This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On finiteness theorems for sums of generalized polygonal numbers

Ben Kane Department of Mathematics, University of Hong Kong, Pokfulam, Hong Kong [email protected]  and  Zichen Yang Department of Mathematics, University of Hong Kong, Pokfulam, Hong Kong [email protected]
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 mjm_{j}-gonal numbers for which lcm(m12,,mr2)𝔐\text{lcm}(m_{1}-2,\dots,m_{r}-2)\leq\mathfrak{M} and we obtain a bound on the asymptotic growth of a constant Γ𝔐\Gamma_{\mathfrak{M}} such that it suffices to check the representability condition for nΓ𝔐n\leq\Gamma_{\mathfrak{M}}.

Key words and phrases:
Shifted lattices, sums of polygonal numbers, theta series
2020 Mathematics Subject Classification:
11E25, 11F27, 11F30
The first author was supported by grants from the Research Grants Council of the Hong Kong SAR, China (project numbers HKU 17303618 and 17314122).

1. Introduction

A milestone in the arithmetic theory of quadratic forms is the Conway–Schneeberger 1515-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 1515. 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 m3m\geq 3 and xx\in\mathbb{Z}, we define the xx-th generalized mm-gonal number to be

Pm(x)(m2)x2(m4)x2.P_{m}(x)\coloneqq\frac{(m-2)x^{2}-(m-4)x}{2}.

A sum of generalized polygonal numbers is a function F:rF\colon\mathbb{Z}^{r}\to\mathbb{Z} of the form

F(x1,,xr)=a1Pm1(x1)++arPmr(xr)F(x_{1},\ldots,x_{r})=a_{1}P_{m_{1}}(x_{1})+\cdots+a_{r}P_{m_{r}}(x_{r})

with integer parameters 1a1ar1\leq a_{1}\leq\cdots\leq a_{r} and m1,,mr3m_{1},\ldots,m_{r}\geq 3. If m1==mr=mm_{1}=\cdots=m_{r}=m for some integer m3m\geq 3, the sum FF is called a sum of mm-gonal numbers. We say that a positive integer nn is represented by a sum FF, if there exist x1,,xrx_{1},\ldots,x_{r}\in\mathbb{Z} such that n=F(x1,,xr)n=F(x_{1},\ldots,x_{r}). Such a tuple (x1,,xr)r(x_{1},\ldots,x_{r})\in\mathbb{Z}^{r} is called a representation of nn by FF. We denote rF(n)r_{F}(n) the number of representations of nn by FF. If every positive integer is represented by FF, we say that FF is universal.

A number of studies are devoted to prove finiteness theorems for sums of mm-gonal numbers. Indeed, by constructing Bhargava’s escalator tree [1], one can prove that for any integer m3m\geq 3, there exists a minimal constant γm>0\gamma_{m}>0 such that every sum of mm-gonal numbers is universal if and only if it represents every positive integers up to γm\gamma_{m} [14, Lemma 2.1]. For small values of the parameter mm, the constant γm\gamma_{m} has been determined. To name a few, Bosma and the first author [3] showed that γ3=γ6=8\gamma_{3}=\gamma_{6}=8. The next case γ4=15\gamma_{4}=15 is a consequence of Conway–Schneeberger’s 1515-theorem. For m=5m=5, Ju [9] showed that γ5=109\gamma_{5}=109. For m=7m=7, Kamaraj, the first author, and Tomiyasu provided an upper bound for γ7\gamma_{7} in [12]. For m=8m=8, Ju and Oh [10] proved that γ8=60\gamma_{8}=60. In a recent paper, the first author and Liu [14, Theorem 1.1(1)] studied the growth of γm\gamma_{m} as mm\to\infty. More precisely, they proved that for any real number ε>0\varepsilon>0, we have

(1.1) m4γmεm7+εm-4\leq\gamma_{m}\ll_{\varepsilon}m^{7+\varepsilon}

where the implied constant is effective. Later, Kim and Park [15] improved the upper bound by showing that the growth of γm\gamma_{m} is at most linear, giving lower and upper bounds for γm\gamma_{m} that are both linear, although the constant or proportionality for the upper bound is not explicitly known.

Generally, for a class 𝒞\mathcal{C} of sums of generalized polygonal numbers, one can construct the corresponding escalator tree T𝒞T_{\mathcal{C}} and try to prove a finiteness theorem with a constant Γ𝒞>0\Gamma_{\mathcal{C}}>0 depending on the class 𝒞\mathcal{C}. For example, the classes considered above are those consisting of sums of mm-gonal numbers for a fixed integer m3m\geq 3. The most general class one could consider is the class 𝒞\mathcal{C}_{\infty} containing all sums of generalized polygonal numbers with arbitrary parameters m1,mr3m_{1}\ldots,m_{r}\geq 3. However, such a uniform constant Γ\Gamma_{\infty} does not exist because by (1.1) we have an arbitrarily large lower bound Γγmm4\Gamma_{\infty}\geq\gamma_{m}\geq m-4.

It is thus natural to consider subclasses of 𝒞\mathcal{C}_{\infty}. For any integer 𝔐1\mathfrak{M}\geq 1, we consider the class 𝒞𝔐\mathcal{C}_{\mathfrak{M}} consisting of sums of generalized polygonal numbers with parameters m1,,mr3m_{1},\ldots,m_{r}\geq 3 such that lcm(m12,,mr2)𝔐\text{lcm}(m_{1}-2,\ldots,m_{r}-2)\leq\mathfrak{M}. The reason we consider lcm(m12,,mr2)\text{lcm}(m_{1}-2,\ldots,m_{r}-2) rather than an a priori more natural choice lcm(m1,,mr)\text{lcm}(m_{1},\ldots,m_{r}) will become apparent later on. For these classes, we have the following finiteness theorems.

Theorem 1.1.

For any integer 𝔐1\mathfrak{M}\geq 1, there exists a minimal constant Γ𝔐>0\Gamma_{\mathfrak{M}}>0 such that every sum of generalized polygonal numbers in 𝒞𝔐\mathcal{C}_{\mathfrak{M}} is universal if and only if it represents every positive integer up to Γ𝔐\Gamma_{\mathfrak{M}}. Moreover, for any real number ε>0\varepsilon>0, we have

Γ𝔐ε𝔐43+ε,\Gamma_{\mathfrak{M}}\ll_{\varepsilon}\mathfrak{M}^{43+\varepsilon},

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 2323 is represented by P4+2P4+3P3P_{4}+2P_{4}+3P_{3} in [13, Theorem 1.8]. Therefore any child of P4+2P4+3P3P_{4}+2P_{4}+3P_{3} in the escalator tree is universal. For details, see the construction of the escalator tree below.

Remark 1.3.

Since 𝒞1\mathcal{C}_{1} is exactly the class of sums of triangular numbers, we know that Γ1=8\Gamma_{1}=8. The class 𝒞2\mathcal{C}_{2} is the class of mixed sums of triangular numbers and squares. Because ternary sums in 𝒞2\mathcal{C}_{2} were classified in [31, 7, 19], the constant Γ2\Gamma_{2} is effective in principle. The second author is going to make it explicit in a forthcoming paper. The constant Γ𝔐\Gamma_{\mathfrak{M}} is ineffective for any integer 𝔐3\mathfrak{M}\geq 3.

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 30523052 universal ternary sums can be determined via a computer search, where sums involving P6P_{6} are excluded because P3P_{3} and P6P_{6} represent the same set of integers. Although it seems impossible to prove the universality of all of them, there are 197197 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.

Table 1.1. The list of ternary sums that are confirmed to be universal.
P3+P3+P3P_{3}+P_{3}+P_{3}: [17] P3+P3+P4P_{3}+P_{3}+P_{4}: [31] P3+P3+P5P_{3}+P_{3}+P_{5}: [32] P3+P3+P7P_{3}+P_{3}+P_{7}: [32] P3+P3+P8P_{3}+P_{3}+P_{8}: [32]
P3+P3+P10P_{3}+P_{3}+P_{10}: [32] P3+P3+P11P_{3}+P_{3}+P_{11} P3+P3+P12P_{3}+P_{3}+P_{12}: [11] P3+P3+2P3P_{3}+P_{3}+2P_{3}: [17] P3+P3+2P4P_{3}+P_{3}+2P_{4}: [31]
P3+P3+2P5P_{3}+P_{3}+2P_{5}: [32] P3+P3+2P8P_{3}+P_{3}+2P_{8} P3+P3+3P5P_{3}+P_{3}+3P_{5} P3+P3+4P3P_{3}+P_{3}+4P_{3}: [17] P3+P3+4P4P_{3}+P_{3}+4P_{4}: [31]
P3+P3+4P5P_{3}+P_{3}+4P_{5}: [32] P3+P3+5P3P_{3}+P_{3}+5P_{3}: [17] P3+P4+P4P_{3}+P_{4}+P_{4}: [31] P3+P4+P5P_{3}+P_{4}+P_{5}: [32] P3+P4+P7P_{3}+P_{4}+P_{7}: [32]
P3+P4+P8P_{3}+P_{4}+P_{8}: [32] P3+P4+P9P_{3}+P_{4}+P_{9}: [11] P3+P4+P10P_{3}+P_{4}+P_{10}: [32] P3+P4+P11P_{3}+P_{4}+P_{11}: [32] P3+P4+P12P_{3}+P_{4}+P_{12}: [32]
P3+P4+2P3P_{3}+P_{4}+2P_{3}: [31] P3+P4+2P4P_{3}+P_{4}+2P_{4}: [31] P3+P4+2P5P_{3}+P_{4}+2P_{5}: [32] P3+P4+2P7P_{3}+P_{4}+2P_{7}: [11] P3+P4+2P8P_{3}+P_{4}+2P_{8}: [32]
P3+P4+3P3P_{3}+P_{4}+3P_{3}: [7] P3+P4+3P4P_{3}+P_{4}+3P_{4}: [7] P3+P4+4P3P_{3}+P_{4}+4P_{3}: [31] P3+P4+4P4P_{3}+P_{4}+4P_{4}: [31] P3+P4+6P3P_{3}+P_{4}+6P_{3}: [7]
P3+P4+8P3P_{3}+P_{4}+8P_{3}: [19] P3+P4+8P4P_{3}+P_{4}+8P_{4}: [31] P3+P5+P5P_{3}+P_{5}+P_{5}: [32] P3+P5+P7P_{3}+P_{5}+P_{7}: [11] P3+P5+P8P_{3}+P_{5}+P_{8}: [32]
P3+P5+P9P_{3}+P_{5}+P_{9}: [32] P3+P5+P11P_{3}+P_{5}+P_{11}: [6] P3+P5+P12P_{3}+P_{5}+P_{12} P3+P5+P13P_{3}+P_{5}+P_{13}: [11] P3+P5+2P3P_{3}+P_{5}+2P_{3}: [32]
P3+P5+2P4P_{3}+P_{5}+2P_{4}: [32] P3+P5+2P5P_{3}+P_{5}+2P_{5} P3+P5+2P7P_{3}+P_{5}+2P_{7}: [32] P3+P5+2P8P_{3}+P_{5}+2P_{8} P3+P5+2P9P_{3}+P_{5}+2P_{9}: [11]
P3+P5+3P3P_{3}+P_{5}+3P_{3}: [32] P3+P5+3P4P_{3}+P_{5}+3P_{4}: [32] P3+P5+4P3P_{3}+P_{5}+4P_{3}: [32] P3+P5+4P4P_{3}+P_{5}+4P_{4}: [32] P3+P5+4P5P_{3}+P_{5}+4P_{5}
P3+P5+6P3P_{3}+P_{5}+6P_{3}: [32] P3+P5+9P3P_{3}+P_{5}+9P_{3}: [11] P3+P7+P7P_{3}+P_{7}+P_{7}: [32] P3+P7+P8P_{3}+P_{7}+P_{8}: [11] P3+P7+P10P_{3}+P_{7}+P_{10}: [32]
P3+P7+2P3P_{3}+P_{7}+2P_{3}: [11] P3+P7+2P5P_{3}+P_{7}+2P_{5}: [11] P3+P7+2P7P_{3}+P_{7}+2P_{7}: [11] P3+P7+5P3P_{3}+P_{7}+5P_{3} P3+P8+P8P_{3}+P_{8}+P_{8}
P3+P8+2P3P_{3}+P_{8}+2P_{3}: [32] P3+P8+2P4P_{3}+P_{8}+2P_{4}: [32] P3+P8+2P5P_{3}+P_{8}+2P_{5} P3+P8+3P3P_{3}+P_{8}+3P_{3} P3+P8+3P4P_{3}+P_{8}+3P_{4}: [32]
P3+P8+4P3P_{3}+P_{8}+4P_{3} P3+P8+6P3P_{3}+P_{8}+6P_{3}: [32] P3+P9+2P3P_{3}+P_{9}+2P_{3}: [11] P3+P10+2P3P_{3}+P_{10}+2P_{3}: [32] P3+P12+2P3P_{3}+P_{12}+2P_{3}: [11]
P3+2P3+2P3P_{3}+2P_{3}+2P_{3}: [17] P3+2P3+2P4P_{3}+2P_{3}+2P_{4}: [31] P3+2P3+2P7P_{3}+2P_{3}+2P_{7}: [32] P3+2P3+2P8P_{3}+2P_{3}+2P_{8}: [11] P3+2P3+3P3P_{3}+2P_{3}+3P_{3}: [17]
P3+2P3+3P4P_{3}+2P_{3}+3P_{4}: [7] P3+2P3+4P3P_{3}+2P_{3}+4P_{3}: [17] P3+2P3+4P4P_{3}+2P_{3}+4P_{4}: [7] P3+2P3+4P5P_{3}+2P_{3}+4P_{5}: [32] P3+2P4+2P4P_{3}+2P_{4}+2P_{4}: [31]
P3+2P4+2P5P_{3}+2P_{4}+2P_{5}: [32] P3+2P4+4P3P_{3}+2P_{4}+4P_{3}: [31] P3+2P4+4P5P_{3}+2P_{4}+4P_{5}: [11] P3+2P5+2P5P_{3}+2P_{5}+2P_{5} P3+2P5+3P3P_{3}+2P_{5}+3P_{3}
P3+2P5+3P4P_{3}+2P_{5}+3P_{4}: [32] P3+2P5+4P3P_{3}+2P_{5}+4P_{3}: [32] P3+2P5+4P4P_{3}+2P_{5}+4P_{4}: [11] P3+2P5+6P3P_{3}+2P_{5}+6P_{3} P3+2P8+4P5P_{3}+2P_{8}+4P_{5}
P4+P4+P5P_{4}+P_{4}+P_{5}: [32] P4+P4+P8P_{4}+P_{4}+P_{8}: [32] P4+P4+P10P_{4}+P_{4}+P_{10}: [32] P4+P4+2P3P_{4}+P_{4}+2P_{3}: [31] P4+P4+2P5P_{4}+P_{4}+2P_{5}
P4+P4+2P8P_{4}+P_{4}+2P_{8}: [9] P4+P5+P5P_{4}+P_{5}+P_{5}: [32] P4+P5+P10P_{4}+P_{5}+P_{10} P4+P5+2P3P_{4}+P_{5}+2P_{3} P4+P5+2P4P_{4}+P_{5}+2P_{4}: [32]
P4+P5+2P5P_{4}+P_{5}+2P_{5} P4+P5+2P8P_{4}+P_{5}+2P_{8} P4+P5+3P3P_{4}+P_{5}+3P_{3}: [32] P4+P5+4P5P_{4}+P_{5}+4P_{5} P4+P5+6P3P_{4}+P_{5}+6P_{3}
P4+P7+2P3P_{4}+P_{7}+2P_{3} P4+P7+2P5P_{4}+P_{7}+2P_{5} P4+P8+P8P_{4}+P_{8}+P_{8}: [32] P4+P8+2P3P_{4}+P_{8}+2P_{3} P4+P8+2P5P_{4}+P_{8}+2P_{5}
P4+P8+3P4P_{4}+P_{8}+3P_{4}: [9] P4+P10+2P3P_{4}+P_{10}+2P_{3} P4+P12+2P4P_{4}+P_{12}+2P_{4} P4+2P3+2P3P_{4}+2P_{3}+2P_{3}: [31] P4+2P3+2P4P_{4}+2P_{3}+2P_{4}: [31]
P4+2P3+2P5P_{4}+2P_{3}+2P_{5} P4+2P3+2P8P_{4}+2P_{3}+2P_{8} P4+2P3+4P3P_{4}+2P_{3}+4P_{3}: [31] P4+2P3+4P4P_{4}+2P_{3}+4P_{4}: [31] P4+2P3+4P5P_{4}+2P_{3}+4P_{5}
P4+2P3+5P3P_{4}+2P_{3}+5P_{3}: [31] P4+2P4+4P3P_{4}+2P_{4}+4P_{3}: [31] P4+2P5+2P5P_{4}+2P_{5}+2P_{5} P4+2P5+3P3P_{4}+2P_{5}+3P_{3} P4+2P5+6P3P_{4}+2P_{5}+6P_{3}
P4+2P8+4P5P_{4}+2P_{8}+4P_{5} P5+P5+P5P_{5}+P_{5}+P_{5}: [8] P5+P5+P10P_{5}+P_{5}+P_{10}: [32] P5+P5+2P3P_{5}+P_{5}+2P_{3} P5+P5+2P4P_{5}+P_{5}+2P_{4}
P5+P5+2P5P_{5}+P_{5}+2P_{5}: [32] P5+P5+2P8P_{5}+P_{5}+2P_{8} P5+P5+3P3P_{5}+P_{5}+3P_{3} P5+P5+3P5P_{5}+P_{5}+3P_{5}: [6] P5+P5+4P3P_{5}+P_{5}+4P_{3}
P5+P5+4P5P_{5}+P_{5}+4P_{5}: [32] P5+P5+5P5P_{5}+P_{5}+5P_{5}: [32] P5+P5+6P5P_{5}+P_{5}+6P_{5}: [18] P5+P5+8P5P_{5}+P_{5}+8P_{5}: [18] P5+P5+9P5P_{5}+P_{5}+9P_{5}: [18]
P5+P5+10P5P_{5}+P_{5}+10P_{5}: [18] P5+P7+3P3P_{5}+P_{7}+3P_{3}: [6] P5+P8+2P3P_{5}+P_{8}+2P_{3} P5+P8+2P5P_{5}+P_{8}+2P_{5} P5+P8+3P3P_{5}+P_{8}+3P_{3}
P5+P8+3P4P_{5}+P_{8}+3P_{4} P5+P9+2P3P_{5}+P_{9}+2P_{3}: [11] P5+P9+3P3P_{5}+P_{9}+3P_{3} P5+P10+2P3P_{5}+P_{10}+2P_{3} P5+P10+3P3P_{5}+P_{10}+3P_{3}
P5+2P3+2P3P_{5}+2P_{3}+2P_{3} P5+2P3+2P8P_{5}+2P_{3}+2P_{8} P5+2P3+3P3P_{5}+2P_{3}+3P_{3}: [32] P5+2P3+3P4P_{5}+2P_{3}+3P_{4}: [32] P5+2P3+4P5P_{5}+2P_{3}+4P_{5}
P5+2P4+3P3P_{5}+2P_{4}+3P_{3} P5+2P4+4P3P_{5}+2P_{4}+4P_{3} P5+2P4+6P4P_{5}+2P_{4}+6P_{4}: [32] P5+2P5+2P5P_{5}+2P_{5}+2P_{5}: [32] P5+2P5+3P3P_{5}+2P_{5}+3P_{3}
P5+2P5+3P5P_{5}+2P_{5}+3P_{5}: [6] P5+2P5+4P5P_{5}+2P_{5}+4P_{5}: [32] P5+2P5+6P3P_{5}+2P_{5}+6P_{3} P5+2P5+6P5P_{5}+2P_{5}+6P_{5}: [6] P5+2P5+8P5P_{5}+2P_{5}+8P_{5}: [18]
P5+2P8+4P5P_{5}+2P_{8}+4P_{5} P5+3P3+3P3P_{5}+3P_{3}+3P_{3} P5+3P3+3P4P_{5}+3P_{3}+3P_{4}: [32] P5+3P3+4P3P_{5}+3P_{3}+4P_{3} P5+3P3+5P5P_{5}+3P_{3}+5P_{5}
P5+3P4+6P3P_{5}+3P_{4}+6P_{3} P5+3P5+3P5P_{5}+3P_{5}+3P_{5}: [6] P5+3P5+4P5P_{5}+3P_{5}+4P_{5}: [6] P5+3P5+6P5P_{5}+3P_{5}+6P_{5}: [32] P5+3P5+7P5P_{5}+3P_{5}+7P_{5}: [18]
P5+3P5+8P5P_{5}+3P_{5}+8P_{5}: [18] P5+3P5+9P5P_{5}+3P_{5}+9P_{5}: [6] P7+P8+2P3P_{7}+P_{8}+2P_{3} P8+P8+2P3P_{8}+P_{8}+2P_{3} P8+P8+2P4P_{8}+P_{8}+2P_{4}: [9]
P8+P8+2P5P_{8}+P_{8}+2P_{5} P8+P10+2P5P_{8}+P_{10}+2P_{5} P8+2P3+2P3P_{8}+2P_{3}+2P_{3} P8+2P3+2P5P_{8}+2P_{3}+2P_{5} P8+2P3+3P3P_{8}+2P_{3}+3P_{3}
P8+2P3+3P4P_{8}+2P_{3}+3P_{4}: [32] P8+2P4+2P5P_{8}+2P_{4}+2P_{5} P8+2P4+3P4P_{8}+2P_{4}+3P_{4}: [9] P8+2P4+4P3P_{8}+2P_{4}+4P_{3} P8+2P5+2P5P_{8}+2P_{5}+2P_{5}
P8+2P5+3P3P_{8}+2P_{5}+3P_{3} P8+2P5+4P3P_{8}+2P_{5}+4P_{3}
Theorem 1.5.

