1. Introduction and statement of results
1·1. Main results
Let $\zeta(s)$ be the Riemann zeta function, and let $N(\sigma,T)$ denote the number of zeros $\beta+i\gamma$ of $\zeta(s)$ with $\beta\geq\sigma\geq 0$ and $|\gamma|\leq T$ . The asymptotic
follows from the argument principle. The Riemann hypothesis (RH) asserts that $\zeta(s)\neq 0$ for $\textrm{Re}(s)>{1}/{2}$ , so $N(\sigma,T)=0$ for $\sigma>{1}/{2}$ . Selberg [ Reference TitchmarshTit86 , theorem 9·19C] proved a delicate zero density estimate that recovers the upper bound in (1·1) at $\sigma={1}/{2}$ , namely
(See Baluyot [ Reference Alan and BaluyotBal17 , theorem 1·2·1] for an improvement.) Selberg’s estimate implies that
As an application of (1·2), Selberg proved a central limit theorem for $\log|\zeta({1}/{2}+it)|$ :
Let $\mathscr{A}$ be the set of cuspidal automorphic representations of $\textrm{GL}_2$ over $\mathbb{Q}$ with unitary central character. For $\pi\in\mathscr{A}$ , let $L(s,\pi)$ be its standard L-function. Define
As with $\zeta(s)$ , the argument principle can be used to prove that
The generalised Riemann hypothesis (GRH) asserts that $L(s, \pi) \neq 0$ for $\textrm{Re}(s) > {1}/{2}$ . Jutila [ Reference JutilaJut83 ] and Selberg [ Reference SelbergSel92 ] observed that analogues of (1·2) and (1·3) should also hold for Hecke–Maaß newforms. When $\pi \in \mathscr{A}$ corresponds with a holomorphic cuspidal newform of even weight $k \geq 2$ , Luo [ Reference LuoLuo95 , theorem 1·1] and Li [ Reference Ford and ZaharescuFZ15 , section 7] proved that $N_{\pi}(\sigma, T) \ll T^{1-\frac{1}{72}(\sigma-\frac{1}{2})}\log T$ . We prove:
Theorem 1·1. Let $\theta\in[0,{7}/{64}]$ be an admissible exponent toward the generalised Ramanujan conjecture for Hecke–Maaß newforms (see (2·1)), and fix $0<c<{1}/{4}-{\theta}/{2}$ . If $\pi \in \mathscr{A}$ , $\sigma \geq {1}/{2}$ , and $T \geq 2$ , then $N_{\pi}(\sigma, T) \ll T^{1-c(\sigma-\frac{1}{2})}\log T$ . The implied constant depends at most on $\pi$ .
Remark 1. Under the generalised Ramanujan conjecture for all Hecke–Maaß newforms, we may take $\theta=0$ . In this case, our result is as strong as Selberg’s zero density estimate (1·2) for $\zeta(s)$ . Currently, the best unconditional bound is $\theta\leq {7}/{64}$ , so we may choose any $c<{25}/{128}$ . This noticeably improves the work of Luo and Li, and it holds for any $\pi\in\mathscr{A}$ . The constant ${1}/{4}-{\theta}/{2}$ is, as of now, the supremum over all $\varpi$ for which we can unconditionally prove an asymptotic for the second mollified moment of $L(s,\pi)$ on $\textrm{Re}(s)={1}/{2}+{1}/{\log T}$ with a mollifier of length $T^{\varpi}$ .
Corollary 1·2. Let $\pi \in \mathscr{A}$ . If $V\in\mathbb{R}$ , then as $T\to\infty$ , we have:
Proof. Bombieri and Hejhal [ Reference Bombieri and HejhalBH95 , theorem B] proved that this follows from a zero density estimate of the quality given by Theorem 1·1.
Remark 2. Radziwiłł and Soundararajan [ Reference Radziwiłł and SoundararajanRS17 ] recently found a second proof of (1·3) which avoids the use of zero density estimates. Their work was recently extended to holomorphic newforms by Das [ Reference DasDas20 ]. The proof relies on both the generalised Ramanujan conjecture and the Sato–Tate conjecture, neither of which is known for any Hecke–Maaß newform.
