal numbers is not universal, then we call the smallest positive integer not represented by the truant of ; for ease of notation, we write if is universal. The children of a node with are the sums of generalized polygonal numbers
with , , and an additional restriction that if to avoid repeated nodes. For any class , we define the set
By construction, this subset has the property that if a sum represents every integer in , then is universal. We also see that is minimal in the sense that it is the smallest subset of with this property. Therefore, if is a finite set, we obtain a finiteness theorem by taking to be the maximal integer contained in .
We denote the escalator tree for the class of arbitrary sums of generalized polygonal numbers. For any integer , we abbreviate for the tree . We will show in Theorem 6.1 that is a finite tree. Therefore, the existence of the constant in Theorem 1.1 is established. Similarly, we abbreviate for the tree for any integers such that . Then the existence of the constant in Theorem 1.6 follows immediately because . To study the growth of the constants and , it is clear that we have to study the truants of non-leaf nodes in the escalator tree .
We begin with an observation on the truants of sums.
Lemma 2.1.
Fix . Let be a sum with . For any integer , the sum has truant , where is a constant depending on with and with .
Proof.
If for some , then there exists such that
Since cannot represent , we see that . Noticing that , we have . Therefore, we have
which is a contradiction. This shows that . For the same reason, the representations of integers are independent of the choice of . Thus, the truant is a constant for any integer . ∎
As a consequence, we obtain the following useful enumeration of the nodes in of depth .
Proposition 2.2.
Suppose that is a node in . Then, and . For each choice of , the truants are summarized in Table 2.1.
Proof.
3 | 4 | 5 | 7 | 8 | 3 | 4 | 5 | 7 | 8 | 9 | 5 | |||
3 | 5 | 8 | 9 | 9 | 12 | 5 | 4 | 5 | 10 | 5 | 4 | 4 | 4 | 6 |
4 | 8 | 3 | 20 | 3 | 3 | 3 | 4 | 5 | 6 | 5 | 4 | 4 | 4 | 6 |
5 | 9 | 20 | 11 | 10 | 4 | 4 | 9 | 7 | 8 | 12 | 6 | 7 | 6 | 9 |
7 | 9 | 3 | 10 | 3 | 3 | 3 | 4 | 5 | 6 | 5 | 4 | 4 | 4 | 6 |
8 | 12 | 3 | 4 | 3 | 3 | 3 | 4 | 5 | 6 | 5 | 4 | 4 | 4 | 6 |
5 | 3 | 4 | 3 | 3 | 3 | 4 | 5 | 6 | 5 | 4 | 4 | 4 | 6 |
The following lemma is a generalization of Lemma 2.1, which is used to bound the truants of sums when varying multiple parameters.
Lemma 2.3.
Fix and parameters . Suppose that there exist integers and such that for any with the property: For any and for any integers for , there exists a non-universal sum with such that for any . Then any sum of the form with for any and for some has the truant
Proof.
Let be any sum of the form with for any and for some . Let be the number of indices such that . By induction on the integer , we prove the lemma jointly with the following inequality
for any . If , without loss of generality, we assume that . By the assumption of the lemma, there exists a sum such that and for any . Using Lemma 2.1, we see that and it follows that for any , we have
as desired. Next we prove the induction step. Assuming the lemma and the inequality hold for , we prove them for . Without loss of generality, we may assume that . Then we choose any sum with . By the inductive hypothesis, we have and
for any . Again by Lemma 2.1, we can conclude that and the sum satisfies the inequality for any . This finishes the proof. ∎
Next we prove upper bounds on the truants of non-leaf nodes of depth and depth .
Proposition 2.4.
Excluding a set of nodes of depth with and , any other node of depth has truant . In particular, a ternary sum is universal only if it is among the excluded nodes.
Proof.
With , we search for the integers satisfying the assumptions of Lemma 2.3 for each choice of integers . By a computer program, we can verify that the integers and satisfy the assumptions for any possible choice of integers of nodes in the escalator tree . By a computer program to check the representations by any node of depth with parameters for any up to , we find that the truant of any node is bounded above by except nodes representing every positive integer up to . Hence, the proposition follows from Lemma 2.3. ∎
Remark 2.5.
Since we can not verify the universality of all of the excluded nodes, we only have an implicit upper bound on the truants of these nodes and the truants of their children by Lemma 2.1, which causes the ineffectiveness of the finiteness theorems.