Excluding sums involving generalized hexagonal numbers, there are at most 30523052 ternary universal sums. Among them, there are 197197 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 𝒞𝔐\mathcal{C}_{\mathfrak{M}}. For any integers 𝔪,𝔐\mathfrak{m},\mathfrak{M}\in\mathbb{Z} such that 𝔐𝔪21\mathfrak{M}\geq\mathfrak{m}-2\geq 1, we define the subclass 𝒞𝔪,𝔐\mathcal{C}_{\mathfrak{m},\mathfrak{M}} as the class of sums of generalized polygonal numbers with parameters m1,,mr3m_{1},\ldots,m_{r}\geq 3 such that lcm(m12,,mr2)𝔐\text{lcm}(m_{1}-2,\ldots,m_{r}-2)\leq\mathfrak{M} and min(m1,,mr)𝔪\min(m_{1},\ldots,m_{r})\geq\mathfrak{m}. For these classes, we can establish effective finiteness theorems.

Theorem 1.6.

For any integers 𝔪,𝔐\mathfrak{m},\mathfrak{M}\in\mathbb{Z} such that 𝔐𝔪21\mathfrak{M}\geq\mathfrak{m}-2\geq 1, there exists a minimal constant Γ𝔪,𝔐>0\Gamma_{\mathfrak{m},\mathfrak{M}}>0 such that every sum of generalized polygonal numbers in 𝒞𝔪,𝔐\mathcal{C}_{\mathfrak{m},\mathfrak{M}} is universal if and only if it represents every positive integer up to Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}}. For any real number ε>0\varepsilon>0, we have the following upper bounds for the constant Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}}:

  1. (1)

    If 𝔪20\mathfrak{m}\geq 20, then we have

    Γ𝔪,𝔐ε𝔐43+ε.\Gamma_{\mathfrak{m},\mathfrak{M}}\ll_{\varepsilon}\mathfrak{M}^{43+\varepsilon}.
  2. (2)

    If 𝔪36\mathfrak{m}\geq 36, then we have

    Γ𝔪,𝔐ε𝔐27+ε.\Gamma_{\mathfrak{m},\mathfrak{M}}\ll_{\varepsilon}\mathfrak{M}^{27+\varepsilon}.

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 Γ𝔐\Gamma_{\mathfrak{M}} and Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}}, 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 𝒞\mathcal{C} of sums of generalized polygonal numbers. The escalator tree T𝒞T_{\mathcal{C}} for universal sums of generalized polygonal numbers in 𝒞\mathcal{C} is a rooted tree constructed inductively as follows. We define the root to be F=0F=0 with depth 0 and then inductively construct the nodes of depth r+1r+1 from the nodes of depth rr as follows. If a node of depth rr is universal, then it is a leaf of the tree. If a sum of generalized polygonal numbers FF is not universal, then we call the smallest positive integer t(F)t(F) not represented by FF the truant of FF; for ease of notation, we write t(F)t(F)\coloneqq\infty if FF is universal. The children of a node F(x1,,xr)=j=1rajPmj(xj)F(x_{1},\ldots,x_{r})=\sum_{j=1}^{r}a_{j}P_{m_{j}}(x_{j}) with t(F)<t(F)<\infty are the sums of generalized polygonal numbers

F(x1,,xr)+ar+1Pmr+1(xr+1)𝒞F(x_{1},\cdots,x_{r})+a_{r+1}P_{m_{r+1}}(x_{r+1})\in\mathcal{C}

with arar+1t(F)a_{r}\leq a_{r+1}\leq t(F), mr+16m_{r+1}\neq 6, and an additional restriction that mrmr+1m_{r}\leq m_{r+1} if ar=ar+1a_{r}=a_{r+1} to avoid repeated nodes. For any class 𝒞\mathcal{C}, we define the set

𝒯𝒞{t(F)<FT𝒞}.\mathscr{T}_{\mathcal{C}}\coloneqq\left\{t(F)<\infty\mid F\in T_{\mathcal{C}}\right\}.

By construction, this subset has the property that if a sum F𝒞F\in\mathcal{C} represents every integer in 𝒯𝒞\mathscr{T}_{\mathcal{C}}, then FF is universal. We also see that 𝒯𝒞\mathscr{T}_{\mathcal{C}} is minimal in the sense that it is the smallest subset of \mathbb{N} with this property. Therefore, if 𝒯𝒞\mathscr{T}_{\mathcal{C}} is a finite set, we obtain a finiteness theorem by taking Γ𝒞\Gamma_{\mathcal{C}} to be the maximal integer contained in 𝒯𝒞\mathscr{T}_{\mathcal{C}}.

We denote TT_{\infty} the escalator tree for the class 𝒞\mathcal{C}_{\infty} of arbitrary sums of generalized polygonal numbers. For any integer 𝔐1\mathfrak{M}\geq 1, we abbreviate T𝔐T_{\mathfrak{M}} for the tree T𝒞𝔐T_{\mathcal{C}_{\mathfrak{M}}}. We will show in Theorem 6.1 that T𝔐T_{\mathfrak{M}} is a finite tree. Therefore, the existence of the constant Γ𝔐>0\Gamma_{\mathfrak{M}}>0 in Theorem 1.1 is established. Similarly, we abbreviate T𝔪,𝔐T_{\mathfrak{m},\mathfrak{M}} for the tree T𝒞𝔪,𝔐T_{\mathcal{C}_{\mathfrak{m},\mathfrak{M}}} for any integers 𝔪,𝔐\mathfrak{m},\mathfrak{M}\in\mathbb{Z} such that 𝔐𝔪21\mathfrak{M}\geq\mathfrak{m}-2\geq 1. Then the existence of the constant Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}} in Theorem 1.6 follows immediately because Γ𝔪,𝔐Γ𝔐\Gamma_{\mathfrak{m},\mathfrak{M}}\leq\Gamma_{\mathfrak{M}}. To study the growth of the constants Γ𝔐\Gamma_{\mathfrak{M}} and Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}}, it is clear that we have to study the truants of non-leaf nodes in the escalator tree TT_{\infty}.

We begin with an observation on the truants of sums.

Lemma 2.1.

Fix 1ir1\leq i\leq r. Let F=a1Pm1++aiPmi++arPmrF^{\prime}=a_{1}P_{m_{1}}+\cdots+a_{i}P_{m_{i}^{\prime}}+\cdots+a_{r}P_{m_{r}} be a sum with t(F)<t(F^{\prime})<\infty. For any integer mi4+t(F)/aim_{i}\geq 4+\lfloor t(F^{\prime})/a_{i}\rfloor, the sum F=a1Pm1++aiPmi++arPmrF=a_{1}P_{m_{1}}+\cdots+a_{i}P_{m_{i}}+\cdots+a_{r}P_{m_{r}} has truant t(F)=ct(F)t(F)=c\leq t(F^{\prime}), where cc is a constant depending on aja_{j} with 1jr1\leq j\leq r and mjm_{j} with jij\neq i.

Proof.

If t(F)>t(F)t(F)>t(F^{\prime}) for some mi4+t(F)/aim_{i}\geq 4+\lfloor t(F^{\prime})/a_{i}\rfloor, then there exists x1,,xrx_{1},\ldots,x_{r}\in\mathbb{Z} such that

a1Pm1(x1)++aiPmi(xi)++arPmr(xr)=t(F).a_{1}P_{m_{1}}(x_{1})+\cdots+a_{i}P_{m_{i}}(x_{i})+\cdots+a_{r}P_{m_{r}}(x_{r})=t(F^{\prime}).

Since FF^{\prime} cannot represent t(F)t(F^{\prime}), we see that xi0,1x_{i}\neq 0,1. Noticing that mi4m_{i}\geq 4, we have Pmi(xi)mi3P_{m_{i}}(x_{i})\geq m_{i}-3. Therefore, we have

a1Pm1(x1)++aiPmi(xi)++arPmr(xr)aiPmi(xi)ai(1+t(F)/ai)>t(F),a_{1}P_{m_{1}}(x_{1})+\cdots+a_{i}P_{m_{i}}(x_{i})+\cdots+a_{r}P_{m_{r}}(x_{r})\geq a_{i}P_{m_{i}}(x_{i})\geq a_{i}(1+\lfloor t(F^{\prime})/a_{i}\rfloor)>t(F^{\prime}),

which is a contradiction. This shows that t(F)t(F)t(F)\leq t(F^{\prime}). For the same reason, the representations of integers nt(F)1n\leq t(F^{\prime})-1 are independent of the choice of mim_{i}. Thus, the truant t(F)t(F) is a constant cc for any integer mi4+t(F)/aim_{i}\geq 4+\lfloor t(F^{\prime})/a_{i}\rfloor. ∎

As a consequence, we obtain the following useful enumeration of the nodes in TT_{\infty} of depth r3r\leq 3.

Proposition 2.2.

Suppose that a1Pm1+a2Pm2a_{1}P_{m_{1}}+a_{2}P_{m_{2}} is a node in TT_{\infty}. Then, a1=1a_{1}=1 and 1a231\leq a_{2}\leq 3. For each choice of a2a_{2}, the truants are summarized in Table 2.1.

Proof.

One can calculate the truants of nodes with m1,m210m_{1},m_{2}\leq 10. The calculation of the truants of other nodes follows from by Lemma 2.1. For example, to show that t(P3+Pm2)=5t(P_{3}+P_{m_{2}})=5 for any integer m29m_{2}\geq 9, one applies Lemma 2.1 with i=2i=2 and F=P3+P6F^{\prime}=P_{3}+P_{6}. ∎

Table 2.1. Truants of nodes of depth 22
a2=1a_{2}=1 a2=2a_{2}=2 a2=3a_{2}=3
m2m_{2} m1m_{1} 3 4 5 7 8 9\geq 9 3 4 5 7 8 9 10\geq 10 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
9\geq 9 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 r1r\geq 1 and parameters a1,,ara_{1},\ldots,a_{r}\in\mathbb{Z}. Suppose that there exist integers AiA_{i} and BiB_{i} such that BiAi3B_{i}\geq A_{i}\geq 3 for any 1ir1\leq i\leq r with the property: For any 1ir1\leq i\leq r and for any integers AjmjBjA_{j}\leq m_{j}\leq B_{j} for 1jir1\leq j\neq i\leq r, there exists a non-universal sum F=a1Pm1++aiPmi++arPmrF^{\prime}=a_{1}P_{m_{1}}+\cdots+a_{i}P_{m_{i}^{\prime}}+\cdots+a_{r}P_{m_{r}} with AimiBiA_{i}\leq m_{i}^{\prime}\leq B_{i} such that t(F)/ak+3Bk\left\lfloor t(F^{\prime})/a_{k}\right\rfloor+3\leq B_{k} for any 1kr1\leq k\leq r. Then any sum FF of the form a1Pm1++arPmra_{1}P_{m_{1}}+\cdots+a_{r}P_{m_{r}} with miAim_{i}\geq A_{i} for any 1ir1\leq i\leq r and mjBj+1m_{j}\geq B_{j}+1 for some 1jr1\leq j\leq r has the truant

t(F)tmaxmax{t(a1Pm1++arPmr)<AimiBi for 1ir}.t(F)\leq t_{\max}\coloneqq\max\{t(a_{1}P_{m_{1}}+\cdots+a_{r}P_{m_{r}})<\infty\mid A_{i}\leq m_{i}\leq B_{i}\text{ for }1\leq i\leq r\}.
Proof.

Let FF be any sum of the form a1Pm1++arPmra_{1}P_{m_{1}}+\cdots+a_{r}P_{m_{r}} with miAim_{i}\geq A_{i} for any 1ir1\leq i\leq r and mjBj+1m_{j}\geq B_{j}+1 for some 1jr1\leq j\leq r. Let ss be the number of indices 1jr1\leq j\leq r such that mjBj+1m_{j}\geq B_{j}+1. By induction on the integer s1s\geq 1, we prove the lemma jointly with the following inequality

t(F)ak+3Bk\left\lfloor\frac{t(F)}{a_{k}}\right\rfloor+3\leq B_{k}

for any 1kr1\leq k\leq r. If s=1s=1, without loss of generality, we assume that m1B1+1m_{1}\geq B_{1}+1. By the assumption of the lemma, there exists a sum F=a1Pm1++arPmrF^{\prime}=a_{1}P_{m_{1}^{\prime}}+\cdots+a_{r}P_{m_{r}} such that t(F)tmaxt(F^{\prime})\leq t_{\max} and t(F)/ak+3Bk\left\lfloor t(F^{\prime})/a_{k}\right\rfloor+3\leq B_{k} for any 1kr1\leq k\leq r. Using Lemma 2.1, we see that t(F)t(F)tmaxt(F)\leq t(F^{\prime})\leq t_{\max} and it follows that for any 1kr1\leq k\leq r, we have

t(F)ak+3t(F)ak+3Bk,\left\lfloor\frac{t(F)}{a_{k}}\right\rfloor+3\leq\left\lfloor\frac{t(F^{\prime})}{a_{k}}\right\rfloor+3\leq B_{k},

as desired. Next we prove the induction step. Assuming the lemma and the inequality hold for s=k1s=k-1, we prove them for 2s=kr2\leq s=k\leq r. Without loss of generality, we may assume that m1B1+1m_{1}\geq B_{1}+1. Then we choose any sum F=a1Pm1++arPmrF^{\prime}=a_{1}P_{m_{1}^{\prime}}+\cdots+a_{r}P_{m_{r}} with A1m1B1A_{1}\leq m_{1}^{\prime}\leq B_{1}. By the inductive hypothesis, we have t(F)tmaxt(F^{\prime})\leq t_{\max} and

t(F)ak+3Bk\left\lfloor\frac{t(F^{\prime})}{a_{k}}\right\rfloor+3\leq B_{k}

for any 1kr1\leq k\leq r. Again by Lemma 2.1, we can conclude that t(F)t(F)tmaxt(F)\leq t(F^{\prime})\leq t_{\max} and the sum FF satisfies the inequality for any 1kr1\leq k\leq r. This finishes the proof. ∎

Next we prove upper bounds on the truants of non-leaf nodes of depth 33 and depth 44.

Proposition 2.4.

Excluding a set of 30523052 nodes of depth 33 with m1,m2161m_{1},m_{2}\leq 161 and m378m_{3}\leq 78, any other node FF of depth 33 has truant t(F)644t(F)\leq 644. In particular, a ternary sum is universal only if it is among the 30523052 excluded nodes.

Proof.

With A1=A2=A3=3A_{1}=A_{2}=A_{3}=3, we search for the integers B1,B2,B3B_{1},B_{2},B_{3} satisfying the assumptions of Lemma 2.3 for each choice of integers a1,a2,a3a_{1},a_{2},a_{3}. By a computer program, we can verify that the integers A1=A2=A3=3A_{1}=A_{2}=A_{3}=3 and B1=B2=161,B3=78B_{1}=B_{2}=161,B_{3}=78 satisfy the assumptions for any possible choice of integers a1,a2,a3a_{1},a_{2},a_{3} of nodes in the escalator tree TT_{\infty}. By a computer program to check the representations by any node of depth 33 with parameters AimiBiA_{i}\leq m_{i}\leq B_{i} for any 1i31\leq i\leq 3 up to 30003000, we find that the truant of any node is bounded above by 644644 except 30523052 nodes representing every positive integer up to 30003000. 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 44, we have the following facts:

  1. (1)

    Suppose that mi1883m_{i}\leq 1883 for at least three indices {1i4}\{1\leq i\leq 4\}. The truant of a non-leaf node a1Pm1+a2Pm2+a3Pm3+a4Pm4a_{1}P_{m_{1}}+a_{2}P_{m_{2}}+a_{3}P_{m_{3}}+a_{4}P_{m_{4}} is bounded above by an implicit constant.

  2. (2)

    Suppose that mi1883m_{i}\leq 1883 for at most two indices {1i4}\{1\leq i\leq 4\}. Excluding a set of finitely many potentially universal nodes, any other node a1Pm1+a2Pm2+a3Pm3+a4Pm4a_{1}P_{m_{1}}+a_{2}P_{m_{2}}+a_{3}P_{m_{3}}+a_{4}P_{m_{4}} is not universal and the truant is bounded above by 18801880.

Proof.

(1) Set M=1883M=1883. For any non-leaf node of depth 44 with miMm_{i}\leq M for at least three indices in {1i4}\{1\leq i\leq 4\}, the truant is bounded by an implicit constant by applying Lemma 2.1, including the potential children of excluded nodes of depth 33.

(2) For any node of depth 44 with miMm_{i}\leq M for at most two indices in {1i4}\{1\leq i\leq 4\}, say 1j,k41\leq j,k\leq 4, we can use a computer program to verify that the integers Ai=3,Bi=MA_{i}=3,B_{i}=M for i=j,ki=j,k and Ai=Bi=MA_{i}=B_{i}=M for ij,ki\neq j,k satisfy the assumptions of Lemma 2.3. Thus, by a computer program, we see that the truant of any node with miMm_{i}\leq M for at most two indices is bounded above by 18801880, 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 mi1883m_{i}\leq 1883 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 p3p\geq 3. Suppose that b1,b2p×b_{1},b_{2}\in\mathbb{Z}_{p}^{\times}, b3pb_{3}\in\mathbb{Z}_{p}, and αpp\alpha\in p\mathbb{Z}_{p} satisfying either αp2p\alpha\in p^{2}\mathbb{Z}_{p} or b3p×b_{3}\in\mathbb{Z}_{p}^{\times}. If the polynomial f(x1,x2,x3)=b1x12+c1x1+b2x22+c2x2+α(b3x32+c3x3)p[x1,x2,x3]f(x_{1},x_{2},x_{3})=b_{1}x_{1}^{2}+c_{1}x_{1}+b_{2}x_{2}^{2}+c_{2}x_{2}+\alpha(b_{3}x_{3}^{2}+c_{3}x_{3})\in\mathbb{Z}_{p}[x_{1},x_{2},x_{3}] represents every class in /p2\mathbb{Z}/p^{2}\mathbb{Z}, then

(b1b2p)=1.\genfrac{(}{)}{}{}{-b_{1}b_{2}}{p}=1.
Proof.

First assume that αp2p\alpha\in p^{2}\mathbb{Z}_{p}. Replacing xix_{i} by xici2bix_{i}-\frac{c_{i}}{2b_{i}} for 1i21\leq i\leq 2, we see that b1x12+b2x22b_{1}x_{1}^{2}+b_{2}x_{2}^{2} represents every class in /p2\mathbb{Z}/p^{2}\mathbb{Z} by the assumption. In particular, there exist w1,w2pw_{1},w_{2}\in\mathbb{Z}_{p} such that b1w12+b2w22p (mod p2)b_{1}w_{1}^{2}+b_{2}w_{2}^{2}\equiv p\text{~{}(mod }p^{2}\text{)}. Hence, we have b1w12+b2w220 (mod p)b_{1}w_{1}^{2}+b_{2}w_{2}^{2}\equiv 0\text{~{}(mod }p\text{)} such that either w1p×w_{1}\in\mathbb{Z}_{p}^{\times} or w2p×w_{2}\in\mathbb{Z}_{p}^{\times}. Without loss of generality, assume that w1p×w_{1}\in\mathbb{Z}_{p}^{\times}. Then, we see that (b2w2/w1)2b1b2 (mod p)(b_{2}w_{2}/w_{1})^{2}\equiv-b_{1}b_{2}\text{~{}(mod }p\text{)}, which is the desired result.

Next assume that b3p×b_{3}\in\mathbb{Z}_{p}^{\times} and αpp×\alpha\in p\mathbb{Z}_{p}^{\times}. Without loss of generality, we may assume that α=p\alpha=p. Replacing xix_{i} by xici2bix_{i}-\frac{c_{i}}{2b_{i}} for 1i31\leq i\leq 3, we see that b1x12+b2x22+pb3x32b_{1}x_{1}^{2}+b_{2}x_{2}^{2}+pb_{3}x_{3}^{2} represents every class in /p2\mathbb{Z}/p^{2}\mathbb{Z} by the assumption. Thus there exist w1,w2,w3pw_{1},w_{2},w_{3}\in\mathbb{Z}_{p} such that b1w12+b2w22+pb3w32kp (mod p2)b_{1}w_{1}^{2}+b_{2}w_{2}^{2}+pb_{3}w_{3}^{2}\equiv kp\text{~{}(mod }p^{2}\text{)} with kp×k\in\mathbb{Z}_{p}^{\times} such that b3kb_{3}k is a non-square modulo pp. Hence we see that b1w12+b2w220 (mod p)b_{1}w_{1}^{2}+b_{2}w_{2}^{2}\equiv 0\text{~{}(mod }p\text{)} such that either w1p×w_{1}\in\mathbb{Z}_{p}^{\times} or w2p×w_{2}\in\mathbb{Z}_{p}^{\times}, and the proof follows as in the previous case. ∎

Lemma 2.9.