We study the distribution of imaginary parts of the nontrivial zeros of $L(s, \pi)$ using Theorem 1·1. First, we give an incomplete history of such results for $\zeta(s)$ . Hlawka [ Reference HlawkaHla75 ] proved that if $\alpha\in\mathbb{R}$ is fixed and $h\colon \mathbb{T}\to\mathbb{C}$ is continuous, then
where $\gamma$ varies over the imaginary parts of the nontrivial zeros of $\zeta(s)$ and $\mathbb{T}=\mathbb{R}/\mathbb{Z}$ . Thus the sequence of fractional parts $\{\gamma\alpha\}$ is equidistributed modulo 1. However, given a rate of convergence, there exist continuous functions h such that the limit in (1·5) cannot be attained with said rate (see also [ Reference Ford and ZaharescuFZ05 , theorem 7]). Therefore, (1·5) is the best that one can say for arbitrary h.
Ford and Zaharescu [ Reference Ford and ZaharescuFZ05 , corollary 2] established the existence of a second order term, proving that if $h\;:\;\mathbb{T}\to\mathbb{C}$ is twice continuously differentiable,Footnote 1 then
where
Despite the limitations on the analytic properties of h, Ford and Zaharescu still conjectured [ Reference Ford and ZaharescuFZ05 , conjecture A] that for any interval $\mathbb{I}\subseteq\mathbb{T}$ of length $|\mathbb{I}|$ , we have
which implies that
Ford, Soundararajan and Zaharescu [ Reference Ford, Soundararajan and ZaharescuFSZ09 ] made some progress toward the conjectured asymptotics (1·8) and (1·9). Unconditionally, they proved that
Assuming RH, they proved that
Along with making some appealing connections between the conjectured asymptotics (1·8) and (1·9) and other intriguing open problems like pair correlation of zeros of $\zeta(s)$ and the distribution of primes in short intervals, they proved analogues for other L-functions of (1·10) (assuming a zero density estimate of the form (1·2)) and (1·11) (assuming GRH).
In this paper, we extend the work in [ Reference Ford and ZaharescuFZ05, Reference Ford, Soundararajan and ZaharescuFSZ09, Reference Ford, Meng and ZaharescuFMZ17, Reference Liu and ZaharescuLZ21 ] to $L(s,\pi)$ for any $\pi\in\mathscr{A}$ using Theorem 1·1. Let $n \ge 1$ . Consider the $\boldsymbol{\alpha}\in\mathbb{R}^n$ for which there exists a constant $C_{\boldsymbol{\alpha}}>0$ such thatFootnote 2
where $\|\boldsymbol{m}\|_p$ is the $\ell^p$ norm on $\mathbb{R}^n$ for $1\leq p\leq\infty$ . This is a technical artifact of our extension to $\mathbb{R}^n$ ; when $n=1$ , the condition reduces to $\alpha\neq 0$ . Our density function $g_{\pi, \boldsymbol{\alpha}}(\boldsymbol{t})$ , which extends (1·7) for $n\geq 2$ , is identically zero unless there exists a matrix $M=(b_{jk}) \in \mathcal{M}_{r\times n}(\mathbb{Z})$ with linearly independent row vectors $\boldsymbol{b}_j$ and $\gcd(b_{j1}, \ldots, b_{jn}) = 1$ for all $1 \leq j \leq r$ ; fully reduced rationals $a_1/q_1, \ldots, a_r/q_r$ ; and distinct primes $p_1, \ldots, p_r$ such that
Among such possible matrices M, choose one with maximal r, which uniquely determines the row vectors $\boldsymbol{b}_j=(b_{j1},\ldots,b_{jn})$ . If such an M exists, then define
where
Let $\pi\in\mathscr{A}$ , $n\geq 1$ , and $\mathbb{B} \subseteq \mathbb{T}^n$ be a product of n subintervals of $\mathbb{T}$ . The conjecture in (1·8) can be extended to $\pi\in\mathscr{A}$ as follows:
where $\gamma$ ranges over the imaginary parts of the nontrivial zeros of $L(s,\pi)$ . As progress toward (1·15), we prove an unconditional n-dimensional version of (1·6) for $\textrm{GL}_2$ L-functions. In what follows, let $C^u(\mathbb{T}^n)$ be the set of u-times continuously differentiable functions $h\colon\mathbb{T}^n\to\mathbb{R}$ . Let $\gamma$ vary over the imaginary parts of the nontrivial zeros of $L(s, \pi)$ .