Proposition 2.6.
For the nodes of depth , we have the following facts:
-
(1)
Suppose that for at least three indices . The truant of a non-leaf node is bounded above by an implicit constant.
-
(2)
Suppose that for at most two indices . Excluding a set of finitely many potentially universal nodes, any other node is not universal and the truant is bounded above by .
Proof.
(1) Set . For any non-leaf node of depth with for at least three indices in , the truant is bounded by an implicit constant by applying Lemma 2.1, including the potential children of excluded nodes of depth .
(2) For any node of depth with for at most two indices in , say , we can use a computer program to verify that the integers for and for satisfy the assumptions of Lemma 2.3. Thus, by a computer program, we see that the truant of any node with for at most two indices is bounded above by , excluding a set of finitely many potentially universal nodes. ∎
Remark 2.7.
From the example in Remark 1.2, in fact we exclude an infinite set of universal nodes with for exactly three indices.
For dealing with nodes of higher depth, the naive enumeration is not practical. Thus we use analytic methods instead. To finish this section, we prove some properties that will be used later.
Lemma 2.8.
Fix a prime number . Suppose that , , and satisfying either or . If the polynomial represents every class in , then
Proof.
First assume that . Replacing by for , we see that represents every class in by the assumption. In particular, there exist such that . Hence, we have such that either or . Without loss of generality, assume that . Then, we see that , which is the desired result.
Next assume that and . Without loss of generality, we may assume that . Replacing by for , we see that represents every class in by the assumption. Thus there exist such that with such that is a non-square modulo . Hence we see that such that either or , and the proof follows as in the previous case. ∎
Lemma 2.9.
Suppose that is a prime number. Let be a node in such that . If one of the following conditions holds
-
(1)
,
-
(2)
, and ,
then we have
Proof.
If , then the construction of implies that . Hence in both cases and thus represents every class in . The claim then follows by Lemma 2.8. ∎
Lemma 2.10.
Let be a node in the tree with . The following facts hold.
-
(1)
If and , then we have . If , then either or .
-
(2)
If and , then we have . If , then or .
-
(3)
If and , then we have . If , then .
-
(4)
If , then we have if , if , if , and if .
Proof.
This follows from a straightforward calculation that uses the fact that either or . ∎
3. Representations by Shifted Lattices
For an integer , a quadratic space of dimension is a vector space over of dimension equipped with a symmetric bilinear form . We can associate with a quadratic map defined by for any vector . A quadratic space is positive-definite if for any non-zero vector . A lattice in a quadratic space of dimension is a free -submodule of of rank . A lattice is integral if . Set . For an integral lattice , the discriminant is defined as and the level is defined as the least positive integer such that for any vector .
A shifted lattice in a quadratic space is a subset of of the form , where is a lattice in and is a vector of . A shifted lattice is positive-definite if the underlying quadratic space is positive-definite. A shifted lattice is integral if . The rank of a shifted lattice is the rank of as a free -module. For an integral shifted lattice , a base lattice of a shifted lattice is an integral lattice containing . An integral shifted lattice always admits a base lattice because the lattice generated by is a base lattice of . The discriminant, the level, and the conductor of an integral shifted lattice relative to a base lattice of are defined as the discriminant of , the level of , and the least positive integer such that . For a shifted lattice of conductor and with , we let be the shifted lattice . Finally, we say that an element is represented by a shifted lattice or equivalently the shifted lattice represents an element if there is a vector such that . Such a vector is called a representation of by . The number of representations of by is denoted by .
Let be a sum of generalized polygonal numbers. In Section 4, we will construct a shifted lattice together with two integers such that for any integer . Therefore, to study the representations by , it is equivalent to study the representations by the shifted lattice . Let be the complex upper half-plane. For a positive-definite integral shifted lattice , the theta series associated to is defined as
By [23, Proposition 2.1] and the theory of modular forms, the theta series is a modular form, which splits into a sum of an Eisenstein series and a cusp form . Let and denote the -th Fourier coefficient of and for any integer , respectively. We have
(3.1) |
By estimating the Fourier coefficients and , we can show that is positive.
To estimate the Eisenstein part, we apply Siegel’s analytic theory of quadratic forms. Note that historically Siegel’s papers [25, 26, 27, 28, 29] did not cover the cases of positive-definite shifted lattices. However, Siegel’s results were later generalized to include positive-definite shifted lattices, for example, see [33, 35, 24].