Suppose that p3p\geq 3 is a prime number. Let F=a1Pm1+a2Pm2+a3Pm3F=a_{1}P_{m_{1}}+a_{2}P_{m_{2}}+a_{3}P_{m_{3}} be a node in TT_{\infty} such that pa1a2(m12)(m22)p\nmid a_{1}a_{2}(m_{1}-2)(m_{2}-2). If one of the following conditions holds

  1. (1)

    p2a3p^{2}\mid a_{3},

  2. (2)

    pa3p\mid a_{3}, p(m32)p\nmid(m_{3}-2) and t(F)p2t(F)\geq p^{2},

then we have

(a1a2(m12)(m22)p)=1.\genfrac{(}{)}{}{}{-a_{1}a_{2}(m_{1}-2)(m_{2}-2)}{p}=1.
Proof.

If p2a3p^{2}\mid a_{3}, then the construction of TT_{\infty} implies that t(a1Pm1+a2Pm2)p2t(a_{1}P_{m_{1}}+a_{2}P_{m_{2}})\geq p^{2}. Hence t(F)p2t(F)\geq p^{2} in both cases and thus FF represents every class in /p2\mathbb{Z}/p^{2}\mathbb{Z}. The claim then follows by Lemma 2.8. ∎

Lemma 2.10.

Let F=Pm1+a2Pm2+a3Pm3F=P_{m_{1}}+a_{2}P_{m_{2}}+a_{3}P_{m_{3}} be a node in the tree TT_{\infty} with 4m1,m2,m34\mid m_{1},m_{2},m_{3}. The following facts hold.

  1. (1)

    If a2=1a_{2}=1 and a3=1a_{3}=1, then we have t(F)46t(F)\leq 46. If t(F)8t(F)\geq 8, then either (m1,m2)=(4,8)(m_{1},m_{2})=(4,8) or (m1,m2,m3)=(4,4,8)(m_{1},m_{2},m_{3})=(4,4,8).

  2. (2)

    If a2=1a_{2}=1 and a3=2a_{3}=2, then we have t(F)71t(F)\leq 71. If t(F)16t(F)\geq 16, then (m1,m2)=(8,12)(m_{1},m_{2})=(8,12) or (m1,m2,m3){(4,4,8),(4,12,4),(4,16,4),(8,8,4),(8,16,8)}(m_{1},m_{2},m_{3})\in\{(4,4,8),(4,12,4),(4,16,4),(8,8,4),(8,16,8)\}.

  3. (3)

    If a2=1a_{2}=1 and a3=3a_{3}=3, then we have t(F)38t(F)\leq 38. If t(F)8t(F)\geq 8, then (m1,m2)=(4,8)(m_{1},m_{2})=(4,8).

  4. (4)

    If a2=2a_{2}=2, then we have t(F)7t(F)\leq 7 if a3=2a_{3}=2, t(F)10t(F)\leq 10 if a3=3a_{3}=3, t(F)15t(F)\leq 15 if a3=4a_{3}=4, and t(F)12t(F)\leq 12 if a3=5a_{3}=5.

Proof.

This follows from a straightforward calculation that uses the fact that either Pm(x){0,1,m3,m,3m8}P_{m}(x)\in\{0,1,m-3,m,3m-8\} or Pm(x)3m3P_{m}(x)\geq 3m-3. ∎

3. Representations by Shifted Lattices

For an integer r1r\geq 1, a quadratic space VV of dimension rr is a vector space over \mathbb{Q} of dimension rr equipped with a symmetric bilinear form B:V×VB\colon V\times V\to\mathbb{Q}. We can associate VV with a quadratic map Q:VQ\colon V\to\mathbb{Q} defined by Q(v)B(v,v)Q(v)\coloneqq B(v,v) for any vector vVv\in V. A quadratic space VV is positive-definite if Q(v)>0Q(v)>0 for any non-zero vector vVv\in V. A lattice LL in a quadratic space VV of dimension rr is a free \mathbb{Z}-submodule of VV of rank rr. A lattice LL is integral if B(L,L)B(L,L)\subseteq\mathbb{Z}. Set L#{vV2B(v,L)}L^{\#}\coloneqq\{v\in V\mid 2B(v,L)\subseteq\mathbb{Z}\}. For an integral lattice LL, the discriminant is defined as [L#:L][L^{\#}\colon L] and the level is defined as the least positive integer NN such that NQ(v)NQ(v)\in\mathbb{Z} for any vector vL#v\in L^{\#}.

A shifted lattice XX in a quadratic space VV is a subset of VV of the form L+νL+\nu, where LL is a lattice in VV and ν\nu is a vector of VV. A shifted lattice is positive-definite if the underlying quadratic space VV is positive-definite. A shifted lattice XX is integral if B(X,X)B(X,X)\subseteq\mathbb{Z}. The rank of a shifted lattice X=L+νX=L+\nu is the rank of LL as a free \mathbb{Z}-module. For an integral shifted lattice XX, a base lattice of a shifted lattice XX is an integral lattice LXL_{X} containing XX. An integral shifted lattice always admits a base lattice because the lattice generated by XX is a base lattice of XX. The discriminant, the level, and the conductor of an integral shifted lattice XX relative to a base lattice LXL_{X} of XX are defined as the discriminant of LXL_{X}, the level of LXL_{X}, and the least positive integer such that MLXLML_{X}\subseteq L. For a shifted lattice X=L+νX=L+\nu of conductor MM and a,da,d\in\mathbb{Z} with ad1 (mod M)ad\equiv 1\text{~{}(mod }M\text{)}, we let X(d)X(d) be the shifted lattice L+aνL+a\nu. Finally, we say that an element nn\in\mathbb{Q} is represented by a shifted lattice XX or equivalently the shifted lattice XX represents an element nn\in\mathbb{Q} if there is a vector vXv\in X such that Q(v)=nQ(v)=n. Such a vector vXv\in X is called a representation of nn by XX. The number of representations of nn by XX is denoted by rX(n)r_{X}(n).

Let FF be a sum of generalized polygonal numbers. In Section 4, we will construct a shifted lattice XX together with two integers μ1,ρ\mu\geq 1,\rho\in\mathbb{Z} such that rF(n)=rX(μn+ρ)r_{F}(n)=r_{X}(\mu n+\rho) for any integer nn\in\mathbb{Z}. Therefore, to study the representations by FF, it is equivalent to study the representations by the shifted lattice XX. Let ={τIm(τ)>0}\mathbb{H}=\{\tau\in\mathbb{C}\mid\operatorname{Im}(\tau)>0\} be the complex upper half-plane. For a positive-definite integral shifted lattice XX, the theta series ΘX:\Theta_{X}\colon\mathbb{H}\to\mathbb{C} associated to XX is defined as

ΘX(τ)vXe2πiτQ(v)=nrX(n)e2πiτn.\Theta_{X}(\tau)\coloneqq\sum_{v\in X}e^{2\pi i\tau Q(v)}=\sum_{n\in\mathbb{Z}}r_{X}(n)e^{2\pi i\tau n}.

By [23, Proposition 2.1] and the theory of modular forms, the theta series ΘX\Theta_{X} is a modular form, which splits into a sum of an Eisenstein series EXE_{X} and a cusp form GXG_{X}. Let aEX(n)a_{E_{X}}(n) and aGX(n)a_{G_{X}}(n) denote the nn-th Fourier coefficient of EXE_{X} and GXG_{X} for any integer n0n\geq 0, respectively. We have

(3.1) rX(n)=aEX(n)+aGX(n).r_{X}(n)=a_{E_{X}}(n)+a_{G_{X}}(n).

By estimating the Fourier coefficients aEX(n)a_{E_{X}}(n) and aGX(n)a_{G_{X}}(n), we can show that rX(n)r_{X}(n) 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 XX is denoted by LXL_{X} and the quadratic map is denoted by QQ. For r2r\geq 2, we have

(3.2) aEX(n)=cr(2π)r2nr21[LX#:LX]12Γ(r2)pβp(n;X),a_{E_{X}}(n)=\frac{c_{r}(2\pi)^{\frac{r}{2}}n^{\frac{r}{2}-1}}{[L_{X}^{\#}\colon L_{X}]^{\frac{1}{2}}\Gamma(\frac{r}{2})}\prod_{p}\beta_{p}(n;X),

where Γ(x)\Gamma(x) is the usual Gamma function and we have cr1c_{r}\coloneqq 1 if r2r\neq 2 and c2=12c_{2}=\frac{1}{2}; the quantities βp(n;X)\beta_{p}(n;X) are the local densities of XX, which is defined as follows. Let VpV_{p}, LX,pL_{X,p}, and XpX_{p} be the localizations of the quadratic space VV, the base lattice LXL_{X}, and the shifted lattice XX at pp, respectively. We choose the unique Haar measures dv\mathrm{d}v on VpV_{p} and dσ\mathrm{d}\sigma on p\mathbb{Q}_{p} such that

LX,pdv=pdσ=1.\int_{L_{X,p}}\mathrm{d}v=\int_{\mathbb{Z}_{p}}\mathrm{d}\sigma=1.

For any positive integer nn and any prime number pp, the local density βp(n;X)\beta_{p}(n;X) of a shifted lattice XX in the quadratic space VV is defined as the integral

βp(n;X)pXpep(σ(Q(v)n))dvdσ.\beta_{p}(n;X)\coloneqq\int_{\mathbb{Q}_{p}}\int_{X_{p}}e_{p}\big{(}\sigma(Q(v)-n)\big{)}\mathrm{d}v\mathrm{d}\sigma.

where the function ep:pe_{p}\colon\mathbb{Q}_{p}\to\mathbb{C} is defined as follows. For any pp-adic number αp\alpha\in\mathbb{Q}_{p}, we define ep(α)e2πiae_{p}(\alpha)\coloneqq e^{-2\pi ia} for some rational number at=1pta\in\bigcup_{t=1}^{\infty}p^{-t}\mathbb{Z} such that αap\alpha-a\in\mathbb{Z}_{p}. 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 βp(n;X)\beta_{p}(n;X) 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 p3p\geq 3. Suppose that we have an integral shifted lattice X=L+νX=L+\nu in a quadratic space VV with the associated quadratic map denoted by QQ and choose a base lattice LXL_{X}. By Jordan canonical form theorem [16, Theorem 5.2.4 and Section 5.3], there exists a basis {η1,,ηr}\{\eta_{1},\ldots,\eta_{r}\} of the localization LpL_{p} such that the Gram matrix of LpL_{p} is a diagonal matrix with entries A1,,ArpA_{1},\ldots,A_{r}\in\mathbb{Z}_{p} and ν=s1η1++srηr\nu=s_{1}\eta_{1}+\cdots+s_{r}\eta_{r} with rational numbers s1,,srps_{1},\ldots,s_{r}\in\mathbb{Q}_{p}. Then, we have

(4.1) βp(n;X)\displaystyle\beta_{p}(n;X) =pXpep(σ(Q(v)n))dvdσ\displaystyle=\int_{\mathbb{Q}_{p}}\int_{X_{p}}e_{p}\big{(}\sigma(Q(v)-n)\big{)}\mathrm{d}v\mathrm{d}\sigma
=pordp([LX:L])pprep(σ(Q(i=1r(xi+si)ηi)n))dx1dxrdσ\displaystyle=p^{-\text{ord}_{p}([L_{X}\colon L])}\int_{\mathbb{Q}_{p}}\int_{\mathbb{Z}_{p}^{r}}e_{p}\bigg{(}\sigma\bigg{(}Q\bigg{(}\sum_{i=1}^{r}(x_{i}+s_{i})\eta_{i}\bigg{)}-n\bigg{)}\bigg{)}\mathrm{d}x_{1}\cdots\mathrm{d}x_{r}\mathrm{d}\sigma
=pordp([LX:L])pprep(σ(i=1r(Aixi2+2Aisixi+Aisi2)n))dx1dxrdσ,\displaystyle=p^{-\text{ord}_{p}([L_{X}\colon L])}\int_{\mathbb{Q}_{p}}\int_{\mathbb{Z}_{p}^{r}}e_{p}\bigg{(}\sigma\bigg{(}\sum_{i=1}^{r}\left(A_{i}x_{i}^{2}+2A_{i}s_{i}x_{i}+A_{i}s_{i}^{2}\right)-n\bigg{)}\bigg{)}\mathrm{d}x_{1}\cdots\mathrm{d}x_{r}\mathrm{d}\sigma,

Thus, it boils down to evaluating the integral

Ip(n;ϕ)pprep(σ(ϕ(𝐱)n))d𝐱dσ,I_{p}(n;\phi)\coloneqq\int_{\mathbb{Q}_{p}}\int_{\mathbb{Z}_{p}^{r}}e_{p}\big{(}\sigma\left(\phi(\mathbf{x})-n\right)\big{)}\mathrm{d}\mathbf{x}\mathrm{d}\sigma,

where 𝐱(x1,,xr)\mathbf{x}\coloneqq(x_{1},\ldots,x_{r}) and ϕ(𝐱)=i=1r(bixi2+cixi)\phi(\mathbf{x})=\sum_{i=1}^{r}(b_{i}x_{i}^{2}+c_{i}x_{i}) with bi,cipb_{i},c_{i}\in\mathbb{Z}_{p} for 1ir1\leq i\leq r.

First, we evaluate the integral

Gp(σ;b,c)pep(σ(bx2+cx))dx,G_{p}(\sigma;b,c)\coloneqq\int_{\mathbb{Z}_{p}}e_{p}\left(\sigma(bx^{2}+cx)\right)\mathrm{d}x,

for σp\sigma\in\mathbb{Q}_{p} and b,cpb,c\in\mathbb{Z}_{p}.

Lemma 4.1.

Set 𝔱min(ordp(b),ordp(c))\mathfrak{t}\coloneqq\min(\text{ord}_{p}(b),\text{ord}_{p}(c)). Suppose that σ=uptp\sigma=up^{t}\in\mathbb{Z}_{p} such that up×u\in\mathbb{Z}_{p}^{\times} and tt\in\mathbb{Z}. Then we have

Gp(σ;b,c)={1,if t+𝔱0,0,if t+𝔱<0,ordp(b)>ordp(c),ep(σc24b)γp(σb)pt+𝔱2,if t+𝔱<0,ordp(b)ordp(c),G_{p}(\sigma;b,c)=\begin{cases}\displaystyle 1,&\text{if }t+\mathfrak{t}\geq 0,\\ \displaystyle 0,&\text{if }t+\mathfrak{t}<0,\text{ord}_{p}(b)>\text{ord}_{p}(c),\\ \displaystyle e_{p}\left(-\frac{\sigma c^{2}}{4b}\right)\cdot\gamma_{p}(\sigma b)\cdot p^{\frac{t+\mathfrak{t}}{2}},&\text{if }t+\mathfrak{t}<0,\text{ord}_{p}(b)\leq\text{ord}_{p}(c),\\ \end{cases}

where for x=uxptxx=u_{x}p^{t_{x}} with uxp×u_{x}\in\mathbb{Z}_{p}^{\times} and txt_{x}\in\mathbb{Z}, we define

γp(x)={1,if tx is even,εp3(uxp),if tx is odd,\gamma_{p}(x)=\begin{dcases}\displaystyle 1,&\text{if }t_{x}\text{ is even},\\ \displaystyle\varepsilon_{p}^{3}\cdot\genfrac{(}{)}{}{}{u_{x}}{p},&\text{if }t_{x}\text{ is odd},\end{dcases}

and

εp{1,if p1 (mod 4),i,if p3 (mod 4).\varepsilon_{p}\coloneqq\begin{dcases}1,&\text{if }p\equiv 1\text{~{}(mod }4\text{)},\\ i,&\text{if }p\equiv 3\text{~{}(mod }4\text{)}.\end{dcases}
Proof.

The first case is obvious. For the second case, replacing xx by x+pt𝔱1x+p^{-t-\mathfrak{t}-1} in the integral, we have Gp(σ;b,c)=ep(σcpt𝔱1)Gp(σ;b,c)G_{p}(\sigma;b,c)=e_{p}(\sigma cp^{-t-\mathfrak{t}-1})\cdot G_{p}(\sigma;b,c), where the extra factors containing bb disappear because ep(x)=1e_{p}(x)=1 for xpx\in\mathbb{Z}_{p}. Since ep(σcpt𝔱1)1e_{p}(\sigma cp^{-t-\mathfrak{t}-1})\neq 1, the integral vanishes. For the last case, replacing xx by xc2bx-\frac{c}{2b} in the integral, we have

Gp(σ;b,c)=ep(σc24b)pep(σbx2)dx.G_{p}(\sigma;b,c)=e_{p}\left(-\frac{\sigma c^{2}}{4b}\right)\cdot\int_{\mathbb{Z}_{p}}e_{p}(\sigma bx^{2})\mathrm{d}x.

Applying [36, Lemma 2.1(1)], we obtain the desired result. ∎

For the sake of simplicty, we define \infty to be a formal symbol with the properties tt\leq\infty and t+=+t=t+\infty=\infty+t=\infty for any integer tt\in\mathbb{Z}. We set a convention that taking the minimum among an empty set outputs the formal symbol \infty.

Theorem 4.2.

Let p3p\geq 3 be an odd prime number. Suppose that npn\in\mathbb{Z}_{p} and ϕ(𝐱)=i=1r(bixi2+cixi)\phi(\mathbf{x})=\sum_{i=1}^{r}(b_{i}x_{i}^{2}+c_{i}x_{i}) with bi,cipb_{i},c_{i}\in\mathbb{Z}_{p} for 1ir1\leq i\leq r. For 1ir1\leq i\leq r, we define timin(ordp(bi),ordp(ci))t_{i}\coloneqq\min(\text{ord}_{p}(b_{i}),\text{ord}_{p}(c_{i})). We set

Dp{1irordp(bi)>ordp(ci)},\displaystyle D_{p}\coloneqq\{1\leq i\leq r\mid\text{ord}_{p}(b_{i})>\text{ord}_{p}(c_{i})\},
Np{1irordp(bi)ordp(ci)},\displaystyle N_{p}\coloneqq\{1\leq i\leq r\mid\text{ord}_{p}(b_{i})\leq\text{ord}_{p}(c_{i})\},

and set 𝔱dmin{tiiDp}\mathfrak{t}_{d}\coloneqq\min\{t_{i}\mid i\in D_{p}\}. We further define

𝔫n+iNpci24bi.\mathfrak{n}\coloneqq n+\sum_{i\in N_{p}}\frac{c_{i}^{2}}{4b_{i}}.

If 𝔫0\mathfrak{n}\neq 0, we assume that 𝔫=𝔲np𝔱n\mathfrak{n}=\mathfrak{u}_{n}p^{\mathfrak{t}_{n}} with 𝔲np×\mathfrak{u}_{n}\in\mathbb{Z}_{p}^{\times} and 𝔱n\mathfrak{t}_{n}\in\mathbb{Z}. Otherwise we set 𝔱n\mathfrak{t}_{n}\coloneqq\infty. For an integer tt\in\mathbb{Z}, we define

p(t){iNptit<0 and odd},p(t)|p(t)|.\mathcal{L}_{p}(t)\coloneqq\{i\in N_{p}\mid t_{i}-t<0\text{ and odd}\},~{}\ell_{p}(t)\coloneqq|\mathcal{L}_{p}(t)|.

Then, we have

Ip(n;ϕ)=1+(11p)1tmin(𝔱d,𝔱n)p(t) evenδp(t)pτp(t)+δp(𝔱n+1)ωppτp(𝔱n+1),I_{p}(n;\phi)=1+\left(1-\frac{1}{p}\right)\sum_{\begin{subarray}{c}1\leq t\leq\min(\mathfrak{t}_{d},\mathfrak{t}_{n})\\ \ell_{p}(t)\text{ even}\end{subarray}}\delta_{p}(t)p^{\tau_{p}(t)}+\delta_{p}(\mathfrak{t}_{n}+1)\omega_{p}p^{\tau_{p}(\mathfrak{t}_{n}+1)},

where for any integer tt\in\mathbb{Z}, we define

δp(t)εp3p(t)ip(t)(uip),τp(t)t+iNpti<ttit2,\delta_{p}(t)\coloneqq\varepsilon_{p}^{3\ell_{p}(t)}\prod_{i\in\mathcal{L}_{p}(t)}\genfrac{(}{)}{}{}{u_{i}}{p},~{}\tau_{p}(t)\coloneqq t+\sum_{\begin{subarray}{c}i\in N_{p}\\ t_{i}<t\end{subarray}}\frac{t_{i}-t}{2},

with uipordp(bi)bip×u_{i}\coloneqq p^{-\text{ord}_{p}(b_{i})}b_{i}\in\mathbb{Z}_{p}^{\times} and define

ωp{0,if 𝔱n𝔱d,1p,if 𝔱n<𝔱d and p(𝔱n+1) is even,εp(𝔲np)1p,if 𝔱n<𝔱d and p(𝔱n+1) is odd.\omega_{p}\coloneqq\begin{dcases}\displaystyle 0,&\text{if }\mathfrak{t}_{n}\geq\mathfrak{t}_{d},\\ \displaystyle-\frac{1}{p},&\text{if }\mathfrak{t}_{n}<\mathfrak{t}_{d}\text{ and }\ell_{p}(\mathfrak{t}_{n}+1)\text{ is even},\\ \displaystyle\varepsilon_{p}\genfrac{(}{)}{}{}{\mathfrak{u}_{n}}{p}\frac{1}{\sqrt{p}},&\text{if }\mathfrak{t}_{n}<\mathfrak{t}_{d}\text{ and }\ell_{p}(\mathfrak{t}_{n}+1)\text{ is odd}.\end{dcases}
Proof.