Theorem 1·3. Let $\pi \in \mathscr{A}$ . Let $\boldsymbol{\alpha} \in \mathbb{R}^n$ satisfy (1·12). If $h \in C^{n+2}(\mathbb{T}^n)$ , then
where ${d}\boldsymbol{t}$ is Lebesgue measure on $\mathbb{T}^n$ . The implied constant depends on $\pi$ , h, and $\boldsymbol{\alpha}$ .
Remark 3. With extra work, we may allow $h\in C^{n+1}(\mathbb{T}^n)$ . Also, under some extra conditions on $\boldsymbol{\alpha}$ which excludes a density zero subset of $\mathbb{R}^n$ , we expect to further quantify the o(T) error term in (1·16).
Define the discrepancy
Our next result follows quickly from Theorem 1·3.
Corollary 1·4. Let $\pi\in\mathscr{A}$ . If $\boldsymbol{\alpha}\in\mathbb{R}^n$ satisfies (1·12), then
When $n=1$ , we recover an unconditional analogue of (1·10) for all $\pi\in\mathscr{A}$ .This special case of Corollary 1·4 was proved in [ Reference Ford, Soundararajan and ZaharescuFSZ09 ] under the hypothesis of a zero density estimate of the form proved in Theorem 1·1.
Remark 4. Let $d\geq 3$ be an integer, and let $L(s,\pi)$ be the standard L-function associated to a cuspidal automorphic representation $\pi$ of $\textrm{GL}_d$ over $\mathbb{Q}$ . Let $N_{\pi}(\sigma,T)$ be the number of nontrivial zeros $\beta+i\gamma$ of $L(s,\pi)$ with $\beta\geq\sigma$ and $|\gamma|\leq T$ . If there exist constants $c_{\pi}>0$ and $d_{\pi}>0$ such that $N_{\pi}(\sigma,T)\ll T^{1-c_{\pi}(\sigma-\frac{1}{2})}(\!\log T)^{d_{\pi}}$ , then one can prove an analogue of Theorem 1·3 and Corollary 1·4 for $L(s,\pi)$ . Such an estimate for $N_{\pi}(\sigma,T)$ is not yet known for any $d\geq 3$ , and it appears to be quite difficult to prove.
1·2. Application to “zero races”
In a letter to Fuss, Chebyshev observed that the generalised Riemann hypothesis implies that primes $p\equiv 3\pmod{4}$ tend to be more numerous than primes $p\equiv 1\pmod{4}$ . Using the generalised Riemann hypothesis and other hypotheses, Rubinstein and Sarnak [ Reference Rubinstein and SarnakRS94 ] began a systematic study of “prime number races” in which they determine how often $\pi(x;4,3)>\pi(x;4,1)$ , where $\pi(x;q,a)$ equals $\#\{p\leq x\colon p\equiv a\pmod{q}\}$ . They proved that
Thus the “bias” toward primes of the form $4n+3$ is quite strong. The literature on prime number races which study such inequities is quite vast. See, for instance, the work of Fiorilli, Ford, Harper, Konyagin, Lamzouri and Martin [ Reference Fiorilli and MartinFM13, Reference Ford and KonyaginFK02, Reference Ford, Lamzouri and KonyaginFLK13, Reference Ford, Harper and LamzouriFHL19 ].
We use Theorem 1·3 to study inequities not between primes in different residue classes, but between zeros of different L-functions. Define
Let $h\in\mathscr{C}^{n+2}(\mathbb{T}^n)$ , and let $\boldsymbol{\alpha}$ satisfy (1·12). We consider two holomorphic cusp forms $f_1$ and $f_2$ with trivial nebentypus, where
have trivial nebentypus, even integral weights $k_j\geq 2$ and squarefree levels $q_j\geq 1$ . We assume that $f_j$ is normalised so that $\lambda_{f_j}(1)=1$ and that $f_j$ is an eigenfunction of all of the Hecke operators. We call such cusp forms newforms (see [ Reference OnoOno04 , section 2·5]). It is classical that there exists $\pi_j\in\mathscr{A}$ such that $L(s,f_j)=L(s,\pi_j)$ , so $\lambda_{f_j}(n) = \lambda_{\pi_j}(n)$ , $g_{f_j,\boldsymbol{\alpha}}(\boldsymbol{t})=g_{\pi_j,\boldsymbol{\alpha}}(\boldsymbol{t})$ , etc.