The main tool is the Siegel–Minkowski formula, which interprets the Fourier coefficient of the Eisenstein part in terms of a product of local densities. We shall use the formulation given in [24, (1.15)]. Suppose that the base lattice of the shifted lattice is denoted by and the quadratic map is denoted by . For , we have
(3.2) |
where is the usual Gamma function and we have if and ; the quantities are the local densities of , which is defined as follows. Let , , and be the localizations of the quadratic space , the base lattice , and the shifted lattice at , respectively. We choose the unique Haar measures on and on such that
For any positive integer and any prime number , the local density of a shifted lattice in the quadratic space is defined as the integral
where the function is defined as follows. For any -adic number , we define for some rational number such that . To estimate the Fourier coefficient of the Eisenstein using (3.2), it is clear that we have to evaluate the local densities, which is the main goal in the coming section.
4. Formulae for local densities and a lower bound on the Eisenstein part
In this section, we derive explicit formulae for local densities by similar arguments in [36, Section 2] and apply them to bound the Eisenstein part.
4.1. An Explicit Formula for Non-dyadic Local Densities
Fix a prime number . Suppose that we have an integral shifted lattice in a quadratic space with the associated quadratic map denoted by and choose a base lattice . By Jordan canonical form theorem [16, Theorem 5.2.4 and Section 5.3], there exists a basis of the localization such that the Gram matrix of is a diagonal matrix with entries and with rational numbers . Then, we have
(4.1) | ||||
Thus, it boils down to evaluating the integral
where and with for .
First, we evaluate the integral
for and .
Lemma 4.1.
Set . Suppose that such that and . Then we have
where for with and , we define
and
Proof.
The first case is obvious. For the second case, replacing by in the integral, we have , where the extra factors containing disappear because for . Since , the integral vanishes. For the last case, replacing by in the integral, we have
Applying [36, Lemma 2.1(1)], we obtain the desired result. ∎
For the sake of simplicty, we define to be a formal symbol with the properties and for any integer . We set a convention that taking the minimum among an empty set outputs the formal symbol .
Theorem 4.2.
Let be an odd prime number. Suppose that and with for . For , we define . We set
and set . We further define
If , we assume that with and . Otherwise we set . For an integer , we define
Then, we have
where for any integer , we define
with and define
4.2. An Explicit Formula for Dyadic Local Densities
The dyadic case is similar but slightly more complicated because the Jordan canonical form is not necessarily diagonal in general. Although a formula for diagonal cases is sufficient for our applications, we shall prove a formula in general for the completeness of the result. By Jordan canonical form theorem [16, Theorem 5.2.5 and Section 5.3], there exists a basis of the localization such that with rational numbers and the Gram matrix of is a block-diagonal matrix with blocks with entries and blocks
with for and satisfying . Arguing in the same way as in the non-dyadic cases, we have
(4.2) |
and it boils down to the calculation of the following integral
where , , , and is a polynomial of the form
First, we evaluate the following Gauss integrals
for and .
Lemma 4.3.
Set . Suppose that such that and . Then we have
where and we extend the Legendre symbol to via the Hilbert symbol by
Proof.
Lemma 4.4.
Set . Suppose that such that and . Then we have
Proof.
The first case is obvious. For the second case, by the symmetry, we may assume that . Therefore, we have
The indicator function vanishes identically because and . Therefore we obtain the desired result. For the third case, we apply the change of the variable to the the second line of the above equation,
as desired. ∎
Lemma 4.5.
Set . Suppose that such that and . Then we have
Proof.
The first case is obvious. For the second case, we may assume that . Therefore by Lemma 4.3, we have
The indicator function vanishes identically because . So we obtain the desired results. For the third case, we apply the changes of variables and . Then we have
where the last equality follows from [36, Lemma 4.4]. ∎
Theorem 4.6.
Suppose that and is a polynomial of the form
For , we define . For , we define . For , we define . We set
and set
We further set
If , assume that such that and . Otherwise we set . For an integer , we define
and
Then, we have
where for any integer , we define
with , define
and define
4.3. Lower Bounds on the Eisenstein Part
Let be a sum of generalized polygonal numbers. We are going to construct a shifted lattice together with integers such that
(4.3) |
for any integer . The choice of such a shifted lattice with the integers and is not necessarily unique. Throughout this paper, we determine them as follows.