By Lemma 4.1, we see that

Ip(n;ϕ)\displaystyle I_{p}(n;\phi) =pep(σn)i=1rGp(σ;bi,ci)dσ\displaystyle=\int_{\mathbb{Q}_{p}}e_{p}(-\sigma n)\prod_{i=1}^{r}G_{p}(\sigma;b_{i},c_{i})\mathrm{d}\sigma
=1+1t𝔱dptp×ep(ptσn)i=1rGp(ptσ,bi,ci)dσ\displaystyle=1+\sum_{1\leq t\leq\mathfrak{t}_{d}}p^{t}\int_{\mathbb{Z}_{p}^{\times}}e_{p}(-p^{-t}\sigma n)\prod_{i=1}^{r}G_{p}(p^{-t}\sigma,b_{i},c_{i})\mathrm{d}\sigma
=1+1t𝔱dδp(t)pτp(t)p×(σp)p(t)ep(ptσ(n+iNpti<tci24bi))dσ.\displaystyle=1+\sum_{1\leq t\leq\mathfrak{t}_{d}}\delta_{p}(t)p^{\tau_{p}(t)}\int_{\mathbb{Z}_{p}^{\times}}\genfrac{(}{)}{}{}{\sigma}{p}^{\ell_{p}(t)}e_{p}\Bigg{(}-p^{-t}\sigma\bigg{(}n+\sum_{\begin{subarray}{c}i\in N_{p}\\ t_{i}<t\end{subarray}}\frac{c_{i}^{2}}{4b_{i}}\bigg{)}\Bigg{)}\mathrm{d}\sigma.

By definition, we have

n+iNpti<tci24bi=𝔫iNptitci24bi𝔫 (mod ptp).n+\sum_{\begin{subarray}{c}i\in N_{p}\\ t_{i}<t\end{subarray}}\frac{c_{i}^{2}}{4b_{i}}=\mathfrak{n}-\sum_{\begin{subarray}{c}i\in N_{p}\\ t_{i}\geq t\end{subarray}}\frac{c_{i}^{2}}{4b_{i}}\equiv\mathfrak{n}\text{~{}(mod }p^{t}\mathbb{Z}_{p}\text{)}.

Thus, by [36, Lemma 2.4], we can conclude that

Ip(n;ϕ)\displaystyle I_{p}(n;\phi) =1+1t𝔱dp(t) even(χptp(𝔫)1pχpt1p(𝔫))δp(t)pτp(t)+1t𝔱dp(t) oddχpt1p×(𝔫)δp(t)ωppτp(t)\displaystyle=1+\sum_{\begin{subarray}{c}1\leq t\leq\mathfrak{t}_{d}\\ \ell_{p}(t)\text{ even}\end{subarray}}\left(\chi_{p^{t}\mathbb{Z}_{p}}(\mathfrak{n})-\frac{1}{p}\chi_{p^{t-1}\mathbb{Z}_{p}}(\mathfrak{n})\right)\delta_{p}(t)p^{\tau_{p}(t)}+\sum_{\begin{subarray}{c}1\leq t\leq\mathfrak{t}_{d}\\ \ell_{p}(t)\text{ odd}\end{subarray}}\chi_{p^{t-1}\mathbb{Z}_{p}^{\times}}(\mathfrak{n})\delta_{p}(t)\omega_{p}p^{\tau_{p}(t)}
=1+(11p)1tmin(𝔱d,𝔱n)p(t) evenδp(t)pτp(t)+δp(𝔱n+1)ωppτp(𝔱n+1),\displaystyle=1+\left(1-\frac{1}{p}\right)\sum_{\begin{subarray}{c}1\leq t\leq\min(\mathfrak{t}_{d},\mathfrak{t}_{n})\\ \ell_{p}(t)\text{ even}\end{subarray}}\delta_{p}(t)p^{\tau_{p}(t)}+\delta_{p}(\mathfrak{t}_{n}+1)\omega_{p}p^{\tau_{p}(\mathfrak{t}_{n}+1)},

where we denote χS(x)\chi_{S}(x) the indicator function of a subset SpS\subseteq\mathbb{Q}_{p} here and throughout. ∎

4.2. An Explicit Formula for Dyadic Local Densities

The dyadic case p=2p=2 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 {η1,,ηr}\{\eta_{1},\ldots,\eta_{r}\} of the localization L2L_{2} such that ν=s1η1++srηr\nu=s_{1}\eta_{1}+\cdots+s_{r}\eta_{r} with rational numbers s1,,sr2s_{1},\ldots,s_{r}\in\mathbb{Q}_{2} and the Gram matrix of L2L_{2} is a block-diagonal matrix with 1×11\times 1 blocks with entries A1,,Ar12A_{1},\ldots,A_{r_{1}}\in\mathbb{Z}_{2} and 2×22\times 2 blocks

B1(0110),,Br2(0110),C1(2112),,Cr3(2112),B_{1}\begin{pmatrix}0&1\\ 1&0\end{pmatrix},\ldots,B_{r_{2}}\begin{pmatrix}0&1\\ 1&0\end{pmatrix},C_{1}\begin{pmatrix}2&1\\ 1&2\end{pmatrix},\ldots,C_{r_{3}}\begin{pmatrix}2&1\\ 1&2\end{pmatrix},

with Bj,Ck2B_{j},C_{k}\in\mathbb{Z}_{2} for 1jr21\leq j\leq r_{2} and 1kr31\leq k\leq r_{3} satisfying r=r1+2(r2+r3)r=r_{1}+2(r_{2}+r_{3}). Arguing in the same way as in the non-dyadic cases, we have

(4.2) β2(n;X)=2ord2([LX:L])I2(n;ϕ)\beta_{2}(n;X)=2^{-\text{ord}_{2}([L_{X}\colon L])}I_{2}(n;\phi)

and it boils down to the calculation of the following integral

I2(n;ϕ)22r1×22r2×22r3e2(σ(ϕ(𝐱,𝐲,𝐳)n))d𝐱d𝐲d𝐳dσ,\displaystyle I_{2}(n;\phi)\coloneqq\int_{\mathbb{Q}_{2}}\int_{\mathbb{Z}_{2}^{r_{1}}\times\mathbb{Z}_{2}^{2r_{2}}\times\mathbb{Z}_{2}^{2r_{3}}}e_{2}\big{(}\sigma\left(\phi(\mathbf{x},\mathbf{y},\mathbf{z})-n\right)\big{)}\mathrm{d}\mathbf{x}\mathrm{d}\mathbf{y}\mathrm{d}\mathbf{z}\mathrm{d}\sigma,

where 𝐱(x1,,xr1)\mathbf{x}\coloneqq(x_{1},\ldots,x_{r_{1}}), 𝐲(y1,1,y1,2,,yr2,1,yr2,2)\mathbf{y}\coloneqq(y_{1,1},y_{1,2},\ldots,y_{r_{2},1},y_{r_{2},2}), 𝐳(z1,1,z1,2,zr3,1,zr3,2)\mathbf{z}\coloneqq(z_{1,1},z_{1,2}\ldots,z_{r_{3},1},z_{r_{3},2}), and ϕ(𝐱,𝐲,𝐳)2[𝐱,𝐲,𝐳]\phi(\mathbf{x},\mathbf{y},\mathbf{z})\in\mathbb{Z}_{2}[\mathbf{x},\mathbf{y},\mathbf{z}] is a polynomial of the form

ϕ(𝐱,𝐲,𝐳)=i=1r1(bixi2+cixi)\displaystyle\phi(\mathbf{x},\mathbf{y},\mathbf{z})=\sum_{i=1}^{r_{1}}(b_{i}x_{i}^{2}+c_{i}x_{i}) +j=1r2(bjyj,1yj,2+cjyj,1+djyj,2)\displaystyle+\sum_{j=1}^{r_{2}}(b_{j}^{\prime}y_{j,1}y_{j,2}+c_{j}^{\prime}y_{j,1}+d_{j}^{\prime}y_{j,2})
+k=1r3(bk′′zk,12+bk′′zk,1zk,2+bk′′zk,22+ck′′zk,1+dk′′zk,2).\displaystyle+\sum_{k=1}^{r_{3}}\left(b_{k}^{\prime\prime}z_{k,1}^{2}+b_{k}^{\prime\prime}z_{k,1}z_{k,2}+b_{k}^{\prime\prime}z_{k,2}^{2}+c_{k}^{\prime\prime}z_{k,1}+d_{k}^{\prime\prime}z_{k,2}\right).

First, we evaluate the following Gauss integrals

G2(σ;b,c)\displaystyle G_{2}(\sigma;b,c)\coloneqq 2e2(σ(bx2+cx))dx,\displaystyle\int_{\mathbb{Z}_{2}}e_{2}\left(\sigma(bx^{2}+cx)\right)\mathrm{d}x,
G2(σ;b,c,d)\displaystyle G_{2}^{\prime}(\sigma;b,c,d)\coloneqq 22e2(σ(by1y2+cy1+dy2))dy1dy2,\displaystyle\int_{\mathbb{Z}_{2}^{2}}e_{2}\left(\sigma(by_{1}y_{2}+cy_{1}+dy_{2})\right)\mathrm{d}y_{1}\mathrm{d}y_{2},
G2′′(σ;b,c,d)\displaystyle G_{2}^{\prime\prime}(\sigma;b,c,d)\coloneqq 22e2(σ(b(z12+z22+z1z2)+cz1+dz2))dz1dz2,\displaystyle\int_{\mathbb{Z}_{2}^{2}}e_{2}\left(\sigma(b(z_{1}^{2}+z_{2}^{2}+z_{1}z_{2})+cz_{1}+dz_{2})\right)\mathrm{d}z_{1}\mathrm{d}z_{2},

for σ2\sigma\in\mathbb{Q}_{2} and b,c,d2b,c,d\in\mathbb{Z}_{2}.

Lemma 4.3.

Set 𝔱min(ord2(b),ord2(c))\mathfrak{t}\coloneqq\min(\text{ord}_{2}(b),\text{ord}_{2}(c)). Suppose that σ=u2t2\sigma=u2^{t}\in\mathbb{Z}_{2} such that u2×u\in\mathbb{Z}_{2}^{\times} and tt\in\mathbb{Z}. Then we have

G2(σ;b,c)={1,if t+𝔱0,0,if t+𝔱<0,ord2(b)>ord2(c),1,if t+𝔱=1,ord2(b)=ord2(c),0,if t+𝔱<1,ord2(b)=ord2(c),0,if t+𝔱=1,ord2(b)<ord2(c),e2(uub8σc24b)(2uub)1+t+𝔱21+t+𝔱2,if t+𝔱<1,ord2(b)<ord2(c),G_{2}(\sigma;b,c)=\begin{dcases}1,&\text{if }t+\mathfrak{t}\geq 0,\\ 0,&\text{if }t+\mathfrak{t}<0,\text{ord}_{2}(b)>\text{ord}_{2}(c),\\ 1,&\text{if }t+\mathfrak{t}=-1,\text{ord}_{2}(b)=\text{ord}_{2}(c),\\ 0,&\text{if }t+\mathfrak{t}<-1,\text{ord}_{2}(b)=\text{ord}_{2}(c),\\ 0,&\text{if }t+\mathfrak{t}=-1,\text{ord}_{2}(b)<\text{ord}_{2}(c),\\ e_{2}\left(\frac{uu_{b}}{8}-\frac{\sigma c^{2}}{4b}\right)\genfrac{(}{)}{}{}{2}{uu_{b}}^{1+t+\mathfrak{t}}2^{\frac{1+t+\mathfrak{t}}{2}},&\text{if }t+\mathfrak{t}<-1,\text{ord}_{2}(b)<\text{ord}_{2}(c),\end{dcases}

where ub2ord2(b)b2×u_{b}\coloneqq 2^{-\text{ord}_{2}(b)}b\in\mathbb{Z}_{2}^{\times} and we extend the Legendre symbol to 2\mathbb{Q}_{2} via the Hilbert symbol by

(2x){(2,x)2,if x2×,0,if x22×.\genfrac{(}{)}{}{}{2}{x}\coloneqq\begin{dcases}(2,x)_{2},&\text{if }x\in\mathbb{Z}_{2}^{\times},\\ 0,&\text{if }x\in\mathbb{Q}_{2}\setminus\mathbb{Z}_{2}^{\times}.\end{dcases}
Proof.

Applying [36, Lemma 4.3(1)], the proof is essentially the same as the proof of Lemma 4.1, except that the cases when ord2(b)=ord2(c)\text{ord}_{2}(b)=\text{ord}_{2}(c) require a different treatment. Assume that ord2(b)=ord2(c)\text{ord}_{2}(b)=\text{ord}_{2}(c). It is clear that G2(σ;b,c)=1G_{2}(\sigma;b,c)=1 if t+𝔱0t+\mathfrak{t}\geq 0. If t+𝔱<0t+\mathfrak{t}<0, we put

k1t𝔱2.k\coloneqq\left\lfloor\frac{1-t-\mathfrak{t}}{2}\right\rfloor.

Then, we have

G2(σ;b,c)\displaystyle G_{2}(\sigma;b,c) =12kr/2ke2(σ(br2+cr))2e2(2kσx(c+2br))dx\displaystyle=\frac{1}{2^{k}}\sum_{r\in\mathbb{Z}/2^{k}\mathbb{Z}}e_{2}(\sigma(br^{2}+cr))\int_{\mathbb{Z}_{2}}e_{2}(2^{k}\sigma x(c+2br))\mathrm{d}x
=χ2(σ2kc)2kr/2ke2(σ(br2+cr)).\displaystyle=\frac{\chi_{\mathbb{Z}_{2}}(\sigma 2^{k}c)}{2^{k}}\sum_{r\in\mathbb{Z}/2^{k}\mathbb{Z}}e_{2}(\sigma(br^{2}+cr)).

Noticing that χ2(σ2kc)0\chi_{\mathbb{Z}_{2}}(\sigma 2^{k}c)\neq 0 if and only if t+𝔱=1t+\mathfrak{t}=-1, we obtain the desired results. ∎

Lemma 4.4.

Set 𝔱min(ord2(b),ord2(c),ord2(d))\mathfrak{t}\coloneqq\min(\text{ord}_{2}(b),\text{ord}_{2}(c),\text{ord}_{2}(d)). Suppose that σ=u2t2\sigma=u2^{t}\in\mathbb{Z}_{2} such that u2×u\in\mathbb{Z}_{2}^{\times} and tt\in\mathbb{Z}. Then we have

G2(σ;b,c,d)={1,if t+𝔱0,0,if t+𝔱<0,ord2(b)>min(ord2(c),ord2(d)),e2(σcdb)2t+𝔱,if t+𝔱<0,ord2(b)min(ord2(c),ord2(d)).G_{2}^{\prime}(\sigma;b,c,d)=\begin{dcases}1,&\text{if }t+\mathfrak{t}\geq 0,\\ 0,&\text{if }t+\mathfrak{t}<0,\text{ord}_{2}(b)>\min(\text{ord}_{2}(c),\text{ord}_{2}(d)),\\ e_{2}\left(-\frac{\sigma cd}{b}\right)2^{t+\mathfrak{t}},&\text{if }t+\mathfrak{t}<0,\text{ord}_{2}(b)\leq\min(\text{ord}_{2}(c),\text{ord}_{2}(d)).\end{dcases}
Proof.

The first case is obvious. For the second case, by the symmetry, we may assume that ord2(c)ord2(d)\text{ord}_{2}(c)\geq\text{ord}_{2}(d). Therefore, we have

G2(σ;b,c,d)=2e2(σcy1)2e2(σ(by1+d)y2)dy1dy2=2e2(σcy1)χ2(σ(by1+d))dy1.G_{2}^{\prime}(\sigma;b,c,d)=\int_{\mathbb{Z}_{2}}e_{2}(\sigma cy_{1})\int_{\mathbb{Z}_{2}}e_{2}(\sigma(by_{1}+d)y_{2})\mathrm{d}y_{1}\mathrm{d}y_{2}=\int_{\mathbb{Z}_{2}}e_{2}(\sigma cy_{1})\chi_{\mathbb{Z}_{2}}(\sigma(by_{1}+d))\mathrm{d}y_{1}.

The indicator function vanishes identically because ord2(b)>ord2(d)\text{ord}_{2}(b)>\text{ord}_{2}(d) and t+𝔱<0t+\mathfrak{t}<0. Therefore we obtain the desired result. For the third case, we apply the change of the variable y1y1dby_{1}\to y_{1}-\frac{d}{b} to the the second line of the above equation,

G2(σ;b,c,d)=e2(σcdb)2t+𝔱2e2(σ2t𝔱cy1)dy1=e2(σcdb)2t+𝔱,G_{2}^{\prime}(\sigma;b,c,d)=e_{2}\left(-\frac{\sigma cd}{b}\right)2^{t+\mathfrak{t}}\int_{\mathbb{Z}_{2}}e_{2}(\sigma 2^{-t-\mathfrak{t}}cy_{1})\mathrm{d}y_{1}=e_{2}\left(-\frac{\sigma cd}{b}\right)2^{t+\mathfrak{t}},

as desired. ∎

Lemma 4.5.

Set 𝔱min(ord2(b),ord2(c),ord2(d))\mathfrak{t}\coloneqq\min(\text{ord}_{2}(b),\text{ord}_{2}(c),\text{ord}_{2}(d)). Suppose that σ=u2t2\sigma=u2^{t}\in\mathbb{Z}_{2} such that u2×u\in\mathbb{Z}_{2}^{\times} and tt\in\mathbb{Z}. Then we have

G2′′(σ;b,c,d)={1,if t+𝔱0,0,if t+𝔱<0,ord2(b)>min(ord2(c),ord2(d)),(1)t+𝔱e2(σ(c2+d2cd)3b)2t+𝔱,if t+𝔱<0,ord2(b)min(ord2(c),ord2(d)).G_{2}^{\prime\prime}(\sigma;b,c,d)=\begin{dcases}1,&\text{if }t+\mathfrak{t}\geq 0,\\ 0,&\text{if }t+\mathfrak{t}<0,\text{ord}_{2}(b)>\min(\text{ord}_{2}(c),\text{ord}_{2}(d)),\\ (-1)^{t+\mathfrak{t}}e_{2}\left(\frac{\sigma(c^{2}+d^{2}-cd)}{3b}\right)2^{t+\mathfrak{t}},&\text{if }t+\mathfrak{t}<0,\text{ord}_{2}(b)\leq\min(\text{ord}_{2}(c),\text{ord}_{2}(d)).\end{dcases}
Proof.

The first case is obvious. For the second case, we may assume that ord2(c)ord2(d)\text{ord}_{2}(c)\geq\text{ord}_{2}(d). Therefore by Lemma 4.3, we have

G2′′(σ;b,c,d)\displaystyle G_{2}^{\prime\prime}(\sigma;b,c,d) =2e2(σ(bz12+cz1))2e2(σ(bz22+(bz1+d)z2))dz2dz1\displaystyle=\int_{\mathbb{Z}_{2}}e_{2}(\sigma(bz_{1}^{2}+cz_{1}))\int_{\mathbb{Z}_{2}}e_{2}(\sigma(bz_{2}^{2}+(bz_{1}+d)z_{2}))\mathrm{d}z_{2}\mathrm{d}z_{1}
=2e2(σ(bz12+cz1))G2(σ;b,bz1+d)dz1=χ2(σd)2e2(σ(bz12+cz1))dz1.\displaystyle=\int_{\mathbb{Z}_{2}}e_{2}(\sigma(bz_{1}^{2}+cz_{1}))G_{2}(\sigma;b,bz_{1}+d)\mathrm{d}z_{1}=\chi_{\mathbb{Z}_{2}}(\sigma d)\int_{\mathbb{Z}_{2}}e_{2}(\sigma(bz_{1}^{2}+cz_{1}))\mathrm{d}z_{1}.

The indicator function vanishes identically because t+𝔱<0t+\mathfrak{t}<0. So we obtain the desired results. For the third case, we apply the changes of variables z1z1+d2c3bz_{1}\to z_{1}+\frac{d-2c}{3b} and z2z2+c2d3bz_{2}\to z_{2}+\frac{c-2d}{3b}. Then we have

G2′′(σ;b,c,d)\displaystyle G_{2}^{\prime\prime}(\sigma;b,c,d) =e2(σ(c2+d2cd)3b)22e2(σb(z12+z22+z1z2))dz1dz2\displaystyle=e_{2}\left(\frac{\sigma(c^{2}+d^{2}-cd)}{3b}\right)\int_{\mathbb{Z}_{2}^{2}}e_{2}(\sigma b(z_{1}^{2}+z_{2}^{2}+z_{1}z_{2}))\mathrm{d}z_{1}\mathrm{d}z_{2}
=(1)t+𝔱e2(σ(c2+d2cd)3b)2t+𝔱,\displaystyle=(-1)^{t+\mathfrak{t}}e_{2}\left(\frac{\sigma(c^{2}+d^{2}-cd)}{3b}\right)2^{t+\mathfrak{t}},

where the last equality follows from [36, Lemma 4.4]. ∎

Theorem 4.6.