We say $f_1$ wins the $(\boldsymbol{\alpha}, h)$ -race (against $f_2$ ) if for all large T, we have
where $\gamma_j$ runs over the imaginary parts of the nontrivial zeros of $L(s, f_j)$ . If neither $f_1$ nor $f_2$ wins the $(\boldsymbol{\alpha}, h)$ -race, we say the race is undecided. The results of [ Reference Liu and ZaharescuLZ21 ] suggest that the winner of a decided $(\boldsymbol{\alpha},h)$ -race is determined by the levels $q_1$ and $q_2$ and the behavior of $g_{f_1, \boldsymbol{\alpha}}$ and $g_{f_2, \boldsymbol{\alpha}}$ .
First, we show that if $q_1 = q_2$ and $h \in \mathscr{C}^{3}(\mathbb{T})$ , then the proportion of primes p such that $f_1$ wins the $({\log p}/{2\pi}, h)$ -race is ${1}/{2}$ .
Corollary 1·5. For $j=1,2$ , let $q_j$ be squarefree and $f_j \in S_{k_j}^{\textrm{new}} (\Gamma_0(q_j))$ be non-CM newforms. Suppose that $f_1\neq f_2\otimes\chi$ for all primitive Dirichlet characters $\chi$ . If $h \in \mathscr{C}^{3}(\mathbb{T})$ , then
Next we look at the distribution of
as $\alpha$ varies over values of ${\log p}/{2\pi}$ for prime values of p. Given $I \subseteq [\!-\!2, 2]$ , let $\mu_{\textrm{ST}}(I)$ be the Sato–Tate measure $1/(2\pi)\int_{I}\sqrt{4-t^2}dt$ , and let $\mu_{ST,2}$ be the product measure defined on boxes $I_1\times I_2\subseteq [\!-\!2,2]\times[\!-\!2,2]$ by $\mu_{\textrm{ST},2}(I_1\times I_2)=\mu_{\textrm{ST}}(I_1)\mu_{\textrm{ST}}(I_2)$ . For $\mathcal{I} \subseteq [\!-\!4,4]$ , we define
For $h \in \mathscr{C}^3(\mathbb{T})$ , we set
In the following statement and throughout the paper, for any interval I and $\beta \in \mathbb{R}$ , we let $\beta I = \{ \beta x\;:\; x \in I \}$ .
Corollary 1·6. For $j=1,2$ , let $q_j$ be squarefree and $f_j \in S_{k_j}^{\textrm{new}} (\Gamma_0(q_j))$ be normalised holomorphic non-CM newforms with trivial nebentypus. Suppose that $f_1\neq f_2\otimes\chi$ for all primitive Dirichlet characters $\chi$ . Let
Assume $h \in \mathscr{C}^3(\mathbb{T})$ is such that $k_h \neq 0$ . Then for any interval $\mathcal{I}\subseteq[\!-\!4,4]$ , we have
with an implied constant independent of $\mathcal{I}$ .
If $q_1 > q_2$ , then in contrast to Corollary 1·5, $f_1$ wins the $({\log p}/{2\pi},h)$ -race for all except finitely many primes p.
Corollary 1·7. For $j=1,2$ , let $f_j \in S_{k_j}^{\textrm{new}} (\Gamma_0(q_j))$ be normalised holomorphic newforms with trivial nebentypus. If $q_1>q_2$ , then for any $h \in \mathscr{C}^3(\mathbb{T})$ , there are at most finitely many primes p such that $\pi_2$ wins the $({\log p}/{2\pi},h)$ -race.