Definition 4.7.
Suppose that is a sum of generalized polygonal numbers. We set
For any , we set
Let be a quadratic space of dimension with a basis such that the Gram matrix with respect to the basis is a diagonal matrix with entries . The shifted lattice corresponding to the sum is defined to be
and the base lattice is defined to be the lattice generated by the basis . It is straightforward to verify that (4.3) holds for the shifted lattice .
In this subsection, we are going to bound the Fourier coefficient of the Eisenstein part in (4.3) by Siegel-Minkowski formula and the explicit formulae for local densities. Plugging the data of the shifted lattice and the base lattice constructed in Definition 4.7 into (4.1) and (4.2), we have
(4.4) |
where the quadratic polynomial is given by
(4.5) |
It is relatively easy to bound the product of the local densities for .
Proposition 4.8.
Suppose that is a sum of generalized polygonal numbers with . Let be the shifted lattice and the integers constructed in Definition 4.7 corresponding to . Then we have
where means for any functions .
Proof.
Suppose that is a prime number such that . Then we can eliminate the linear terms in the quadratic polynomial by applying a linear change of variables. Moreover by Jordan canonical form theorem [16, Theorem 5.2.4 and Section 5.3], we may further assume that
where . Apply Theorem 4.2 to , we have and for every and for every integer . Hence it follows from a straightforward calculation that
provided that . Finally, bounding against special values of the zeta function, we obtain the desired asymptotics. ∎
It remains to bound the local densities for prime numbers such that .
Proposition 4.9.
Fix an odd prime number . Suppose that is a node of depth in the escalator tree . Let be the shifted lattice and the integers constructed in Definition 4.7 corresponding to . For any -adic integer , we have
where .
Proof.
We apply Theorem 4.2 to with defined in (4.5). Following the notation in Theorem 4.2, we have for any . Moreover, we have if and if . Reindexing by a permutation , we may assume that and if for any . Now, we show that one of the following condition
-
(1)
;
-
(2)
, ;
-
(3)
, ;
-
(4)
, ;
-
(5)
, , ;
-
(6)
, , .
Clearly by definition. If , then condition (1) holds. So we may assume that . In this case, we have . Indeed, by Table 2.1, we have , and while if , then and . Therefore implies that , which is a contradiction. Hence we have , and then and for any . If , then condition (2) holds. So we may assume that and in particular . In this case, Table 2.1 implies that and . Assuming that , then we have and . If , then condition (3) holds, while if , then either condition (3) or condition (4) holds. Now assume that and . Then and by Lemma 2.9, we see that condition (5) holds. So we have exhausted all of these cases with . Assuming that , we see that condition (6) holds by Lemma 2.9.
Next we bound using Theorem 4.2. If condition (1) holds, then we have . If condition (2) holds, we see that for and therefore for any integer . Thus, we have
Suppose that condition (3) holds next. If , then we have . If , then we have .
If condition (4) holds, we have for . It follows that for any integer . There are four different cases we need to consider separately. First, if and , then we have
Second, if and , then the first sum runs over and . Similar arguments yield that . Third, if and , then we have
Lastly, if and , then we have for any integer . Then, we have
If condition (5) holds, then we see that . If , then we have . If , then we have for any integer and we conclude that
Finally we suppose that condition (6) holds. Then we have . If , then we have . If , then we have . If and , then we have . If and , then for and we have
Combining the calculations with (4.4), we obtain the desired results. ∎
Proposition 4.10.
Suppose that is a node of depth in the escalator tree . Let be the shifted lattice and the integers constructed in Definition 4.7 corresponding to . For any -adic integer , we have
where .
Proof.
We apply Theorem 4.6 to with defined in (4.5). Follow the notation in Theorem 4.6. We have for any where
Moreover, we have if , if , and if . Reindexing by a permutation , we may assume that and if for any . Now, we show that one of the following condition:
-
(1)
;
-
(2)
, , , , if then , if then ;
-
(3)
, , ;
-
(4)
, , ;
-
(5)
, , ;
-
(6)
, , if then , for , ;
-
(7)
, , if then , if then , for , .