Suppose that n2n\in\mathbb{Z}_{2} and ϕ(𝐱,𝐲,𝐳)2[𝐱,𝐲,𝐳]\phi(\mathbf{x},\mathbf{y},\mathbf{z})\in\mathbb{Z}_{2}[\mathbf{x},\mathbf{y},\mathbf{z}] is a polynomial of the form

ϕ(𝐱,𝐲,𝐳)=i=1r1(bixi2+cixi)\displaystyle\phi(\mathbf{x},\mathbf{y},\mathbf{z})=\sum_{i=1}^{r_{1}}(b_{i}x_{i}^{2}+c_{i}x_{i}) +j=1r2(bjyj,1yj,2+cjyj,1+djyj,2)\displaystyle+\sum_{j=1}^{r_{2}}(b_{j}^{\prime}y_{j,1}y_{j,2}+c_{j}^{\prime}y_{j,1}+d_{j}^{\prime}y_{j,2})
+k=1r3(bk′′zk,12+bk′′zk,1zk,2+bk′′zk,22+ck′′zk,1+dk′′zk,2).\displaystyle+\sum_{k=1}^{r_{3}}\left(b_{k}^{\prime\prime}z_{k,1}^{2}+b_{k}^{\prime\prime}z_{k,1}z_{k,2}+b_{k}^{\prime\prime}z_{k,2}^{2}+c_{k}^{\prime\prime}z_{k,1}+d_{k}^{\prime\prime}z_{k,2}\right).

For 1ir11\leq i\leq r_{1}, we define timin(ord2(bi),ord2(ci))t_{i}\coloneqq\min(\text{ord}_{2}(b_{i}),\text{ord}_{2}(c_{i})). For 1jr21\leq j\leq r_{2}, we define tjmin(ord2(bj),ord2(cj),ord2(dj))t_{j}^{\prime}\coloneqq\min(\text{ord}_{2}(b_{j}^{\prime}),\text{ord}_{2}(c_{j}^{\prime}),\text{ord}_{2}(d_{j}^{\prime})). For 1kr31\leq k\leq r_{3}, we define tk′′min(ord2(bk′′),ord2(ck′′),ord2(dk′′))t_{k}^{\prime\prime}\coloneqq\min(\text{ord}_{2}(b_{k}^{\prime\prime}),\text{ord}_{2}(c_{k}^{\prime\prime}),\text{ord}_{2}(d_{k}^{\prime\prime})). We set

D2\displaystyle D_{2}\coloneqq {1ir1ord2(bi)>ord2(ci)},\displaystyle\{1\leq i\leq r_{1}\mid\text{ord}_{2}(b_{i})>\text{ord}_{2}(c_{i})\},
E2\displaystyle E_{2}\coloneqq {1ir1ord2(bi)=ord2(ci)},\displaystyle\{1\leq i\leq r_{1}\mid\text{ord}_{2}(b_{i})=\text{ord}_{2}(c_{i})\},
N2\displaystyle N_{2}\coloneqq {1ir1ord2(bi)<ord2(ci)},\displaystyle\{1\leq i\leq r_{1}\mid\text{ord}_{2}(b_{i})<\text{ord}_{2}(c_{i})\},
D2\displaystyle D_{2}^{\prime}\coloneqq {1jr2ord2(bj)>min(ord2(cj),ord2(dj))},\displaystyle\{1\leq j\leq r_{2}\mid\text{ord}_{2}(b_{j}^{\prime})>\min(\text{ord}_{2}(c_{j}^{\prime}),\text{ord}_{2}(d_{j}^{\prime}))\},
N2\displaystyle N_{2}^{\prime}\coloneqq {1jr2ord2(bj)min(ord2(cj),ord2(dj))},\displaystyle\{1\leq j\leq r_{2}\mid\text{ord}_{2}(b_{j}^{\prime})\leq\min(\text{ord}_{2}(c_{j}^{\prime}),\text{ord}_{2}(d_{j}^{\prime}))\},
D2′′\displaystyle D_{2}^{\prime\prime}\coloneqq {1kr3ord2(bk′′)>min(ord2(ck′′),ord2(dk′′))},\displaystyle\{1\leq k\leq r_{3}\mid\text{ord}_{2}(b_{k}^{\prime\prime})>\min(\text{ord}_{2}(c_{k}^{\prime\prime}),\text{ord}_{2}(d_{k}^{\prime\prime}))\},
N2′′\displaystyle N_{2}^{\prime\prime}\coloneqq {1kr3ord2(bk′′)min(ord2(ck′′),ord2(dk′′))},\displaystyle\{1\leq k\leq r_{3}\mid\text{ord}_{2}(b_{k}^{\prime\prime})\leq\min(\text{ord}_{2}(c_{k}^{\prime\prime}),\text{ord}_{2}(d_{k}^{\prime\prime}))\},

and set

𝔱dmin{{tiiD2}{ti+1iE2}{tjjD2}{tk′′kD2′′}}.\mathfrak{t}_{d}\coloneqq\min\Big{\{}\{t_{i}\mid i\in D_{2}\}\cup\{t_{i}+1\mid i\in E_{2}\}\cup\{t_{j}^{\prime}\mid j\in D_{2}^{\prime}\}\cup\{t_{k}^{\prime\prime}\mid k\in D_{2}^{\prime\prime}\}\Big{\}}.

We further set

𝔫n+iN2ci24bi+jN2cjdjbj+kN2′′ck′′2+dk′′2ck′′dk′′3bk′′.\mathfrak{n}\coloneqq n+\sum_{i\in N_{2}}\frac{c_{i}^{2}}{4b_{i}}+\sum_{j\in N_{2}^{\prime}}\frac{c_{j}^{\prime}d_{j}^{\prime}}{b_{j}^{\prime}}+\sum_{k\in N_{2}^{\prime\prime}}\frac{c_{k}^{\prime\prime 2}+d_{k}^{\prime\prime 2}-c_{k}^{\prime\prime}d_{k}^{\prime\prime}}{3b_{k}^{\prime\prime}}.

If 𝔫0\mathfrak{n}\neq 0, assume that 𝔫=𝔲n2𝔱n\mathfrak{n}=\mathfrak{u}_{n}2^{\mathfrak{t}_{n}} such that 𝔲n2×\mathfrak{u}_{n}\in\mathbb{Z}_{2}^{\times} and 𝔱n\mathfrak{t}_{n}\in\mathbb{Z}. Otherwise we set 𝔱n\mathfrak{t}_{n}\coloneqq\infty. For an integer tt\in\mathbb{Z}, we define

2(t){iN2tit<0 and odd},2(t)|2(t)|,\mathcal{L}_{2}(t)\coloneqq\{i\in N_{2}\mid t_{i}-t<0\text{ and odd}\},~{}\ell_{2}(t)\coloneqq|\mathcal{L}_{2}(t)|,

and

2′′(t){kN2′′tk′′t<0 and odd},2′′(t)|2′′(t)|.\mathcal{L}_{2}^{\prime\prime}(t)\coloneqq\{k\in N_{2}^{\prime\prime}\mid t_{k}^{\prime\prime}-t<0\text{ and odd}\},~{}\ell_{2}^{\prime\prime}(t)\coloneqq|\mathcal{L}_{2}^{\prime\prime}(t)|.

Then, we have

I2(n;ϕ)=1+(112)1tmin(𝔱d,𝔱n+3)tti+1 for iN2δ2(t)2τ2(t),I_{2}(n;\phi)=1+\left(1-\frac{1}{2}\right)\sum_{\begin{subarray}{c}1\leq t\leq\min(\mathfrak{t}_{d},\mathfrak{t}_{n}+3)\\ t\neq t_{i}+1\text{ for }i\in N_{2}\end{subarray}}\delta_{2}(t)2^{\tau_{2}(t)},

where for any integer tt\in\mathbb{Z}, we define

ξ2(t)iN2ti+1<tui23t𝔫,\xi_{2}(t)\coloneqq\sum_{\begin{subarray}{c}i\in N_{2}\\ t_{i}+1<t\end{subarray}}u_{i}-2^{3-t}\mathfrak{n},

with ui2ord2(bi)biu_{i}\coloneqq 2^{-\text{ord}_{2}(b_{i})}b_{i}, define

δ2(t){(1)2′′(t)χ42(ξ2(t))e2(ξ2(t)8)(2i2(t1)ui),if 2(t1) is even,(1)2′′(t)(2ξ2(t)i2(t1)ui)12,if 2(t1) is odd,\delta_{2}(t)\coloneqq\begin{dcases}(-1)^{\ell_{2}^{\prime\prime}(t)}\chi_{4\mathbb{Z}_{2}}(\xi_{2}(t))e_{2}\left(\frac{\xi_{2}(t)}{8}\right)\genfrac{(}{)}{}{}{2}{\prod_{i\in\mathcal{L}_{2}(t-1)}u_{i}},&\text{if }\ell_{2}(t-1)\text{ is even},\\ (-1)^{\ell_{2}^{\prime\prime}(t)}\genfrac{(}{)}{}{}{2}{\xi_{2}(t)\prod_{i\in\mathcal{L}_{2}(t-1)}u_{i}}\frac{1}{\sqrt{2}},&\text{if }\ell_{2}(t-1)\text{ is odd},\end{dcases}

and define

τ2(t)t+12iN2ti+1<t(1+tit)+jN2tj<t(tjt)+kN2′′tk′′<t(tk′′t).\tau_{2}(t)\coloneqq t+\frac{1}{2}\sum_{\begin{subarray}{c}i\in N_{2}\\ t_{i}+1<t\end{subarray}}(1+t_{i}-t)+\sum_{\begin{subarray}{c}j\in N_{2}^{\prime}\\ t_{j}^{\prime}<t\end{subarray}}(t_{j}^{\prime}-t)+\sum_{\begin{subarray}{c}k\in N_{2}^{\prime\prime}\\ t_{k}^{\prime\prime}<t\end{subarray}}(t_{k}^{\prime\prime}-t).
Proof.

By Lemma 4.3, Lemma 4.4, and Lemma 4.5, it is easy to see that

I2(n;ϕ)=1+1tmin(𝔱d,𝔱n+3)tti+1 for iN2(1)2′′(t)(2i2(t1)ui)2τ2(t)2×(2σ)2(t1)e2(σξ28)dσ.I_{2}(n;\phi)=1+\sum_{\begin{subarray}{c}1\leq t\leq\min(\mathfrak{t}_{d},\mathfrak{t}_{n}+3)\\ t\neq t_{i}+1\text{ for }i\in N_{2}\end{subarray}}(-1)^{\ell_{2}^{\prime\prime}(t)}\genfrac{(}{)}{}{}{2}{\prod_{i\in\mathcal{L}_{2}(t-1)}u_{i}}2^{\tau_{2}(t)}\int_{\mathbb{Z}_{2}^{\times}}\genfrac{(}{)}{}{}{2}{\sigma}^{\ell_{2}(t-1)}e_{2}\left(\frac{\sigma\xi_{2}}{8}\right)\mathrm{d}\sigma.

If t>𝔱n+3t>\mathfrak{t}_{n}+3, then ξ2(t)2\xi_{2}(t)\not\in\mathbb{Z}_{2}. Then the integral vanishes by [36, Lemma 4.3(2)]. Suppose that t𝔱n+3t\leq\mathfrak{t}_{n}+3 and 2(t1)\ell_{2}(t-1) is even, we have

2×e2(σξ2(t)8)dσ=12χ42(ξ2(t))e2(ξ2(t)8).\int_{\mathbb{Z}_{2}^{\times}}e_{2}\left(\frac{\sigma\xi_{2}(t)}{8}\right)\mathrm{d}\sigma=\frac{1}{2}\chi_{4\mathbb{Z}_{2}}(\xi_{2}(t))e_{2}\left(\frac{\xi_{2}(t)}{8}\right).

Suppose that t𝔱n+3t\leq\mathfrak{t}_{n}+3 and 2(t1)\ell_{2}(t-1) is odd. Again by [36, Lemma 4.3(2)], we have

2×(2σ)e2(σξ2(t)8)dσ=122(2ξ2(t)),\int_{\mathbb{Z}_{2}^{\times}}\genfrac{(}{)}{}{}{2}{\sigma}e_{2}\left(\frac{\sigma\xi_{2}(t)}{8}\right)\mathrm{d}\sigma=\frac{1}{2\sqrt{2}}\genfrac{(}{)}{}{}{2}{\xi_{2}(t)},

as desired. ∎

4.3. Lower Bounds on the Eisenstein Part

Let F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} be a sum of generalized polygonal numbers. We are going to construct a shifted lattice XX together with integers μ1,ρ\mu\geq 1,\rho\in\mathbb{Z} such that

(4.3) rF(n)=rX(μn+ρ)=aEX(μn+ρ)+aGX(μn+ρ),r_{F}(n)=r_{X}(\mu n+\rho)=a_{E_{X}}(\mu n+\rho)+a_{G_{X}}(\mu n+\rho),

for any integer nn\in\mathbb{Z}. The choice of such a shifted lattice XX with the integers μ\mu and ρ\rho is not necessarily unique. Throughout this paper, we determine them as follows.

Definition 4.7.

Suppose that F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} is a sum of generalized polygonal numbers. We set

Λlcm(m12,,mr2),μ8Λ,ρi=1rai(mi4)2Λmi2.\Lambda\coloneqq\text{lcm}(m_{1}-2,\ldots,m_{r}-2),~{}\mu\coloneqq 8\Lambda,~{}\rho\coloneqq\sum_{i=1}^{r}\frac{a_{i}(m_{i}-4)^{2}\Lambda}{m_{i}-2}.

For any 1ir1\leq i\leq r, we set

αiaiΛmi2,μi2(mi2),ρi4mi.\alpha_{i}\coloneqq\frac{a_{i}\Lambda}{m_{i}-2},~{}\mu_{i}\coloneqq 2(m_{i}-2),~{}\rho_{i}\coloneqq 4-m_{i}.

Let VV be a quadratic space of dimension rr with a basis {e1,,er}\{e_{1},\ldots,e_{r}\} such that the Gram matrix with respect to the basis is a diagonal matrix with entries α1,,αr\alpha_{1},\ldots,\alpha_{r}. The shifted lattice XX corresponding to the sum FF is defined to be

XL+νi=1rμiei+i=1rρiei,X\coloneqq L+\nu\coloneqq\bigoplus_{i=1}^{r}\mathbb{Z}\langle\mu_{i}e_{i}\rangle+\sum_{i=1}^{r}\rho_{i}e_{i},

and the base lattice LXL_{X} is defined to be the lattice generated by the basis {e1,,er}\{e_{1},\ldots,e_{r}\}. It is straightforward to verify that (4.3) holds for the shifted lattice XX.

In this subsection, we are going to bound the Fourier coefficient aEX(μn+ρ)a_{E_{X}}(\mu n+\rho) 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) βp(μn+ρ;X)=pordp(iμi)Ip(8Λn;4Λϕ)=pordp(iμi)+ordp(4Λ)Ip(2n;ϕ),\beta_{p}(\mu n+\rho;X)=p^{-\text{ord}_{p}\left(\prod_{i}\mu_{i}\right)}I_{p}(8\Lambda n;4\Lambda\phi)=p^{-\text{ord}_{p}\left(\prod_{i}\mu_{i}\right)+\text{ord}_{p}(4\Lambda)}I_{p}(2n;\phi),

where the quadratic polynomial is given by

(4.5) ϕ(𝐱)i=1r(ai(mi2)xi2ai(mi4)xi).\phi(\mathbf{x})\coloneqq\sum_{i=1}^{r}(a_{i}(m_{i}-2)x_{i}^{2}-a_{i}(m_{i}-4)x_{i}).

It is relatively easy to bound the product of the local densities for p2iai(mi2)p\nmid 2\prod_{i}a_{i}(m_{i}-2).

Proposition 4.8.

Suppose that F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} is a sum of generalized polygonal numbers with r5r\geq 5. Let X,μ,ρX,\mu,\rho be the shifted lattice and the integers constructed in Definition 4.7 corresponding to FF. Then we have

p2iai(mi2)βp(μn+ρ;X)1,\prod_{p\nmid 2\prod_{i}a_{i}(m_{i}-2)}\beta_{p}(\mu n+\rho;X)\sim 1,

where a(n)b(n)a(n)\sim b(n) means b(n)a(n)b(n)b(n)\ll a(n)\ll b(n) for any functions a,b:a,b\colon\mathbb{N}\to\mathbb{R}.

Proof.

Suppose that pp is a prime number such that p2iai(mi2)p\nmid 2\prod_{i}a_{i}(m_{i}-2). Then we can eliminate the linear terms in the quadratic polynomial ϕ\phi 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

ϕ(𝐱)=x12++xr12+Dxr2\phi(\mathbf{x})=x_{1}^{2}+\cdots+x_{r-1}^{2}+Dx_{r}^{2}

where Diai(mi2)p×D\coloneqq\prod_{i}a_{i}(m_{i}-2)\in\mathbb{Z}_{p}^{\times}. Apply Theorem 4.2 to Ip(2n;ϕ)I_{p}(2n;\phi), we have ti=0t_{i}=0 and iNpi\in N_{p} for every 1ir1\leq i\leq r and τp(t)=(1r2)t\tau_{p}(t)=\left(1-\frac{r}{2}\right)t for every integer t1t\geq 1. Hence it follows from a straightforward calculation that

1p2Ip(2n;ϕ)1+p2,1-p^{-2}\leq I_{p}(2n;\phi)\leq 1+p^{-2},

provided that r5r\geq 5. Finally, bounding against special values of the zeta function, we obtain the desired asymptotics. ∎

It remains to bound the local densities for prime numbers pp such that p2iai(mi2)p\mid 2\prod_{i}a_{i}(m_{i}-2).

Proposition 4.9.

Fix an odd prime number pp. Suppose that F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} is a node of depth r5r\geq 5 in the escalator tree TT_{\infty}. Let X,μ,ρX,\mu,\rho be the shifted lattice and the integers constructed in Definition 4.7 corresponding to FF. For any pp-adic integer npn\in\mathbb{Z}_{p}, we have

2(1p12)pmax(ordp(ai))βp(μn+ρ;X)pordp(Λ)ordp(i(mi2))3,2(1-p^{-\frac{1}{2}})p^{-\max(\text{ord}_{p}(a_{i}))}\leq\frac{\beta_{p}(\mu n+\rho;X)}{p^{\text{ord}_{p}(\Lambda)-\text{ord}_{p}\left(\prod_{i}(m_{i}-2)\right)}}\leq 3,

where Λlcm(m12,,mr2)\Lambda\coloneqq\text{lcm}(m_{1}-2,\ldots,m_{r}-2).

Proof.

We apply Theorem 4.2 to Ip(2n;ϕ)I_{p}(2n;\phi) with ϕ\phi defined in (4.5). Following the notation in Theorem 4.2, we have ti=ordp(ai)t_{i}=\text{ord}_{p}(a_{i}) for any 1ir1\leq i\leq r. Moreover, we have iDpi\in D_{p} if p(mi2)p\mid(m_{i}-2) and iNpi\in N_{p} if p(mi2)p\nmid(m_{i}-2). Reindexing by a permutation λSr\lambda\in S_{r}, we may assume that tλ(1)tλ(2)tλ(r)t_{\lambda(1)}\leq t_{\lambda(2)}\leq\cdots\leq t_{\lambda(r)} and λ(i)λ(j)\lambda(i)\leq\lambda(j) if tλ(i)=tλ(j)t_{\lambda(i)}=t_{\lambda(j)} for any 1ijr1\leq i\leq j\leq r. Now, we show that one of the following condition

  1. (1)

    𝔱d=0\mathfrak{t}_{d}=0;

  2. (2)

    𝔱d1\mathfrak{t}_{d}\geq 1, tλ(1)=tλ(2)=tλ(3)=0t_{\lambda(1)}=t_{\lambda(2)}=t_{\lambda(3)}=0;

  3. (3)

    𝔱d=1\mathfrak{t}_{d}=1, tλ(1)=tλ(2)=0,tλ(3)=1t_{\lambda(1)}=t_{\lambda(2)}=0,t_{\lambda(3)}=1;

  4. (4)

    𝔱d2\mathfrak{t}_{d}\geq 2, tλ(1)=tλ(2)=0,tλ(3)=tλ(4)=1t_{\lambda(1)}=t_{\lambda(2)}=0,t_{\lambda(3)}=t_{\lambda(4)}=1;

  5. (5)

    𝔱d2\mathfrak{t}_{d}\geq 2, tλ(1)=tλ(2)=0,tλ(3)=1,tλ(4)2t_{\lambda(1)}=t_{\lambda(2)}=0,t_{\lambda(3)}=1,t_{\lambda(4)}\geq 2, (aλ(1)aλ(2)(mλ(1)2)(mλ(2)2)/p)=1(-a_{\lambda(1)}a_{\lambda(2)}(m_{\lambda(1)}-2)(m_{\lambda(2)}-2)/p)=1;

  6. (6)

    𝔱d2\mathfrak{t}_{d}\geq 2, tλ(1)=tλ(2)=0,tλ(3)=2t_{\lambda(1)}=t_{\lambda(2)}=0,t_{\lambda(3)}=2, (aλ(1)aλ(2)(mλ(1)2)(mλ(2)2)/p)=1(-a_{\lambda(1)}a_{\lambda(2)}(m_{\lambda(1)}-2)(m_{\lambda(2)}-2)/p)=1.