Corollary 1·7 shows that it is rare for $\pi_1$ to win an $(\alpha,h)$ race against $\pi_2$ if $q_2>q_1$ , but we can show this occurs infinitely often. Our result can be stated neatly in terms of local races rather than in terms of the $(\alpha, h)$ -races described above. For $t_0 \in \mathbb{T}$ , we say that $f_1$ wins the local $(\alpha, t_0)$ -race against $f_2$ if there exists a neighbourhood U of $t_0 \in \mathbb{T}$ such that the $(\alpha, h)$ -race is won by $f_1$ for all $h \in \mathscr{C}^{3}(\mathbb{T})$ which are supported on U.
Corollary 1·8. For $j=1,2$ , let $q_j$ be squarefree and $f_j \in S_{k_j}^{\textrm{new}} (\Gamma_0(q_j))$ be normalised holomorphic non-CM newforms with trivial nebentypus. Suppose that $f_1\neq f_2\otimes\chi$ for all primitive Dirichlet characters $\chi$ . Fix $t_0 \in [0,1)$ . Let $q_1$ be sufficiently large, and let $q_2 \in (q_1, q_1 + q_1^{1/2}]$ , and let $k_1$ and $k_2$ be fixed. There exists $\alpha \in \mathbb{R}$ such that $f_1$ wins the local $(\alpha, t_0)$ -race against $f_2$ .
In addition to Theorem 1·3, the proofs of these corollaries rely on a quantifiable understanding of the joint distribution of $\lambda_{f_1}(p)$ and $\lambda_{f_2}(p)$ as p varies over the primes. Such an understanding follows from the effective version of the Sato–Tate conjecture which counts the number of primes $p\leq X$ such that $(\lambda_{f_1}(p),\lambda_{f_2}(p))\in I_1\times I_2$ proved by the third author in [ Reference ThornerTho21 ] (see Theorem 2·1 below). One can prove analogues of Corollaries 1·5, 1·6 and 1·8 for Dirichlet L-functions by replacing the effective Sato–Tate estimates with results on primes in arithmetic progressions.
2. Preliminaries
2·1. $\operatorname{GL}_2$ L-functions over $\mathbb{Q}$
Let $\pi\in\mathscr{A}$ . Here, we state the essential properties of L-functions of $L(s,\pi)$ that we use throughout our proofs. See [ Reference Iwaniec and KowalskiIK04 , chapter 5] for a convenient summary. Given $\pi\in\mathscr{A}$ with level $q_{\pi}$ , there exist suitable complex numbers $\alpha_{1,\pi}(p)$ and $\alpha_{2,\pi}(p)$ such that
The sum and product both converge absolutely for $\textrm{Re}(s)>1$ . There also exist spectral parameters $\kappa_{\pi}(1)$ and $\kappa_{\pi}(2)$ such that if we define
then the completed L-function $\Lambda(s,\pi)\;:\!=\;q_{\pi}^{s/2}L(s,\pi)L(s,\pi_{\infty})$ is entire of order 1.
Let $\tilde{\pi}\in\mathscr{A}$ be the contragredient representation. We have $\alpha_{j,\tilde{\pi}}(p)=\overline{\alpha_{j,\pi}(p)}$ and $\kappa_{\tilde{\pi}}(j)=\overline{\kappa_{\pi}(j)}$ for $j=1,2$ . Moreover, there exists a complex number $W(\pi)$ of modulus 1 such that for all $s\in\mathbb{C}$ , we have
Building on work of Kim and Sarnak [ Reference KimKim03 , appendix], Blomer and Brumley [ Reference Blomer and BrumleyBB11 ] proved that there exists $\theta\in[0,7/64]$ such that we have the uniform bounds
The generalised Ramanujan conjecture and the Selberg eigenvalue conjecture assert that (2·1) holds with $\theta=0$ .