If , then condition (1) holds. Suppose that next. From this assumption, we have and . Moreover, we have from Table 2.1. Then it follows that . We first deal with the case . Since , we have , depending on whether there exists with odd, in which case . If , we can conclude that . Since by Table 2.1, when . If , we have by Lemma 2.10. Thus, we have , , and and if then . Thus, condition (2) holds. If , then . In this case, condition (2) also holds.
If , then and by our choice of . If , then condition (3) holds. If , we shall show that condition (4) or condition (5) holds. Since , we have from Table 2.1 and from our assumption . If , we have , , and . By Lemma 2.10, we see that . So, we have and . Therefore condition (4) holds. If , we have since . It follows that . If , then condition (4) holds. If , we see that and by our choice of . Moreover, by Lemma 2.10, we have . Therefore we have . This shows that condition (5) holds.
Next we deal with the case . Then and . From Table 2.1, we notice that . Thus we have and by our choice of . If , then Lemma 2.10 (1) and (3) implies that , with only possible if and , giving condition (6) in that case. If , then condition (2) holds. If , then condition (6) holds.
Now suppose that . By Lemma 2.10, we have , with only possible if and , giving condition (7) in that case. If , then condition (2) holds. If , then condition (7) holds. Finally, if , then either condition (2) holds or condition (7) holds, depending on whether or .
Then we bound using the explicit formula from Theorem 4.6. If , then . Otherwise, we have , where we define
If condition (1) holds, then . Suppose that condition (2) holds. If , then . If and , we have for and for . If , then
If , then for . We have
Similarly if and then and . Then, we see that,
If condition (3) holds, we have . If condition (4) or condition (5) holds, arguing as the case of condition (2), we see that
Suppose that condition (6) holds. If with , we have . If with or with , then
where if and for some and otherwise. Notice that the right hand side is determined by , and . By a computer program and by using Lemma 2.10 to get rid of certain choices of , we can conclude that . If with or with , we can argue in a similar manner to conclude that . If , we have for any integer . Thus, for , we have
Combining with the calculation of the cases , we have
The case of condition (7) is similar to that of condition (6). Using Lemma 2.10, we can prove that . Combining the calculations with (4.4), we obtain the desired results. ∎
Now we can prove a lower bound on the Eisenstein part.
Proposition 4.11.
Suppose that is a node of depth in the escalator tree . Let be the shifted lattice and the integers constructed in Definition 4.7 corresponding to . For any real number , we have
for any integer , where .
Proof.
By our construction of the base lattice in Definition 4.7, we have
Plugging this upper bound and Proposition 4.8 into (3.2), we have
Using Proposition 4.9 and Proposition 4.10 to bound the remaining local densities and using Robin’s bound [21, Theorem 11] on the divisor function, it is not hard to obtain the desired bound. ∎
5. Upper Bounds on the Cuspidal Part
In this section, we prove upper bounds on the Fourier coefficient of the cuspidal part appearing in (4.3). This requires the theory of modular forms. We will use the following congruence subgroups of :
and
Fix . Assume that is a congruence subgroup such that when and is a Dirichlet character. Letting be the standard unary theta function of weight , we say that a function is a (holomorphic) modular form of weight and Nebentypus character (for the -multiplier) on if for every we have
and grows at most polynomially in as ; we furthermore call a cusp form if vanishes as for all . We let denote the space of modular forms of weight for the congruence subgroup with the Nebentypus and let denote the subspace of cusp forms in .
In order to tie together the theory of modular forms and the algebraic theory of shifted lattices, we require the Siegel–Weil formula, which rewrites the Fourier coefficients of the Eisenstein series as the number of representations by the genus. To be precise, suppose that we have a shifted lattice in a quadratic space . Let denote the orthogonal group of , the adelization of , and the subgroup of consisting of isometries such that , respectively. By [30, Lemma 1.2], and act on the set of shifted lattices in a quadratic space . The genus of a shifted lattice is the orbit of under the action of , denoted by . The isometry class of a shifted lattice is the orbit of under the action of , denoted by . By [30, Corollary 2.3], the genus of splits into a finite number of isometry classes. So we take a set of representatives of isometry classes in . The Siegel–Weil formula states that the Fourier coefficient is a weighted sum of the number of representations by shifted lattices in as follows,
where the quantity is the mass of , defined by
Let be a positive-definite integral shifted lattice of rank , level , and conductor . By [23, Proposition 2.1], for any shifted lattice in the genus of , the theta series is a modular form in for some Dirichlet character . Therefore we see that .