Clearly 𝔱d0\mathfrak{t}_{d}\geq 0 by definition. If 𝔱d0\mathfrak{t}_{d}\geq 0, then condition (1) holds. So we may assume that 𝔱d1\mathfrak{t}_{d}\geq 1. In this case, we have pa1a2p\nmid a_{1}a_{2}. Indeed, by Table 2.1, we have pa1p\nmid a_{1}, and while if pa2p\mid a_{2}, then p=3p=3 and m1=5m_{1}=5. Therefore pa2p\mid a_{2} implies that 𝔱d=0\mathfrak{t}_{d}=0, which is a contradiction. Hence we have pa1a2p\nmid a_{1}a_{2}, and then λ(i)=i\lambda(i)=i and tλ(i)=0t_{\lambda(i)}=0 for any i{1,2}i\in\{1,2\}. If tλ(3)=0t_{\lambda(3)}=0, then condition (2) holds. So we may assume that tλ(3)1t_{\lambda(3)}\geq 1 and in particular pa3p\mid a_{3}. In this case, Table 2.1 implies that 3p193\leq p\leq 19 and ordp(a3)2\text{ord}_{p}(a_{3})\leq 2. Assuming that ordp(a3)=1\text{ord}_{p}(a_{3})=1, then we have λ(3)=3\lambda(3)=3 and tλ(3)=1t_{\lambda(3)}=1. If p(m32)p\mid(m_{3}-2), then condition (3) holds, while if tλ(4)=1t_{\lambda(4)}=1, then either condition (3) or condition (4) holds. Now assume that p(m32)p\nmid(m_{3}-2) and tλ(4)2t_{\lambda(4)}\geq 2. Then 𝔱d2\mathfrak{t}_{d}\geq 2 and by Lemma 2.9, we see that condition (5) holds. So we have exhausted all of these cases with ordp(a3)=1\text{ord}_{p}(a_{3})=1. Assuming that ord3(a3)=2\text{ord}_{3}(a_{3})=2, we see that condition (6) holds by Lemma 2.9.

Next we bound Ip(2n;ϕ)I_{p}(2n;\phi) using Theorem 4.2. If condition (1) holds, then we have Ip(2n;ϕ)=1I_{p}(2n;\phi)=1. If condition (2) holds, we see that λ(i)Np\lambda(i)\in N_{p} for 1i31\leq i\leq 3 and therefore τp(t)t2\tau_{p}(t)\leq-\frac{t}{2} for any integer t1t\geq 1. Thus, we have

|Ip(2n;ϕ)1|(1p1)1t𝔱npt2+|ωp|p𝔱n+12(1p1)1t𝔱n+1pt2p12+p1.|I_{p}(2n;\phi)-1|\leq\left(1-p^{-1}\right)\sum_{1\leq t\leq\mathfrak{t}_{n}}p^{-\frac{t}{2}}+|\omega_{p}|\cdot p^{-\frac{\mathfrak{t}_{n}+1}{2}}\leq\left(1-p^{-1}\right)\sum_{1\leq t\leq\mathfrak{t}_{n}+1}p^{-\frac{t}{2}}\leq p^{-\frac{1}{2}}+p^{-1}.

Suppose that condition (3) holds next. If 𝔱n=0\mathfrak{t}_{n}=0, then we have 1p1Ip(2n;ϕ)1+p11-p^{-1}\leq I_{p}(2n;\phi)\leq 1+p^{-1}. If 𝔱n1\mathfrak{t}_{n}\geq 1, then we have |Ip(2n;ϕ)1|1p1|I_{p}(2n;\phi)-1|\leq 1-p^{-1}.

If condition (4) holds, we have λ(i)Np\lambda(i)\in N_{p} for 1i41\leq i\leq 4. It follows that τp(t)1t\tau_{p}(t)\leq 1-t for any integer t1t\geq 1. There are four different cases we need to consider separately. First, if 𝔱d\mathfrak{t}_{d}\neq\infty and 𝔱n<𝔱d\mathfrak{t}_{n}<\mathfrak{t}_{d}, then we have

|Ip(2n;ϕ)1|(1p1)1t𝔱np1t+|ωp|p𝔱n1p𝔱n+11p𝔱d.|I_{p}(2n;\phi)-1|\leq\left(1-p^{-1}\right)\sum_{1\leq t\leq\mathfrak{t}_{n}}p^{1-t}+|\omega_{p}|\cdot p^{-\mathfrak{t}_{n}}\leq 1-p^{-\mathfrak{t}_{n}+1}\leq 1-p^{-\mathfrak{t}_{d}}.

Second, if 𝔱d\mathfrak{t}_{d}\neq\infty and 𝔱n𝔱d\mathfrak{t}_{n}\geq\mathfrak{t}_{d}, then the first sum runs over 1t𝔱d1\leq t\leq\mathfrak{t}_{d} and ωp=0\omega_{p}=0. Similar arguments yield that |Ip(2n;ϕ)1|1p𝔱d|I_{p}(2n;\phi)-1|\leq 1-p^{-\mathfrak{t}_{d}}. Third, if 𝔱d=\mathfrak{t}_{d}=\infty and 𝔱ntλ(5)\mathfrak{t}_{n}\leq t_{\lambda(5)}, then we have

|Ip(2n;ϕ)1|(1p1)1t𝔱np1t+p12𝔱n=1(1p12)p𝔱n1(1p12)ptλ(5).\displaystyle|I_{p}(2n;\phi)-1|\leq\left(1-p^{-1}\right)\sum_{1\leq t\leq\mathfrak{t}_{n}}p^{1-t}+p^{-\frac{1}{2}-\mathfrak{t}_{n}}=1-(1-p^{-\frac{1}{2}})p^{-\mathfrak{t}_{n}}\leq 1-(1-p^{-\frac{1}{2}})p^{-t_{\lambda(5)}}.

Lastly, if 𝔱d=\mathfrak{t}_{d}=\infty and 𝔱ntλ(5)+1\mathfrak{t}_{n}\geq t_{\lambda(5)}+1, then we have τp(t)12t\tau_{p}(t)\leq\frac{1}{2}-t for any integer ttλ(5)+1t\geq t_{\lambda(5)}+1. Then, we have

|Ip(2n;ϕ)1|\displaystyle|I_{p}(2n;\phi)-1| (1p1)(1ttλ(5)p1t+tλ(5)+1t𝔱np12t)+p𝔱n1\displaystyle\leq\left(1-p^{-1}\right)\bigg{(}\sum_{1\leq t\leq t_{\lambda(5)}}p^{1-t}+\sum_{t_{\lambda(5)}+1\leq t\leq\mathfrak{t}_{n}}p^{\frac{1}{2}-t}\bigg{)}+p^{-\mathfrak{t}_{n}-1}
=1ptλ(5)+ptλ(5)12p𝔱n12+p𝔱n11(1p12)ptλ(5).\displaystyle=1-p^{-t_{\lambda(5)}}+p^{-t_{\lambda(5)}-\frac{1}{2}}-p^{-\mathfrak{t}_{n}-\frac{1}{2}}+p^{-\mathfrak{t}_{n}-1}\leq 1-\left(1-p^{-\frac{1}{2}}\right)p^{-t_{\lambda(5)}}.

If condition (5) holds, then we see that δp(1)=1\delta_{p}(1)=1. If 𝔱n=0\mathfrak{t}_{n}=0, then we have Ip(2n,ϕ)=1p1I_{p}(2n,\phi)=1-p^{-1}. If 𝔱n1\mathfrak{t}_{n}\geq 1, then we have τp(t)1t2\tau_{p}(t)\leq\frac{1-t}{2} for any integer t1t\geq 1 and we conclude that

|Ip(2n;ϕ)2+p1|\displaystyle\left|I_{p}(2n;\phi)-2+p^{-1}\right| (1p1)2t𝔱np1t2+p1+𝔱n2\displaystyle\leq\left(1-p^{-1}\right)\sum_{2\leq t\leq\mathfrak{t}_{n}}p^{\frac{1-t}{2}}+p^{-\frac{1+\mathfrak{t}_{n}}{2}}
=p12(1+p12)(1p𝔱n2)+p1+𝔱n2p12+p1.\displaystyle=p^{-\frac{1}{2}}\left(1+p^{-\frac{1}{2}}\right)\left(1-p^{-\frac{\mathfrak{t}_{n}}{2}}\right)+p^{-\frac{1+\mathfrak{t}_{n}}{2}}\leq p^{-\frac{1}{2}}+p^{-1}.

Finally we suppose that condition (6) holds. Then we have δp(1)=δp(2)=1\delta_{p}(1)=\delta_{p}(2)=1. If 𝔱n=0\mathfrak{t}_{n}=0, then we have Ip(2n;ϕ)=1p1I_{p}(2n;\phi)=1-p^{-1}. If 𝔱n=1\mathfrak{t}_{n}=1, then we have |Ip(2n;ϕ)1|=1p1|I_{p}(2n;\phi)-1|=1-p^{-1}. If 𝔱n2\mathfrak{t}_{n}\geq 2 and 𝔱d=2\mathfrak{t}_{d}=2, then we have Ip(2n;ϕ)=32p1I_{p}(2n;\phi)=3-2p^{-1}. If 𝔱n2\mathfrak{t}_{n}\geq 2 and 𝔱d3\mathfrak{t}_{d}\geq 3, then τp(t)1t2\tau_{p}(t)\leq 1-\frac{t}{2} for t3t\geq 3 and we have

|Ip(2n;ϕ)3+2p1|\displaystyle\left|I_{p}(2n;\phi)-3+2p^{-1}\right| (1p1)3t𝔱np1t2+p𝔱n2\displaystyle\leq\left(1-p^{-1}\right)\sum_{3\leq t\leq\mathfrak{t}_{n}}p^{1-\frac{t}{2}}+p^{-\frac{\mathfrak{t}_{n}}{2}}
=(p12+p1)(1p1𝔱n2)+p𝔱n2p12+p1.\displaystyle=\left(p^{-\frac{1}{2}}+p^{-1}\right)\left(1-p^{1-\frac{\mathfrak{t}_{n}}{2}}\right)+p^{-\frac{\mathfrak{t}_{n}}{2}}\leq p^{-\frac{1}{2}}+p^{-1}.

Combining the calculations with (4.4), we obtain the desired results. ∎

Proposition 4.10.

Suppose that F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} is a node of depth r5r\geq 5 in the escalator tree TT_{\infty}. Let X,μ,ρX,\mu,\rho be the shifted lattice and the integers constructed in Definition 4.7 corresponding to FF. For any 22-adic integer n2n\in\mathbb{Z}_{2}, we have

2max(ord2(ai))1β2(μn+ρ;X)2ord2(Λ)ordp(i2(mi2))5,2^{-\max(\text{ord}_{2}(a_{i}))-1}\leq\frac{\beta_{2}(\mu n+\rho;X)}{2^{\text{ord}_{2}(\Lambda)-\text{ord}_{p}\left(\prod_{i}2(m_{i}-2)\right)}}\leq 5,

where Λlcm(m12,,mr2)\Lambda\coloneqq\text{lcm}(m_{1}-2,\ldots,m_{r}-2).

Proof.

We apply Theorem 4.6 to I2(2n;ϕ)I_{2}(2n;\phi) with ϕ\phi defined in (4.5). Follow the notation in Theorem 4.6. We have ti=ord2(ai)+εt_{i}=\text{ord}_{2}(a_{i})+\varepsilon for any 1ir1\leq i\leq r where

ε{1,if iD2N2;0,if iE2.\varepsilon\coloneqq\begin{dcases}1,&\text{if }i\in D_{2}\cup N_{2};\\ 0,&\text{if }i\in E_{2}.\end{dcases}

Moreover, we have iD2i\in D_{2} if mi2 (mod 4)m_{i}\equiv 2\text{~{}(mod }4\text{)}, iE2i\in E_{2} if mi1 (mod 2)m_{i}\equiv 1\text{~{}(mod }2\text{)}, and iN2i\in N_{2} if mi0 (mod 4)m_{i}\equiv 0\text{~{}(mod }4\text{)}. Reindexing by a permutation λSr\lambda\in S_{r}, we may assume that tλ(1)tλ(2)tλ(r)t_{\lambda(1)}\leq t_{\lambda(2)}\leq\cdots\leq t_{\lambda(r)} and λ(i)λ(j)\lambda(i)\leq\lambda(j) if tλ(i)=tλ(j)t_{\lambda(i)}=t_{\lambda(j)} for any 1ijr1\leq i\leq j\leq r. Now, we show that one of the following condition:

  1. (1)

    𝔱d2\mathfrak{t}_{d}\leq 2;

  2. (2)

    𝔱d3\mathfrak{t}_{d}\geq 3, tλ(1)=tλ(2)=1,1tλ(3)2t_{\lambda(1)}=t_{\lambda(2)}=1,1\leq t_{\lambda(3)}\leq 2, tλ(3)tλ(4)4t_{\lambda(3)}\leq t_{\lambda(4)}\leq 4, λ(3)N2\lambda(3)\in N_{2}, if tλ(4)=4t_{\lambda(4)}=4 then λ(4)D2N2\lambda(4)\in D_{2}\cup N_{2}, if tλ(4)3t_{\lambda(4)}\geq 3 then tλ(3)=2t_{\lambda(3)}=2;

  3. (3)

    𝔱d=3\mathfrak{t}_{d}=3, tλ(1)=1,tλ(2)=2t_{\lambda(1)}=1,t_{\lambda(2)}=2, λ(2)N2\lambda(2)\in N_{2};

  4. (4)

    𝔱d4\mathfrak{t}_{d}\geq 4, tλ(1)=1,tλ(2)=2,tλ(3)=2,2tλ(4)3t_{\lambda(1)}=1,t_{\lambda(2)}=2,t_{\lambda(3)}=2,2\leq t_{\lambda(4)}\leq 3, λ(4)N2\lambda(4)\in N_{2};

  5. (5)

    𝔱d4\mathfrak{t}_{d}\geq 4, tλ(1)=1,tλ(2)=2,tλ(3)=3,3tλ(4)4t_{\lambda(1)}=1,t_{\lambda(2)}=2,t_{\lambda(3)}=3,3\leq t_{\lambda(4)}\leq 4, λ(3)N2\lambda(3)\in N_{2};

  6. (6)

    𝔱d3\mathfrak{t}_{d}\geq 3, tλ(1)=tλ(2)=tλ(3)=1,3tλ(4)6t_{\lambda(1)}=t_{\lambda(2)}=t_{\lambda(3)}=1,3\leq t_{\lambda(4)}\leq 6, if tλ(4)=6t_{\lambda(4)}=6 then λ(4)D2N2\lambda(4)\in D_{2}\cup N_{2}, λ(i)=i\lambda(i)=i for 1i31\leq i\leq 3, (a1,a2,a3)=(1,1,1),(1,1,3)(a_{1},a_{2},a_{3})=(1,1,1),(1,1,3);

  7. (7)

    𝔱d3\mathfrak{t}_{d}\geq 3, tλ(1)=tλ(2)=1,tλ(3)=2,4tλ(4)7t_{\lambda(1)}=t_{\lambda(2)}=1,t_{\lambda(3)}=2,4\leq t_{\lambda(4)}\leq 7, if tλ(4)=4t_{\lambda(4)}=4 then λ(4)E2\lambda(4)\in E_{2}, if tλ(4)=7t_{\lambda(4)}=7 then λ(4)D2N2\lambda(4)\in D_{2}\cup N_{2}, λ(i)=i\lambda(i)=i for 1i31\leq i\leq 3, (a1,a2,a3)=(1,1,2)(a_{1},a_{2},a_{3})=(1,1,2).

If 𝔱d2\mathfrak{t}_{d}\leq 2, then condition (1) holds. Suppose that 𝔱d3\mathfrak{t}_{d}\geq 3 next. From this assumption, we have tλ(1)=1t_{\lambda(1)}=1 and λ(1)=1\lambda(1)=1. Moreover, we have 1a221\leq a_{2}\leq 2 from Table 2.1. Then it follows that m1,m20 (mod 4)m_{1},m_{2}\equiv 0\text{~{}(mod }4\text{)}. We first deal with the case a2=2a_{2}=2. Since t2=2t_{2}=2, we have tλ(2){1,2}t_{\lambda(2)}\in\{1,2\}, depending on whether there exists 3iN23\leq i\in N_{2} with aia_{i} odd, in which case tλ(2)=1t_{\lambda(2)}=1. If tλ(2)=1t_{\lambda(2)}=1, we can conclude that λ(2)2\lambda(2)\neq 2. Since 2a352\leq a_{3}\leq 5 by Table 2.1, λ(2)=3\lambda(2)=3 when a3=3,5a_{3}=3,5. If λ(2)=3\lambda(2)=3, we have a3a412a_{3}\leq a_{4}\leq 12 by Lemma 2.10. Thus, we have 1tλ(3)21\leq t_{\lambda(3)}\leq 2, λ(3)N2\lambda(3)\in N_{2}, and tλ(3)tλ(4)4t_{\lambda(3)}\leq t_{\lambda(4)}\leq 4 and if tλ(4)=4t_{\lambda(4)}=4 then λ(4)D2N2\lambda(4)\in D_{2}\cup N_{2}. Thus, condition (2) holds. If λ(2)3\lambda(2)\neq 3, then a3=2,4a_{3}=2,4. In this case, condition (2) also holds.

If tλ(2)=2t_{\lambda(2)}=2, then λ(2)=2\lambda(2)=2 and λ(2)N2\lambda(2)\in N_{2} by our choice of λ\lambda. If 𝔱d=3\mathfrak{t}_{d}=3, then condition (3) holds. If 𝔱d4\mathfrak{t}_{d}\geq 4, we shall show that condition (4) or condition (5) holds. Since m1,m20 (mod 4)m_{1},m_{2}\equiv 0\text{~{}(mod }4\text{)}, we have a3=2,4a_{3}=2,4 from Table 2.1 and from our assumption tλ(2)=2t_{\lambda(2)}=2. If a3=2a_{3}=2, we have tλ(3)=2t_{\lambda(3)}=2, λ(3)=3\lambda(3)=3, and m30 (mod 4)m_{3}\equiv 0\text{~{}(mod }4\text{)}. By Lemma 2.10, we see that 2a472\leq a_{4}\leq 7. So, we have 2tλ(4)32\leq t_{\lambda(4)}\leq 3 and λ(4)N2\lambda(4)\in N_{2}. Therefore condition (4) holds. If a3=4a_{3}=4, we have m30 (mod 4)m_{3}\equiv 0\text{~{}(mod }4\text{)} since 𝔱d4\mathfrak{t}_{d}\geq 4. It follows that 2tλ(3)32\leq t_{\lambda(3)}\leq 3. If tλ(3)=2t_{\lambda(3)}=2, then condition (4) holds. If tλ(3)=3t_{\lambda(3)}=3, we see that λ(3)=3\lambda(3)=3 and λ(3)N2\lambda(3)\in N_{2} by our choice of λ\lambda. Moreover, by Lemma 2.10, we have 4a4154\leq a_{4}\leq 15. Therefore we have 3tλ(4)43\leq t_{\lambda(4)}\leq 4. This shows that condition (5) holds.

Next we deal with the case a2=1a_{2}=1. Then tλ(2)=1t_{\lambda(2)}=1 and λ(2)=2\lambda(2)=2. From Table 2.1, we notice that a33a_{3}\leq 3. Thus we have 1tλ(3)21\leq t_{\lambda(3)}\leq 2 and λ(3)N2\lambda(3)\in N_{2} by our choice of λ\lambda. If tλ(3)=1t_{\lambda(3)}=1, then Lemma 2.10 (1) and (3) implies that tλ(3)tλ(4)6t_{\lambda(3)}\leq t_{\lambda(4)}\leq 6, with tλ(4)=6t_{\lambda(4)}=6 only possible if λ(4)=4D2N2\lambda(4)=4\in D_{2}\cup N_{2} and a4=32a_{4}=32, giving condition (6) in that case. If tλ(4)2t_{\lambda(4)}\leq 2, then condition (2) holds. If 3tλ(4)63\leq t_{\lambda(4)}\leq 6, then condition (6) holds.

Now suppose that tλ(3)=2t_{\lambda(3)}=2. By Lemma 2.10, we have 2=tλ(3)tλ(4)72=t_{\lambda(3)}\leq t_{\lambda(4)}\leq 7, with tλ(4)=7t_{\lambda(4)}=7 only possible if t4=64t_{4}=64 and λ(4)=4D2N2\lambda(4)=4\in D_{2}\cup N_{2}, giving condition (7) in that case. If 2tλ(4)32\leq t_{\lambda(4)}\leq 3, then condition (2) holds. If 5tλ(4)75\leq t_{\lambda(4)}\leq 7, then condition (7) holds. Finally, if tλ(4)=4t_{\lambda(4)}=4, then either condition (2) holds or condition (7) holds, depending on whether λ(4)D2N2\lambda(4)\in D_{2}\cup N_{2} or λ(4)E2\lambda(4)\in E_{2}.