The Rankin–Selberg L-functionFootnote 3
factors as $\zeta(s)L(s,\textrm{Ad}^2\pi)$ , where the adjoint square lift $\textrm{Ad}^2\pi$ is an automorphic representation of $\textrm{GL}_3(\mathbb{A}_{\mathbb{Q}})$ . Thus, $L(s,\textrm{Ad}^2\pi)$ is an entire automorphic L-function. This fact and the bound $|\lambda_{\pi}(n)|^2\leq \lambda_{\pi\times\tilde{\pi}}(n)$ [ Reference Jiang, Lü and WangJLW21 , lemma 3·1], enable us to prove via contour integration that
It follows from [ Reference Iwaniec and KowalskiIK04 , theorem 5·42] applied to $\zeta(s)$ and $L(s,\textrm{Ad}^2\pi)$ that there exists an effectively computable constant $c_{\pi}>0$ such that $L(s,\pi\times\tilde{\pi})\neq 0$ in the region
2·2. Holomorphic newforms
Many of our corollaries pertain specifically to $\pi\in\mathscr{A}$ corresponding to holomorphic newforms. As above, let
be a holomorphic cuspidal newform (normalised so that $\lambda_f(1)=1$ ) of even integral weight $k\geq 2$ , level $q\geq 1$ , and trivial nebentypus. If $\pi_f\in\mathscr{A}$ corresponds with f, then $L(s,f)=L(s,\pi_f)$ , $\lambda_f(n)=\lambda_{\pi_f}(n)$ , etc.
For these newforms, it follows from Deligne’s proof of the Weil conjectures that the generalised Ramanujan conjecture holds. Thus, for f a holomorphic cuspidal newform as above, we may take $\theta=0$ in (2·1). Since we assume that f has trivial central character, Deligne’s bound implies that there exists $\theta_p\in[0,\pi]$ such that
The Sato–Tate conjecture, now a theorem due to Barnet-Lamb, Geraghty, Harris, and Taylor [ Reference Barnet-Lamb, Geraghty, Harris and TaylorBLGHT11 ], states that the sequence $(\theta_p)$ is equidistributed in the interval $[\!-\!2,2]$ with respect to the measure ${2}/{\pi}(\sin t)^2 dt$ . In other words, if $I\subseteq[0,\pi]$ is a subinterval, then
After a change of variables, this implies that for any interval $I \subseteq [\!-\!2,2]$ , we have
A recent paper by Thorner [ Reference ThornerTho21 ] provides both an unconditional and a GRH-conditional rate of convergence in (2·4).
Our corollaries of Theorem 1·3 require a natural refinement of the Sato–Tate conjecture. For $j=1,2$ , let $f_j\in S_{k_j}^{\textrm{new}}(\Gamma_0(q_j))$ be a holomorphic cuspidal newform as above. Suppose that $f_1\neq f_2\otimes\chi$ for all primitive nontrivial Dirichlet characters $\chi$ . Building on work of Harris [ Reference HarrisHar09 ], Wong [ Reference WongWon19 ] proved that the sequences $(\lambda_{f_1}(p))$ and $(\lambda_{f_2}(p))$ exhibit a joint distribution: If $I_1,I_2\subseteq [\!-\!2,2]$ , then
Our corollaries of Theorem 1·3 require a nontrivial unconditional bound on the rate of convergence in (2·5). Such a bound was recently proved in [ Reference ThornerTho21 , theorem 1·2].
Theorem 2·1. For $j=1,2$ , let $q_j$ be squarefree and $f_j \in S_{k_j}^{\textrm{new}}(\Gamma_0(q_j))$ be a normalised holomorphic cuspidal newform with trivial nebentypus. Suppose that $f_1\neq f_2\otimes\chi$ for all primitive nontrivial Dirichlet character $\chi$ . Let $I_1, I_2\subseteq [\!-\!2, 2]$ be subintervals. There exists an absolute and effectively computable constant $c>0$ such that
We use the following result to prove our corollaries of Theorem 1·3.
Corollary 2·2. Let $f_1$ and $f_2$ be as in Theorem 2·1. Recall the definition of $\nu$ in (1·17). If $\mathcal{I} \subseteq[\!-\!4,4]$ , then
with an implied constant independent of $\mathcal{I}$ .
Proof. Let
We will define rectangles whose unions approximate S. Let $x_j = -2 + j \lfloor {4}/{g(X)} \rfloor$ . If $\mathcal{I} = ( c, d)$ , set $R_j = [x_j, x_{j+1}] \times [x_j - d, x_{j+1} - c]$ , and similarly $T_j \;:\!=\; [x_j, x_{j+1}] \times [x_{j+1} - d, x_j - c]$ . By construction, we have $\cup_j T_j \subseteq S \subseteq \cup_j R_j$ , which implies that
We apply Theorem 2·1 to count the primes p with $(\lambda_{f_1} (p), \lambda_{f_2}(p))$ in $\cup_j R_j$ .