We decompose the cuspidal part by [4, Theorem 2.5] as follows,
(5.1) |
where the sum runs over all Dirichlet character modulo . For each Dirichlet character modulo , we have . We call the -component of the shifted lattice . Thus, we reduce to the cases when the cusp forms are on the congruence subgroup .
We shall use the following approach to give upper bounds on the Fourier coefficients of the cusp forms. We can normalize a cusp form with respect to the Petersson norm, which is induced by the Petersson inner product. For two modular forms such that is a cusp form, the Petersson inner product is defined as
where we write with . For a cusp form , the Petersson norm of is defined as
We note that while depends on the choice of , the normalization is independent of the choice of ; both normalizations appear in the literature and prove useful in different settings, so we keep the dependence on in our notation to avoid confusion with the normalization that is independent of .
The next proposition reduces the problem of bounding the Fourier coefficients of a cusp form to bounding its Petersson norm.
Proposition 5.1.
Fix such that . Let be an integer such that if and be a Dirichlet character modulo . Let be a cusp form in . If , for any real number , we have
for any integer . If , for any real number , we have
for any integer . Moreover, both the implied constants are effective.
By Proposition 5.1, it suffices to bound the Petersson norm of a cusp form. To do so, we adopt the method of Blomer in [2]. In the paper [2], the method was designed for lattices. But it can be generalized to the cases of shifted lattices. Here we imitate the proof of [12, Lemma 3.2].
Proposition 5.2.
Let be the -component of the cuspidal part of the theta series associated to an integral positive-definite shifted lattice of rank , discriminant , level , and conductor . For any real number , we have
where the implied constant is effective.
Proof.
Let denote the usual fundamental domain of acting on the upper half-plane . Let be a set of representatives of the cosets . Let denote the parabolic subgroup of generated by , namely
Finally, let be a set of representative of the double cosets .
Set
By definition of the Petersson norm, we have
where and we denote
The cusp associated to the matrix is . Recall that every cusp of has cusp width , so the Fourier expansion of at may be written in the form (for some )
Plugging in the Fourier expansions, we have
where is the incomplte gamma function defined as
for a complex variable and a real variable such that and , and for the last step we use an asympotic upper bound as . Therefore, it remains to bound the absolute value of the Fourier coefficients of at each cusp.
Then, by [4, Theorem 2.5] and Siegel–Weil formula, we see that the -component is constructed explicitly as follows,
where is a matrix whose lower right entry is congruent to modulo . Thus, by the triangle inequality and the linearity of the slash operator, it suffices to bound the Fourier coefficient of the theta series at the cusp for each class .
For any shifted lattice , we are going to estimate the Fourier coefficient . There exists a base lattice of such that is of level and conductor relative to . We choose a basis of such that is a basis of for integers and with integers . Let be the Hessian matrix of the bilinear form with respect to the basis . We need a modular transformation formula for the shifted lattice . Using Shimura’s formulation of theta series [23, (2.0)] with the trivial spherical function , we put
where is a real symmetric matrix, is a vector of dimension , and is a positive integer. The theta series is related to Shimura’s theta series as follows,
It is easy to verify that for Shimura’s theta series , a stronger assumption than that in [23, (2.1)] is satisfied since the matrix is even integral. Under such a stronger assumption, we can imitate the arguments in [23, Pages 454–455] to see that, for any matrix with , we have
where is defined as follows,
By [12, Page 15], we have , where are the diagonal entries of the Smith normal form of the Hessian matrix . Hence, plugging it back to the transformation formula of , we have
where is the -th Fourier coefficient of at the cusp for a matrix . Since the explicit construction of the -component is a weighted average of the theta series , the same bound holds for the Fourier coefficient for every . Finally, plugging it back to the Petersson norm of , we obtain
Taking the square root, we obtain the desired result. ∎
Now we can state the main results of this section.
Theorem 5.3.
Let be the cuspidal part of the theta series associated to a positive-definite integral shifted lattice of rank , level , and conductor . For any real number and any integer , we have the following upper bounds:
-
(1)
If is odd, then we have
-
(2)
If is even, then we have
Here the implied constants are effective and denotes Euler’s totient function.
Applying Theorem 5.3 to the shifted lattice constructed in Definition 4.7, we obtain the following upper bounds.