Then we bound I2(2n;ϕ)I_{2}(2n;\phi) using the explicit formula from Theorem 4.6. If 𝔱d=0\mathfrak{t}_{d}=0, then I2(2n;ϕ)=1I_{2}(2n;\phi)=1. Otherwise, we have I2(2n;ϕ)=2+R2(2n;ϕ)I_{2}(2n;\phi)=2+R_{2}(2n;\phi), where we define

R2(2n;ϕ)2tmin(𝔱d,𝔱n+3)tti+1 for iN2δ2(t)2τ2(t)1.R_{2}(2n;\phi)\coloneqq\sum_{\begin{subarray}{c}2\leq t\leq\min(\mathfrak{t}_{d},\mathfrak{t}_{n}+3)\\ t\neq t_{i}+1\text{ for }i\in N_{2}\end{subarray}}\delta_{2}(t)2^{\tau_{2}(t)-1}.

If condition (1) holds, then R2(2n;ϕ)=0R_{2}(2n;\phi)=0. Suppose that condition (2) holds. If λ(4)D2E2\lambda(4)\in D_{2}\cup E_{2}, then |R2(2n;ϕ)|1|R_{2}(2n;\phi)|\leq 1. If λ(4)N2\lambda(4)\in N_{2} and tλ(4)tλ(3)+1t_{\lambda(4)}\leq t_{\lambda(3)}+1, we have δ2(t)=0\delta_{2}(t)=0 for 2ttλ(4)+12\leq t\leq t_{\lambda(4)}+1 and τ2(t)1t+tλ(3)+tλ(4)+42\tau_{2}(t)-1\leq-t+\frac{t_{\lambda(3)}+t_{\lambda(4)}+4}{2} for ttλ(4)+2t\geq t_{\lambda(4)}+2. If λ(5)D2E2\lambda(5)\in D_{2}\cup E_{2}, then

|R2(2n;ϕ)|tλ(4)+2ttλ(5)+12τ2(t)122tλ(5)+tλ(3)+tλ(4)+22.|R_{2}(2n;\phi)|\leq\sum_{t_{\lambda(4)}+2\leq t\leq t_{\lambda(5)}+1}2^{\tau_{2}(t)-1}\leq 2-2^{-t_{\lambda(5)}+\frac{t_{\lambda(3)}+t_{\lambda(4)}+2}{2}}.

If λ(5)N2\lambda(5)\in N_{2}, then τ2(t)13t2+tλ(3)+tλ(4)+tλ(5)+52\tau_{2}(t)-1\leq-\frac{3t}{2}+\frac{t_{\lambda(3)}+t_{\lambda(4)}+t_{\lambda(5)}+5}{2} for ttλ(5)+2t\geq t_{\lambda(5)}+2. We have

|R2(2n;ϕ)|\displaystyle|R_{2}(2n;\phi)| tλ(4)+2ttλ(5)2τ2(t)1+tλ(5)+2t2τ2(t)1\displaystyle\leq\sum_{t_{\lambda(4)}+2\leq t\leq t_{\lambda(5)}}2^{\tau_{2}(t)-1}+\sum_{t_{\lambda(5)}+2\leq t\leq\infty}2^{\tau_{2}(t)-1}
max(0,22tλ(5)+tλ(3)+tλ(4)+42)+2tλ(5)+tλ(3)+tλ(4)+12\displaystyle\leq\max\left(0,2-2^{-t_{\lambda(5)}+\frac{t_{\lambda(3)}+t_{\lambda(4)}+4}{2}}\right)+2^{-t_{\lambda(5)}+\frac{t_{\lambda(3)}+t_{\lambda(4)}+1}{2}}
22tλ(5)+tλ(3)+tλ(4)+22.\displaystyle\leq 2-2^{-t_{\lambda(5)}+\frac{t_{\lambda(3)}+t_{\lambda(4)}+2}{2}}.

Similarly if λ(4)N2\lambda(4)\in N_{2} and tλ(4)=tλ(3)+2t_{\lambda(4)}=t_{\lambda(3)}+2 then tλ(3)=2t_{\lambda(3)}=2 and tλ(4)=4t_{\lambda(4)}=4. Then, we see that,

|R2(2n;ϕ)||δ2(4)|2τ2(4)1+tλ(4)+2ttλ(5)+12τ2(t)122tλ(5)+tλ(4)+42.|R_{2}(2n;\phi)|\leq|\delta_{2}(4)|\cdot 2^{\tau_{2}(4)-1}+\sum_{t_{\lambda(4)}+2\leq t\leq t_{\lambda(5)}+1}2^{\tau_{2}(t)-1}\leq 2-2^{-t_{\lambda(5)}+\frac{t_{\lambda(4)}+4}{2}}.

If condition (3) holds, we have R2(2n;ϕ)=0R_{2}(2n;\phi)=0. If condition (4) or condition (5) holds, arguing as the case of condition (2), we see that

|R2(2n;ϕ)|22tλ(5)+tλ(4)+42.|R_{2}(2n;\phi)|\leq 2-2^{-t_{\lambda(5)}+\frac{t_{\lambda(4)}+4}{2}}.

Suppose that condition (6) holds. If tλ(4)=3t_{\lambda(4)}=3 with λ(4)D2\lambda(4)\in D_{2}, we have |R2(2n;ϕ)|1|R_{2}(2n;\phi)|\leq 1. If tλ(4)=4t_{\lambda(4)}=4 with λ(4)D2\lambda(4)\in D_{2} or tλ(4)=3t_{\lambda(4)}=3 with λ(4)E2\lambda(4)\in E_{2}, then

R2(2n;ϕ)=δ2(3)2τ2(3)1+ϵδ2(4)2τ2(4)1,R_{2}(2n;\phi)=\delta_{2}(3)2^{\tau_{2}(3)-1}+\epsilon\cdot\delta_{2}(4)2^{\tau_{2}(4)-1},

where ϵ0\epsilon\coloneqq 0 if ti=3t_{i}=3 and iN2i\in N_{2} for some 1ir1\leq i\leq r and ϵ1\epsilon\coloneqq 1 otherwise. Notice that the right hand side is determined by ϵ\epsilon, mλ(1),mλ(2),mλ(3) (mod 16)m_{\lambda(1)},m_{\lambda(2)},m_{\lambda(3)}\text{~{}(mod }16\text{)} and n (mod 8)n\text{~{}(mod }8\text{)}. By a computer program and by using Lemma 2.10 to get rid of certain choices of mλ(1),mλ(2),mλ(3)m_{\lambda(1)},m_{\lambda(2)},m_{\lambda(3)}, we can conclude that 1R2(2n;ϕ)2-1\leq R_{2}(2n;\phi)\leq 2. If 5tλ(4)65\leq t_{\lambda(4)}\leq 6 with λ(4)D2\lambda(4)\in D_{2} or 4tλ(4)54\leq t_{\lambda(4)}\leq 5 with λ(4)E2\lambda(4)\in E_{2}, we can argue in a similar manner to conclude that 1R2(2n;ϕ)2-1\leq R_{2}(2n;\phi)\leq 2. If λ(4)N2\lambda(4)\in N_{2}, we have τ2(t)1=t+5+tλ(4)2\tau_{2}(t)-1=-t+\frac{5+t_{\lambda(4)}}{2} for any integer ttλ(4)+2t\geq t_{\lambda(4)}+2. Thus, for 3tλ(4)63\leq t_{\lambda(4)}\leq 6, we have

|tλ(4)+2t𝔱d2τ2(t)1|12tλ(5)+tλ(4)+32.\left|\sum_{t_{\lambda(4)}+2\leq t\leq\mathfrak{t}_{d}}2^{\tau_{2}(t)-1}\right|\leq 1-2^{-t_{\lambda(5)}+\frac{t_{\lambda(4)}+3}{2}}.

Combining with the calculation of the cases λ(4)D2E2\lambda(4)\in D_{2}\cup E_{2}, we have

2tλ(5)+tλ(4)+322R2(2n;ϕ)32tλ(5)+tλ(4)+32.2^{-t_{\lambda(5)}+\frac{t_{\lambda(4)}+3}{2}}-2\leq R_{2}(2n;\phi)\leq 3-2^{-t_{\lambda(5)}+\frac{t_{\lambda(4)}+3}{2}}.

The case of condition (7) is similar to that of condition (6). Using Lemma 2.10, we can prove that 32R2(2n;ϕ)52-\frac{3}{2}\leq R_{2}(2n;\phi)\leq\frac{5}{2}. 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 F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} is a node of depth r5r\geq 5 in the escalator tree TT_{\infty}. Let X,μ,ρX,\mu,\rho be the shifted lattice and the integers constructed in Definition 4.7 corresponding to FF. For any real number ε>0\varepsilon>0, we have

aEX(μn+ρ)r(μn+ρ)r21Λr1+ε(iai)32+ε,a_{E_{X}}(\mu n+\rho)\gg_{r}\frac{(\mu n+\rho)^{\frac{r}{2}-1}}{\Lambda^{r-1+\varepsilon}\cdot\Big{(}\prod_{i}a_{i}\Big{)}^{\frac{3}{2}+\varepsilon}},

for any integer n0n\geq 0, where Λlcm(m12,,mr2)\Lambda\coloneqq\text{lcm}(m_{1}-2,\ldots,m_{r}-2).

Proof.

By our construction of the base lattice LXL_{X} in Definition 4.7, we have

[LX#:LX]rΛriai(mi2)1.[L_{X}^{\#}\colon L_{X}]\ll_{r}\Lambda^{r}\prod_{i}a_{i}(m_{i}-2)^{-1}.

Plugging this upper bound and Proposition 4.8 into (3.2), we have

aEX(μn+ρ)r(μn+ρ)r21Λr1(iai)12piai(mi2)βp(μn+ρ;X).a_{E_{X}}(\mu n+\rho)\gg_{r}\frac{(\mu n+\rho)^{\frac{r}{2}-1}}{\Lambda^{r-1}\cdot\Big{(}\prod_{i}a_{i}\Big{)}^{\frac{1}{2}}}\prod_{p\mid\prod_{i}a_{i}(m_{i}-2)}\beta_{p}(\mu n+\rho;X).

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 aGX(μn+ρ)a_{G_{X}}(\mu n+\rho) of the cuspidal part GXG_{X} appearing in (4.3). This requires the theory of modular forms. We will use the following congruence subgroups of SL2()\text{SL}_{2}(\mathbb{Z}):

Γ0(N)={(abcd)SL2()|c0 (mod N)},\Gamma_{0}(N)=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{SL}_{2}(\mathbb{Z})\Biggm{|}c\equiv 0\text{~{}(mod }N\text{)}\right\},
Γ1(N)={(abcd)Γ0(N)|a1 (mod N)},\Gamma_{1}(N)=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma_{0}(N)\Biggm{|}a\equiv 1\text{~{}(mod }N\text{)}\right\},

and

Γ(N)={(abcd)Γ1(N)|b0 (mod N)}.\Gamma(N)=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma_{1}(N)\Biggm{|}b\equiv 0\text{~{}(mod }N\text{)}\right\}.

Fix k12k\in\frac{1}{2}\mathbb{Z}. Assume that Γ\Gamma is a congruence subgroup such that ΓΓ0(4)\Gamma\subseteq\Gamma_{0}(4) when k12k\in\frac{1}{2}\mathbb{Z}-\mathbb{Z} and χ\chi is a Dirichlet character. Letting θ(τ)ne2πin2τ\theta(\tau)\coloneqq\sum_{n\in\mathbb{Z}}e^{2\pi in^{2}\tau} be the standard unary theta function of weight 12\frac{1}{2}, we say that a function f:f:\mathbb{H}\to\mathbb{C} is a (holomorphic) modular form of weight kk and Nebentypus character χ\chi (for the θ\theta-multiplier) on Γ\Gamma if for every γ=(abcd)Γ\gamma=\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in\Gamma we have

f(γτ)=χ(d)(θ(γτ)θ(τ))2kf(τ)f(\gamma\tau)=\chi(d)\left(\frac{\theta(\gamma\tau)}{\theta(\tau)}\right)^{2k}f(\tau)

and f(γτ)f(\gamma\tau) grows at most polynomially in Im(τ)\operatorname{Im}(\tau) as τi\tau\to i\infty; we furthermore call ff a cusp form if f(γτ)f(\gamma\tau) vanishes as τi\tau\to i\infty for all γΓ\gamma\in\Gamma. We let Mk(Γ,χ)M_{k}(\Gamma,\chi) denote the space of modular forms of weight kk for the congruence subgroup Γ\Gamma with the Nebentypus χ\chi and let Sk(Γ,χ)S_{k}(\Gamma,\chi) denote the subspace of cusp forms in Mk(Γ,χ)M_{k}(\Gamma,\chi).

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 XX in a quadratic space VV. Let O(V),O𝔸(V),O(X)O(V),O_{\mathbb{A}}(V),O(X) denote the orthogonal group of VV, the adelization of O(V)O(V), and the subgroup of O(V)O(V) consisting of isometries σ\sigma such that σ(X)=X\sigma(X)=X, respectively. By [30, Lemma 1.2], O(V)O(V) and O𝔸(V)O_{\mathbb{A}}(V) act on the set of shifted lattices in a quadratic space VV. The genus of a shifted lattice XX is the orbit of XX under the action of O𝔸(V)O_{\mathbb{A}}(V), denoted by gen(X)\text{gen}(X). The isometry class of a shifted lattice XX is the orbit of XX under the action of O(V)O(V), denoted by cls(X)\text{cls}(X). By [30, Corollary 2.3], the genus of XX splits into a finite number of isometry classes. So we take a set of representatives X1,,XgX_{1},\ldots,X_{g} of isometry classes in gen(X)\text{gen}(X). The Siegel–Weil formula states that the Fourier coefficient aEX(n)a_{E_{X}}(n) is a weighted sum of the number of representations by shifted lattices in gen(X)\text{gen}(X) as follows,

aEX(n)=m(X)1i=1grXi(n)|O(Xi)|,a_{E_{X}}(n)=m(X)^{-1}\sum_{i=1}^{g}\frac{r_{X_{i}}(n)}{|O(X_{i})|},

where the quantity m(X)m(X) is the mass of XX, defined by

m(X)i=1g1|O(Xi)|.m(X)\coloneqq\sum_{i=1}^{g}\frac{1}{|O(X_{i})|}.

Let XX be a positive-definite integral shifted lattice of rank rr, level NN, and conductor MM. By [23, Proposition 2.1], for any shifted lattice YY in the genus of XX, the theta series ΘY\Theta_{Y} is a modular form in Mr2(Γ0(M2N)Γ1(M),χ)M_{\frac{r}{2}}(\Gamma_{0}(M^{2}N)\cap\Gamma_{1}(M),\chi) for some Dirichlet character χ\chi. Therefore we see that GX=Sr2(Γ0(M2N)Γ1(M),χ)G_{X}=\in S_{\frac{r}{2}}(\Gamma_{0}(M^{2}N)\cap\Gamma_{1}(M),\chi).

We decompose the cuspidal part GXG_{X} by [4, Theorem 2.5] as follows,

(5.1) GX=ψGX,ψ,G_{X}=\sum_{\psi}G_{X,\psi},

where the sum runs over all Dirichlet character ψ\psi modulo MM. For each Dirichlet character ψ\psi modulo MM, we have GX,ψSr2(Γ0(M2N),χψ)G_{X,\psi}\in S_{\frac{r}{2}}(\Gamma_{0}(M^{2}N),\chi\psi). We call GX,ψG_{X,\psi} the ψ\psi-component of the shifted lattice XX. Thus, we reduce to the cases when the cusp forms are on the congruence subgroup Γ0(M2N)\Gamma_{0}(M^{2}N).

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 f,gMk(Γ,χ)f,g\in M_{k}(\Gamma,\chi) such that fgfg is a cusp form, the Petersson inner product f,gΓ\langle f,g\rangle_{\Gamma} is defined as

f,gΓΓ\f(τ)g(τ)¯vkdudvv2,\langle f,g\rangle_{\Gamma}\coloneqq\int_{\Gamma\backslash\mathbb{H}}f(\tau)\overline{g(\tau)}v^{k}\frac{\mathrm{d}u\mathrm{d}v}{v^{2}},

where we write τ=u+iv\tau=u+iv with u,vu,v\in\mathbb{R}. For a cusp form fSk(Γ,χ)f\in S_{k}(\Gamma,\chi), the Petersson norm of ff is defined as

fΓf,fΓ.\lVert f\rVert_{\Gamma}\coloneqq\sqrt{\langle f,f\rangle_{\Gamma}}.

We note that while f,gΓ\langle f,g\rangle_{\Gamma} depends on the choice of Γ\Gamma, the normalization f,gΓ[SL2():Γ]\frac{\langle f,g\rangle_{\Gamma}}{[\text{SL}_{2}(\mathbb{Z})\colon\Gamma]} is independent of the choice of Γ\Gamma; both normalizations appear in the literature and prove useful in different settings, so we keep the dependence on Γ\Gamma in our notation to avoid confusion with the normalization that is independent of Γ\Gamma.

The next proposition reduces the problem of bounding the Fourier coefficients of a cusp form to bounding its Petersson norm.

Proposition 5.1.

Fix k12k\in\frac{1}{2}\mathbb{Z} such that k2k\geq 2. Let N1N\geq 1 be an integer such that 4N4\mid N if k12k\in\frac{1}{2}\mathbb{Z}-\mathbb{Z} and χ\chi be a Dirichlet character modulo NN. Let ff be a cusp form in Sk(Γ0(N),χ)S_{k}(\Gamma_{0}(N),\chi). If k12k\in\frac{1}{2}\mathbb{Z}-\mathbb{Z}, for any real number ε>0\varepsilon>0, we have

|af(n)|ε,kfΓ0(N)Nεnk214+ε((n,N)N)116,|a_{f}(n)|\ll_{\varepsilon,k}\lVert f\rVert_{\Gamma_{0}(N)}N^{\varepsilon}n^{\frac{k}{2}-\frac{1}{4}+\varepsilon}\left(\frac{(n,N)}{N}\right)^{\frac{1}{16}},

for any integer n0n\geq 0. If kk\in\mathbb{Z}, for any real number ε>0\varepsilon>0, we have

|af(n)|ε,kfΓ0(N)N12+εnk212+ε|a_{f}(n)|\ll_{\varepsilon,k}\lVert f\rVert_{\Gamma_{0}(N)}N^{\frac{1}{2}+\varepsilon}n^{\frac{k}{2}-\frac{1}{2}+\varepsilon}

for any integer n0n\geq 0. Moreover, both the implied constants are effective.

Proof.

If kk\in\mathbb{Z}, it is exactly [22, Theorem 12]. If k12k\in\frac{1}{2}\mathbb{Z}-\mathbb{Z}, it follows from [34, Theorem 1]. ∎

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 GG be the ψ\psi-component of the cuspidal part of the theta series associated to an integral positive-definite shifted lattice XX of rank r4r\geq 4, discriminant DD, level NN, and conductor MM. For any real number ε>0\varepsilon>0, we have

GΓ0(M2N)ε,r(M2N)r+1+εM,\lVert G\rVert_{\Gamma_{0}(M^{2}N)}\ll_{\varepsilon,r}\frac{(M^{2}N)^{r+1+\varepsilon}}{M},

where the implied constant is effective.

Proof.

Let \mathcal{F} denote the usual fundamental domain of SL2()\text{SL}_{2}(\mathbb{Z}) acting on the upper half-plane \mathbb{H}. Let {γ1,,γs}\{\gamma_{1},\ldots,\gamma_{s}\} be a set of representatives of the cosets Γ(M2N)\SL2()\Gamma(M^{2}N)\backslash\text{SL}_{2}(\mathbb{Z}). Let Γ\Gamma_{\infty} denote the parabolic subgroup of Γ(M2N)\SL2()\Gamma(M^{2}N)\backslash\text{SL}_{2}(\mathbb{Z}) generated by T(1101)T\coloneqq\left(\begin{smallmatrix}1&1\\ 0&1\end{smallmatrix}\right), namely

Γ{(1001),(1101),,(1M2N101)}.\Gamma_{\infty}\coloneqq\left\{\begin{pmatrix}1&0\\ 0&1\end{pmatrix},\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\ldots,\begin{pmatrix}1&M^{2}N-1\\ 0&1\end{pmatrix}\right\}.

Finally, let {ρ1,,ρt}\{\rho_{1},\ldots,\rho_{t}\} be a set of representative of the double cosets Γ(M2N)\SL2()/Γ\Gamma(M^{2}N)\backslash\text{SL}_{2}(\mathbb{Z})/\Gamma_{\infty}.

Set

CM,N[SL2():Γ0(M2N)][SL2():Γ(M2N)].C_{M,N}\coloneqq\frac{[\text{SL}_{2}(\mathbb{Z})\colon\Gamma_{0}(M^{2}N)]}{[\text{SL}_{2}(\mathbb{Z})\colon\Gamma(M^{2}N)]}.