Since the area of $\cup_j R_j \setminus S$ is at most $g(X) \lfloor {4}/{g(X)} \rfloor^2$ , we have
and we conclude that
The same argument shows that (2·7) holds with $R_j$ replaced with $T_j$ on the left-hand side. The lemma now follows from (2·6).
Finally, we require a refinement of (1·4), namely
This is [ Reference Iwaniec and KowalskiIK04 , theorem 5·8] applied to L(s,f).
3. Proof of Theorem 1·1
Let $\pi\in\mathscr{A}$ . We detect the zeros of $L(s,\pi)$ by estimating a mollified second moment of $L(s,\pi)$ near the line $\textrm{Re}(s)={1}/{2}$ . To describe our mollifier, we define $\mu_{\pi}(n)$ by the convolution identity
so that
Let $T>0$ be a large parameter, and let $0<\varpi<{1}/{4}$ . Define P(t) by
Our mollifier is
As a proxy for detecting zeros of $L(s,\pi)$ near $\textrm{Re}(s)={1}/{2}$ , where $L(s,\pi)$ oscillates wildly, we detect zeros of the mollified L-function $L(s,\pi)M_{\pi}(s, T^{\varpi})$ near $\textrm{Re}(s)={1}/{2}$ . To this end, we let $w\;:\;\mathbb{R}\to\mathbb{R}$ be an infinitely differentiable function whose support is a compact subset of $[T/4,2T]$ and whose j-th derivative satisfies $|w^{(j)}(t)|\ll_{w,j}((\!\log T)/T)^j$ for all $j\geq 0$ . Also, let $w(t)=1$ for $t\in[T/2,T]$ . We will estimate
eventually choosing $\alpha$ and $\beta$ to equal $1/\log T$ . Define
As in [ Reference BernardBer15, Reference Andersen and ThornerAT21 ], it follows from the approximate functional equation (see [ Reference Iwaniec and KowalskiIK04 , section 5.2] also) that
where we have split the sum into diagonal terms
and off-diagonal terms
Lemma 3·1. Let $\varepsilon>0$ . If $\alpha,\beta\in\mathbb{C}$ satisfy $|\alpha|,|\beta|\ll {1}/{\log T}$ and $|\alpha+\beta|\gg {1}/{\log T}$ , then for any integers $a,b\geq 1$ , we have that $N_{a,b}^{\pm}(\alpha,\beta)\ll_{\varepsilon}(ab)^{\frac{1}{2}}T^{\frac{1}{2}+\theta}(abT)^{\varepsilon}$ .
Proof. This is [ Reference Andersen and ThornerAT21 , proposition 3·4].
Corollary 3·2. Fix $0< \varpi< {1}/{4}-{\theta}/{2}$ . There exists a constant $\delta>0$ such that
Proof. Let $\varepsilon>0$ . First, observe that $|\mu_{\pi}(a)|\ll 1+|\lambda_{\pi}(a)|\ll 1+|\lambda_{\pi}(a)|^2$ . We then apply Lemma 3·1 and bound everything else trivially to obtain
By (2·2), this is $\ll_{f,\varepsilon} T^{\frac{1}{2}+\theta + 2\varpi +\varepsilon}$ . If $\varpi \leq {1}/{4} - {\theta}{2}-\varepsilon$ , then the above display is $\ll T^{1-\varepsilon}$ .
Proposition 3·3. If T is sufficiently large and $\varpi\in(0, {1}/{4}-{\theta}/{2})$ is fixed, then
Proof. In light of Corollary 3·2, it remains to bound the diagonal contribution. We first note by a calculation identitcal to [ Reference BernardBer15 , lemma 11] that
So it suffices for us to estimate
with $\alpha=\beta={1}/{\log T}$ .
Let $\sigma_0 = {1}/{2}+{1}/{\log T}$ . We observe via the Mellin inversion that $I^D(\alpha,\beta)$ equals
By a computation identical to [ Reference BernardBer15 , lemma 6], there exists a product of half-planes containing an open neighborhood of the point $u=v=s=0$ and an Euler product $A_{\alpha,\beta}(u,v,s)$ , absolutely convergent for (u,v,s) in said product of half-planes, such that
By Möbius inversion and a continuity argument, one can prove that $A_{0,0}(0,0,0)=1$ [ Reference BernardBer15 , lemma 7].