Proposition 5.4.
Suppose that is a node of depth in the escalator tree . Let be the shifted lattice and the integers constructed in Definition 4.7 corresponding to . Set . For any real number and any integer , the following inequalities hold:
-
(1)
If is odd, then we have
-
(2)
If is even, we have
In both cases, the implied constants are effective.
Proof.
Suppose that the level and the conductor of the shifted lattice are denoted by and , respectively. The conductor is bounded above by and the level is bounded above by . Plugging them into Theorem 5.3, we obtain the desired bounds. ∎
6. Proof of the Main Results
For any integers , the existence of the constants and follows from the following theorem.
Theorem 6.1.
For any integer , the tree is finite.
Proof.
By induction on the depth, we see that the parameters and are bounded for any and there are finitely many nodes of a fixed depth in . It suffices to show that the depth of the tree is finite. Using Proposition 4.11 and Proposition 5.4 to compare the Eisenstein part and the cuspidal part in (4.3), the integers not represented by a node of depth is bounded below by a uniform constant only depending on . Since any child node represents at least one more integer than its parent node, the depth of every subtree rooted at a node of depth must be finite. This implies that the depth of the tree is finite. ∎
Now we prove the main theorems.
Proof of Theorem 1.1 and Theorem 1.4.
The existence of the constant follows from Theorem 6.1. When , the calculation in [3] reveals that . For the upper bound on , by Proposition 2.2, Proposition 2.4, and Proposition 2.6, we can bound the truants of the non-leaf nodes of depth in by an implicit constant. So we arrive at depth . The parameters are also bounded by this implicit constant. Thus, by comparing the bounds in Proposition 4.11 and Proposition 5.4, we see that any node of depth in represents every positive integer such that
This implies that the truant of any non-leaf node of depth in is bounded above by the right hand side. Since any non-leaf node of depth will represent more integers than its ancestor of depth , its truant must satisfy the same bound. This finishes the proof of Theorem 1.1. The proof of Theorem 1.4 is clear from the above proof, because the ineffectiveness is due to the implicit upper bound for the truants of nodes of depth . ∎
Proof of Theorem 1.6.
Since we have , the existence of the constant is clear for any integers such that . If , any node of depth with for can not represent . Thus the constant is effective and it satisfies the same bound in Theorem 1.1 with an effective implied constant under the assumption. If , any node of depth with for can not represent . For nodes of depth , we can use a stronger upper bound for the cuspidal part when is even. Therefore we have
as desired. ∎
Proof of Theorem 1.5.
The first statement follows from Proposition 2.4 and a computer search. For the second statement, there are cases that are proved to be universal. For the reference, see Table LABEL:tbl::confirmedcases. For the remaining cases, it is not hard to show that the genus of its corresponding shifted lattice contains only one class. Hence the cuspidal part is trivial by Siegel-Weil formula and the universality can be verified easily via local computations. ∎
References
- [1] Manjul Bhargava. On the conway-schneeberger fifteen theorem. Contemporary Mathematics, 272:27–38, 2000.
- [2] Valentin Blomer. Uniform bounds for fourier coefficients of theta-series with arithmetic applications. Acta Arithmetica, 114:1–21, 2004.
- [3] Wieb Bosma and Ben Kane. The triangular theorem of eight and representation by quadratic polynomials. Proceedings of the American Mathematical Society, 141(5):1473–1486, 2013.
- [4] Bumkyu Cho. On the number of representations of integers by quadratic forms with congruence conditions. Journal of Mathematical Analysis and Applications, 462(1):999–1013, 2018.
- [5] John H Conway. Universal quadratic forms and the fifteen theorem. Quadratic forms and their applications, page 23, 1999.
- [6] Fan Ge and Zhi-Wei Sun. On some universal sums of generalized polygonal numbers. In Colloquium Mathematicum, volume 145, pages 149–155. Instytut Matematyczny Polskiej Akademii Nauk, 2016.
- [7] Song Guo, Hao Pan, and Zhi-Wei Sun. Mixed sums of squares and triangular numbers (II). Integers, 7(1):Paper–A56, 2007.
- [8] Richard K Guy. Every number is expressible as the sum of how many polygonal numbers? The American Mathematical Monthly, 101(2):169–172, 1994.
- [9] Jangwon Ju. Universal sums of generalized pentagonal numbers. The Ramanujan Journal, 51(3):479–494, 2020.