By definition of the Petersson norm, we have

GΓ0(M2N)2=\displaystyle\lVert G\rVert_{\Gamma_{0}(M^{2}N)}^{2}= CM,NΓ(M2N)\|G(τ)|2vr2dudvv2=CM,Ni=1sγi|G(τ)|2vr2dudvv2\displaystyle~{}C_{M,N}\int_{\Gamma(M^{2}N)\backslash\mathbb{H}}|G(\tau)|^{2}v^{\frac{r}{2}}\frac{\mathrm{d}u\mathrm{d}v}{v^{2}}=C_{M,N}\sum_{i=1}^{s}\int_{\gamma_{i}\mathcal{F}}|G(\tau)|^{2}v^{\frac{r}{2}}\frac{\mathrm{d}u\mathrm{d}v}{v^{2}}
=\displaystyle= CM,Ni=1s|G|kγi(τ)|2vr2dudvv2=CM,Nj=1tγΓγ|G|kρj(τ)|2vr2dudvv2,\displaystyle~{}C_{M,N}\sum_{i=1}^{s}\int_{\mathcal{F}}\Big{|}G|_{k}\gamma_{i}(\tau)\Big{|}^{2}v^{\frac{r}{2}}\frac{\mathrm{d}u\mathrm{d}v}{v^{2}}=C_{M,N}\sum_{j=1}^{t}\int_{\bigcup_{\gamma\in\Gamma_{\infty}}\gamma\mathcal{F}}\Big{|}G|_{k}\rho_{j}(\tau)\Big{|}^{2}v^{\frac{r}{2}}\frac{\mathrm{d}u\mathrm{d}v}{v^{2}},

where kr2k\coloneqq\frac{r}{2} and we denote

f|kγ(τ)(Θ(γτ)Θ(τ))2kf(γτ).f|_{k}\gamma(\tau)\coloneqq\left(\frac{\Theta(\gamma\tau)}{\Theta(\tau)}\right)^{-2k}f(\gamma\tau).

The cusp associated to the matrix ρj\rho_{j} is ρj(i)\rho_{j}(i\infty). Recall that every cusp of Γ(M2N)\Gamma(M^{2}N) has cusp width M2NM^{2}N, so the Fourier expansion of GG at ρj(i)\rho_{j}(i\infty) may be written in the form (for some aj(n)a_{j}(n)\in\mathbb{C})

G|kρj(τ)=n=1aj(n)e2πiτnM2N.G|_{k}\rho_{j}(\tau)=\sum_{n=1}^{\infty}a_{j}(n)e^{\frac{2\pi i\tau n}{M^{2}N}}.

Plugging in the Fourier expansions, we have

γΓγ|G|kρj(τ)|2vr2\displaystyle\int_{\bigcup_{\gamma\in\Gamma_{\infty}}\gamma\mathcal{F}}\Big{|}G|_{k}\rho_{j}(\tau)\Big{|}^{2}v^{\frac{r}{2}} dudvv23212M2N12|G|kρj(τ)|2vr2dudvv2\displaystyle\frac{\mathrm{d}u\mathrm{d}v}{v^{2}}\leq\int_{\frac{\sqrt{3}}{2}}^{\infty}\int_{-\frac{1}{2}}^{M^{2}N-\frac{1}{2}}\Big{|}G|_{k}\rho_{j}(\tau)\Big{|}^{2}v^{\frac{r}{2}}\frac{\mathrm{d}u\mathrm{d}v}{v^{2}}
=n=1m=1aj(n)aj(m)¯32e2π(n+m)vM2Nvr2212M2N12e2πi(nm)uM2Ndudv\displaystyle=\sum_{n=1}^{\infty}\sum_{m=1}^{\infty}a_{j}(n)\overline{a_{j}(m)}\int_{\frac{\sqrt{3}}{2}}^{\infty}e^{-\frac{2\pi(n+m)v}{M^{2}N}}v^{\frac{r}{2}-2}\int_{-\frac{1}{2}}^{M^{2}N-\frac{1}{2}}e^{\frac{2\pi i(n-m)u}{M^{2}N}}\mathrm{d}u\mathrm{d}v
=M2Nn=1|aj(n)|232e4πnvM2Nvr22dv\displaystyle=M^{2}N\sum_{n=1}^{\infty}|a_{j}(n)|^{2}\int_{\frac{\sqrt{3}}{2}}^{\infty}e^{-\frac{4\pi nv}{M^{2}N}}v^{\frac{r}{2}-2}\mathrm{d}v
r(M2N)r2n=1|aj(n)|2n1r2Γ(r21,23πnM2N)\displaystyle\ll_{r}(M^{2}N)^{\frac{r}{2}}\sum_{n=1}^{\infty}|a_{j}(n)|^{2}n^{1-\frac{r}{2}}\Gamma\left(\frac{r}{2}-1,\frac{2\sqrt{3}\pi n}{M^{2}N}\right)
r(M2N)2n=1e23πnM2Nn|aj(n)|2,\displaystyle\ll_{r}(M^{2}N)^{2}\sum_{n=1}^{\infty}\frac{e^{-\frac{2\sqrt{3}\pi n}{M^{2}N}}}{n}|a_{j}(n)|^{2},

where Γ(s,y)\Gamma(s,y) is the incomplte gamma function defined as

Γ(s,y)yetts1dt\Gamma(s,y)\coloneqq\int_{y}^{\infty}e^{-t}t^{s-1}\mathrm{d}t

for a complex variable ss\in\mathbb{C} and a real variable yy\in\mathbb{R} such that (s)>0\real(s)>0 and y>0y>0, and for the last step we use an asympotic upper bound Γ(s,y)seyys1\Gamma(s,y)\ll_{s}e^{-y}y^{s-1} as yy\to\infty. Therefore, it remains to bound the absolute value of the Fourier coefficients aj(n)a_{j}(n) of GG at each cusp.

Then, by [4, Theorem 2.5] and Siegel–Weil formula, we see that the ψ\psi-component GG is constructed explicitly as follows,

G\displaystyle G =1ϕ(M)d (mod M)ψ(d)¯χΔ(d)¯GX|γd\displaystyle=\frac{1}{\phi(M)}\sum_{d\text{~{}(mod }M\text{)}}\overline{\psi(d)}\cdot\overline{\chi_{\Delta}(d)}\cdot G_{X}|\gamma_{d}
=1ϕ(M)d (mod M)ψ(d)¯m(X(d))cls(Y)gen(X(d))ΘX(d)ΘY|O(Y)|,\displaystyle=\frac{1}{\phi(M)}\sum_{d\text{~{}(mod }M\text{)}}\frac{\overline{\psi(d)}}{m(X(d))}\sum_{\text{cls}(Y)\subseteq\text{gen}(X(d))}\frac{\Theta_{X(d)}-\Theta_{Y}}{|O(Y)|},

where γdΓ0(M2N)\gamma_{d}\in\Gamma_{0}(M^{2}N) is a matrix whose lower right entry is congruent to dd modulo MM. Thus, by the triangle inequality and the linearity of the slash operator, it suffices to bound the Fourier coefficient of the theta series ΘY\Theta_{Y} at the cusp ρj(i)\rho_{j}(i\infty) for each class cls(Y)gen(X(d))\text{cls}(Y)\subseteq\text{gen}(X(d)).

For any shifted lattice Y=K+μgen(X(d))Y=K+\mu\in\text{gen}(X(d)), we are going to estimate the Fourier coefficient aj(n)a_{j}(n). There exists a base lattice LYL_{Y} of YY such that YY is of level NN and conductor MM relative to LYL_{Y}. We choose a basis {e1,,er}\{e_{1},\ldots,e_{r}\} of LYL_{Y} such that {μ1e1,,μrer}\{\mu_{1}e_{1},\ldots,\mu_{r}e_{r}\} is a basis of KK for integers μ1,,μr1\mu_{1},\ldots,\mu_{r}\geq 1 and μ=ρ1e1++ρrer\mu=\rho_{1}e_{1}+\cdots+\rho_{r}e_{r} with integers ρ1,,ρr\rho_{1},\ldots,\rho_{r}\in\mathbb{Z}. Let AA be the Hessian matrix of the bilinear form BB with respect to the basis {e1,,er}\{e_{1},\ldots,e_{r}\}. We need a modular transformation formula for the shifted lattice YY. Using Shimura’s formulation of theta series [23, (2.0)] with the trivial spherical function PP, we put

θ(τ;A,h,N)xh (mod N)e2πiτx𝗍Ax2N2,\theta(\tau;A,h,N)\coloneqq\sum_{x\equiv h\text{~{}(mod }N\text{)}}e^{2\pi i\tau\frac{x^{\mathsf{t}}Ax}{2N^{2}}},

where AA is a r×rr\times r real symmetric matrix, hh is a vector of dimension rr, and NN is a positive integer. The theta series ΘY\Theta_{Y} is related to Shimura’s theta series as follows,

ΘY(τ)=\displaystyle\Theta_{Y}(\tau)= xiρimodμie2πiτx𝗍Ax2=h (mod M)hiρi (mod μi)xh (mod M)e2πiτx𝗍Ax2\displaystyle\sum_{x_{i}\equiv\rho_{i}\mod{\mu_{i}}}e^{2\pi i\tau\frac{x^{\mathsf{t}}Ax}{2}}=\sum_{\begin{subarray}{c}h\text{~{}(mod }M\text{)}\\ h_{i}\equiv\rho_{i}\text{~{}(mod }\mu_{i}\text{)}\end{subarray}}\sum_{x\equiv h\text{~{}(mod }M\text{)}}e^{2\pi i\tau\frac{x^{\mathsf{t}}Ax}{2}}
=\displaystyle= h (mod M)hiρi (mod μi)θ(τ;M2A,MNh,M2N).\displaystyle\sum_{\begin{subarray}{c}h\text{~{}(mod }M\text{)}\\ h_{i}\equiv\rho_{i}\text{~{}(mod }\mu_{i}\text{)}\end{subarray}}\theta(\tau;M^{2}A,MNh,M^{2}N).

It is easy to verify that for Shimura’s theta series θ(τ;M2A,MNh,M2N)\theta(\tau;M^{2}A,MNh,M^{2}N), a stronger assumption than that in [23, (2.1)] is satisfied since the matrix M2AM^{2}A is even integral. Under such a stronger assumption, we can imitate the arguments in [23, Pages 454–455] to see that, for any matrix γ=(abcd)SL2()\gamma=\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in\text{SL}_{2}(\mathbb{Z}) with c>0c>0, we have

ΘY(γ(τ))=(i(cτ+d))r2cr2Mrdet(A)12h (mod M)hiρi (mod μi)k (mod M2N)Ak0 (mod N)Φ(MNh,M2A,k,M2N)θ(τ;M2A,k,M2N),\Theta_{Y}(\gamma(\tau))=\frac{(-i(c\tau+d))^{\frac{r}{2}}}{c^{\frac{r}{2}}M^{r}\det(A)^{\frac{1}{2}}}\sum_{\begin{subarray}{c}h\text{~{}(mod }M\text{)}\\ h_{i}\equiv\rho_{i}\text{~{}(mod }\mu_{i}\text{)}\end{subarray}}\sum_{\begin{subarray}{c}k\text{~{}(mod }M^{2}N\text{)}\\ Ak\equiv 0\text{~{}(mod }N\text{)}\end{subarray}}\Phi(MNh,M^{2}A,k,M^{2}N)\theta(\tau;M^{2}A,k,M^{2}N),

where Φ(h,S,k,m)\Phi(h,S,k,m) is defined as follows,

Φ(h,S,k,m)g (mod cm)gh (mod m)e2πi(ag𝗍Sg+2k𝗍Sg+dk𝗍Sk2cm2).\Phi(h,S,k,m)\coloneqq\sum_{\begin{subarray}{c}g\text{~{}(mod }cm\text{)}\\ g\equiv h\text{~{}(mod }m\text{)}\end{subarray}}e^{2\pi i\left(a\frac{g^{\mathsf{t}}Sg+2k^{\mathsf{t}}Sg+dk^{\mathsf{t}}Sk}{2cm^{2}}\right)}.

By [12, Page 15], we have |Φ(MNh,M2A,k,M2N)|cr2i=1rgcd(M2bj,c)12|\Phi(MNh,M^{2}A,k,M^{2}N)|\leq c^{\frac{r}{2}}\prod_{i=1}^{r}\gcd(M^{2}b_{j},c)^{\frac{1}{2}}, where b1,,brb_{1},\ldots,b_{r}\in\mathbb{Z} are the diagonal entries of the Smith normal form of the Hessian matrix AA. Hence, plugging it back to the transformation formula of ΘY\Theta_{Y}, we have

|aY,γ(n)|\displaystyle|a_{Y,\gamma}(n)| Mrμ1μrk (mod M2N)Ak0 (mod N)|{xrx𝗍Ax=2Nn,xk (mod M2N)}|\displaystyle\leq\frac{M^{r}}{\mu_{1}\cdots\mu_{r}}\sum_{\begin{subarray}{c}k\text{~{}(mod }M^{2}N\text{)}\\ Ak\equiv 0\text{~{}(mod }N\text{)}\end{subarray}}\left|\{x\in\mathbb{Z}^{r}\mid x^{\mathsf{t}}Ax=2Nn,x\equiv k\text{~{}(mod }M^{2}N\text{)}\}\right|
Mrμ1μr|{xrx𝗍Ax2Nn}|(M2N)r2Mnr2,\displaystyle\leq\frac{M^{r}}{\mu_{1}\cdots\mu_{r}}\left|\{x\in\mathbb{Z}^{r}\mid x^{\mathsf{t}}Ax\leq 2Nn\}\right|\leq\frac{(M^{2}N)^{\frac{r}{2}}}{M}n^{\frac{r}{2}},

where aY,γ(n)a_{Y,\gamma}(n) is the nn-th Fourier coefficient of ΘY\Theta_{Y} at the cusp γ(i)\gamma(i\infty) for a matrix γSL2()\gamma\in\text{SL}_{2}(\mathbb{Z}). Since the explicit construction of the ψ\psi-component GG is a weighted average of the theta series ΘY\Theta_{Y}, the same bound holds for the Fourier coefficient aj(n)a_{j}(n) for every 1jt1\leq j\leq t. Finally, plugging it back to the Petersson norm of GG, we obtain

GΓ0(M2N)2rCM,Nj=1t(M2N)r+2M2n=1e23πnM2Nnr1r,ε(M2N)2r+2+εM2.\lVert G\rVert_{\Gamma_{0}(M^{2}N)}^{2}\ll_{r}C_{M,N}\sum_{j=1}^{t}\frac{(M^{2}N)^{r+2}}{M^{2}}\sum_{n=1}^{\infty}e^{-\frac{2\sqrt{3}\pi n}{M^{2}N}}n^{r-1}\ll_{r,\varepsilon}\frac{(M^{2}N)^{2r+2+\varepsilon}}{M^{2}}.

Taking the square root, we obtain the desired result. ∎

Now we can state the main results of this section.

Theorem 5.3.

Let GXG_{X} be the cuspidal part of the theta series associated to a positive-definite integral shifted lattice XX of rank r4r\geq 4, level NN, and conductor MM. For any real number ε>0\varepsilon>0 and any integer n1n\geq 1, we have the following upper bounds:

  1. (1)

    If rr is odd, then we have

    |aGX(n)|ε,rϕ(M)(M2N)r+1+εMnr414+ε.|a_{G_{X}}(n)|\ll_{\varepsilon,r}\frac{\phi(M)(M^{2}N)^{r+1+\varepsilon}}{M}n^{\frac{r}{4}-\frac{1}{4}+\varepsilon}.
  2. (2)

    If rr is even, then we have

    |aGX(n)|ε,rϕ(M)(M2N)r+32+εMnr412+ε.|a_{G_{X}}(n)|\ll_{\varepsilon,r}\frac{\phi(M)(M^{2}N)^{r+\frac{3}{2}+\varepsilon}}{M}n^{\frac{r}{4}-\frac{1}{2}+\varepsilon}.

Here the implied constants are effective and ϕ\phi denotes Euler’s totient function.

Proof.

It follows from Proposition 5.1, Proposition 5.2, and (5.1). ∎

Applying Theorem 5.3 to the shifted lattice constructed in Definition 4.7, we obtain the following upper bounds.

Proposition 5.4.

Suppose that F=i=1raiPmiF=\sum_{i=1}^{r}a_{i}P_{m_{i}} is a node of depth r4r\geq 4 in the escalator tree TT_{\infty}. Let X,μ,ρX,\mu,\rho be the shifted lattice and the integers constructed in Definition 4.7 corresponding to FF. Set Λlcm(m12,,mr2)\Lambda\coloneqq\text{lcm}(m_{1}-2,\ldots,m_{r}-2). For any real number ε>0\varepsilon>0 and any integer n0n\geq 0, the following inequalities hold:

  1. (1)

    If rr is odd, then we have

    |aGX(μn+ρ)|ε,r(iai)r+1+εΛ3r+3+ε(μn+ρ)r414+ε.|a_{G_{X}}(\mu n+\rho)|\ll_{\varepsilon,r}\Big{(}\prod_{i}a_{i}\Big{)}^{r+1+\varepsilon}\Lambda^{3r+3+\varepsilon}(\mu n+\rho)^{\frac{r}{4}-\frac{1}{4}+\varepsilon}.
  2. (2)

    If rr is even, we have

    |aGX(μn+ρ)|ε,r(iai)r+32+εΛ3r+92+ε(μn+ρ)r412+ε.|a_{G_{X}}(\mu n+\rho)|\ll_{\varepsilon,r}\Big{(}\prod_{i}a_{i}\Big{)}^{r+\frac{3}{2}+\varepsilon}\Lambda^{3r+\frac{9}{2}+\varepsilon}(\mu n+\rho)^{\frac{r}{4}-\frac{1}{2}+\varepsilon}.

In both cases, the implied constants are effective.

Proof.

Suppose that the level and the conductor of the shifted lattice XX are denoted by NN and MM, respectively. The conductor MM is bounded above by 2Λ2\Lambda and the level NN is bounded above by 4Λiai4\Lambda\prod_{i}a_{i}. Plugging them into Theorem 5.3, we obtain the desired bounds. ∎

6. Proof of the Main Results

For any integers 𝔐𝔪21\mathfrak{M}\geq\mathfrak{m}-2\geq 1, the existence of the constants Γ𝔐\Gamma_{\mathfrak{M}} and Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}} follows from the following theorem.

Theorem 6.1.

For any integer 𝔐1\mathfrak{M}\geq 1, the tree T𝔐T_{\mathfrak{M}} is finite.

Proof.

By induction on the depth, we see that the parameters aia_{i} and mim_{i} are bounded for any i1i\geq 1 and there are finitely many nodes of a fixed depth in T𝔐T_{\mathfrak{M}}. 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 55 is bounded below by a uniform constant only depending on 𝔐\mathfrak{M}. 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 55 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 Γ𝔐\Gamma_{\mathfrak{M}} follows from Theorem 6.1. When 𝔐=1\mathfrak{M}=1, the calculation in [3] reveals that Γ𝔐=8\Gamma_{\mathfrak{M}}=8. For the upper bound on Γ𝔐\Gamma_{\mathfrak{M}}, by Proposition 2.2, Proposition 2.4, and Proposition 2.6, we can bound the truants of the non-leaf nodes of depth r4r\leq 4 in TT by an implicit constant. So we arrive at depth 55. The parameters a1,,a5a_{1},\ldots,a_{5} 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 55 in T𝔐T_{\mathfrak{M}} represents every positive integer nn such that

nε𝔐43+ε.n\gg_{\varepsilon}\mathfrak{M}^{43+\varepsilon}.

This implies that the truant of any non-leaf node of depth 55 in T𝔐T_{\mathfrak{M}} is bounded above by the right hand side. Since any non-leaf node of depth r6r\geq 6 will represent more integers than its ancestor of depth 55, 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 r4r\leq 4. ∎

Proof of Theorem 1.6.

Since we have Γ𝔪,𝔐Γ𝔐\Gamma_{\mathfrak{m},\mathfrak{M}}\leq\Gamma_{\mathfrak{M}}, the existence of the constant Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}} is clear for any integers 𝔪,𝔐\mathfrak{m},\mathfrak{M} such that 𝔐𝔪21\mathfrak{M}\geq\mathfrak{m}-2\geq 1. If 𝔪20\mathfrak{m}\geq 20, any node of depth r4r\leq 4 with mi𝔪m_{i}\geq\mathfrak{m} for 1ir1\leq i\leq r can not represent 1616. Thus the constant Γ𝔪,𝔐\Gamma_{\mathfrak{m},\mathfrak{M}} is effective and it satisfies the same bound in Theorem 1.1 with an effective implied constant under the assumption. If 𝔪36\mathfrak{m}\geq 36, any node of depth r5r\leq 5 with mi𝔪m_{i}\geq\mathfrak{m} for 1ir1\leq i\leq r can not represent 3232. For nodes of depth r=6r=6, we can use a stronger upper bound for the cuspidal part when rr is even. Therefore we have

Γ𝔪,𝔐ε𝔐27+ε,\Gamma_{\mathfrak{m},\mathfrak{M}}\ll_{\varepsilon}\mathfrak{M}^{27+\varepsilon},

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 125125 cases that are proved to be universal. For the reference, see Table LABEL:tbl::confirmedcases. For the remaining 7272 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 mm-gonal numbers. International Mathematics Research Notices, 2020(20):6999–7036, 2020.
  • [15] Byeong Moon Kim and Dayoon Park. A finiteness theorem for universal mm-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 θ\theta sur le k-ième nombre premier et grandes valeurs de la fonction ω\omega(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.