- [10] Jangwon Ju and Byeong-Kweon Oh. Universal sums of generalized octagonal numbers. Journal of Number Theory, 190:292–302, 2018.
- [11] Jangwon Ju, Byeong-Kweon Oh, and Bangnam Seo. Ternary universal sums of generalized polygonal numbers. International Journal of Number Theory, 15(04):655–675, 2019.
- [12] Ramanujam Kamaraj, Ben Kane, and Ryoko Tomiyasu. Universal sums of generalized heptagonal numbers. arXiv preprint arXiv:2203.14678, 2022.
- [13] Ben Kane. On two conjectures about mixed sums of squares and triangular numbers. J. Combin. Number Theory, 1:77–90, 2009.
- [14] Ben Kane and Jingbo Liu. Universal sums of -gonal numbers. International Mathematics Research Notices, 2020(20):6999–7036, 2020.
- [15] Byeong Moon Kim and Dayoon Park. A finiteness theorem for universal -gonal forms. arXiv preprint arXiv:2011.02650, 2020.
- [16] Yoshiyuki Kitaoka. Arithmetic of quadratic forms, volume 106. Cambridge University Press, 1999.
- [17] J Liouville. Nouveaux théorémes concernant les nombres triangulaires. Journal de Mathématiques pures et appliquées, 8:73–84, 1863.
- [18] Byeong-Kweon Oh. Ternary universal sums of generalized pentagonal numbers. The Korean Mathematical Society, 48(4):837–847, 2011.
- [19] Byeong-Kweon Oh and Zhi-Wei Sun. Mixed sums of squares and triangular numbers (III). Journal of Number Theory, 129(4):964–969, 2009.
- [20] Ken Ono and Kannan Soundararajan. Ramanujan’s ternary quadratic form. Inventiones mathematicae, 130(3):415–454, 1997.
- [21] Guy Robin. Estimation de la fonction de tchebychef sur le k-ième nombre premier et grandes valeurs de la fonction (n) nombre de diviseurs premiers de n. Acta Arithmetica, 42(4):367–389, 1983.
- [22] Rainer Schulze-Pillot and Abdullah Yenirce. Petersson products of bases of spaces of cusp forms and estimates for fourier coefficients. International Journal of Number Theory, 14(08):2277–2290, 2018.
- [23] Goro Shimura. On modular forms of half integral weight. Annals of Mathematics, pages 440–481, 1973.
- [24] Goro Shimura. Inhomogeneous quadratic forms and triangular numbers. American journal of mathematics, 126(1):191–214, 2004.
- [25] Carl Ludwig Siegel. Uber die analytische theorie der quadratischen formen. Annals of Mathematics, pages 527–606, 1935.
- [26] Carl Ludwig Siegel. Uber die analytische theorie der quadratischen formen ii. Annals of mathematics, pages 230–263, 1936.
- [27] Carl Ludwig Siegel. Uber die analytische theorie der quadratischen formen iii. Annals of Mathematics, pages 212–291, 1937.
- [28] Carl Ludwig Siegel. Indefinite quadratische formen und funktionentheorie i. Mathematische Annalen, 124(1):17–54, 1951.
- [29] Carl Ludwig Siegel. Indefinite quadratische formen und funktionentheorie ii. Mathematische Annalen, 124(1):364–387, 1951.
- [30] Liang Sun. Class numbers of quadratic diophantine equations. Journal of Number Theory, 166:181–192, 2016.
- [31] Zhi-Wei Sun. Mixed sums of squares and triangular numbers. Acta Arithmetica, 127:103–113, 2007.
- [32] Zhi-Wei Sun. On universal sums of polygonal numbers. Science China Mathematics, 7(58):1367–1396, 2015.
- [33] F Van der Blij. On the theory of quadratic forms. Annals of Mathematics, pages 875–883, 1949.
- [34] Fabian Waibel. Fourier coefficients of half-integral weight cusp forms and waring’s problem. The Ramanujan Journal, 47(1):185–200, 2018.
- [35] André Weil. Sur la formule de siegel dans la théorie des groupes classiques. Acta mathematica, 113(1):1–87, 1965.
- [36] Tonghai Yang. An explicit formula for local densities of quadratic forms. Journal of Number Theory, 2(72):309–356, 1998.