Upon choosing $\delta$ sufficiently small and shifting the contours to $\textrm{Re}(u)=\delta$ , $\textrm{Re}(v)=\delta$ , and $\textrm{Re}(s)=-\delta/2$ , we find that
The contribution from the pole at $u=v=0$ determines the magnitude of the double integral over u and v, and this magnitude is $\asymp_{\pi}(\!\log T)^2$ for $\alpha$ and $\beta$ in our prescribed range. (This follows once we push the u- and v-contours to the left using (2·3).) Since $\int_{\mathbb{R}}w(t)dt\asymp T$ by hypothesis, we combine all preceding contributions and choose $\alpha=\beta=1/\log T$ to conclude the desired bound.
Corollary 3·4. If T is sufficiently large and $\varpi\in(0, {1}/{4}-{\theta}/{2})$ is fixed, then
Proof. This follows from Proposition 3·3 and the Cauchy–Schwarz inequality.
We prove a corresponding estimate on a vertical line to the right of $\textrm{Re}(s)=1$ .
Lemma 3·5. If $\varpi\in(0, {1}/{4}-{\theta}/{2})$ and $A>\theta$ are fixed and T is sufficiently large, then
Proof. Let $A>\theta$ . It follows immediately from [ Reference Montgomery and VaughanMV74 , corollary 3] that if $T>1$ and $(b_n)$ is any sequence of complex numbers satisfying $\sum_{n}n|b_n|^2<\infty$ , then
We define $a_n$ by the identity
By (3·1), (3·2), and (3·3), we have that $a_n=0$ for all $n\leq T^{\varpi/2}$ and $|a_n|^2\ll (n^{\theta}d_4(n))^2\ll n^{2\theta}d_{16}(n)$ , where $d_k(n)$ is the n-th Dirichlet coefficient of $\zeta(s)^k$ . The desired result follows once we apply (3·4) with $b_n = a_n n^{-1-A}$ .
Proof of Theorem 1·1 . We use Gabriel’s convexity principle [ Reference TitchmarshTit86 , section 7·8] to interpolate the bounds in Corollary 3·4 and Lemma 3·5. In particular, if $0<\varpi< {1}/{4}-{\theta}/{2}$ and $A>\theta$ are fixed and $c=\varpi(1+2A-2\theta)/(1+2A)$ , then in the range ${1}/{2}+{1}/{\log T}\leq\sigma\leq 1+A$ , we have
Define $\Phi(s) \;:\!=\; 1- (1- L(s,\pi) M_{\pi}(s,T^{\varpi}))^2$ . By construction, if $\alpha\in\mathbb{C}$ , then
For any $M \ge 1$ , let $C_M$ be the rectangular contour with corners $\sigma + iT/2$ , $\sigma + iT$ , $M + iT/2$ , and $M + i T$ . Applying Littlewood’s lemma ([ Reference TitchmarshTit58 , pp. 132–133]) and letting $M \to \infty$ , we have
In view of the bound $\log|1 + z| \le |z|$ , it follows from (3·5) that
For the second and third integrals, we consider the integrals over $x \le 1$ and over $x >1$ separately. For $x \in (1, \infty)$ we can trivially bound the Dirichlet series as
where $(a_n)$ is a certain sequence of complex numbers such that $|a_n|\ll_{\varepsilon}n^{\theta+\varepsilon}$ for any fixed $\varepsilon>0$ and all $n\geq 2$ . Thus, we have
so
To handle the integrals for ${1}/{2}+{1}/{\log T}\leq x\leq 1$ , we use the trivial bound
A proof of the final bound is contained within the proof of [ Reference Iwaniec and KowalskiIK04 , theorem 5·8]. The corresponding integral for $\Phi(x+iT)$ has the same bound.
By the preceding work, it follows by dyadic decomposition that
This estimate, the mean value theorem for integrals, and the fact that $N_{\pi}(\sigma,T)$ is monotonically decreasing as $\sigma$ increases together imply that
If ${1}/{2}\leq \sigma\leq {1}/2 +{1}/{\log T}/2$ , then (1·4) implies that $N_{\pi}(\sigma,T)\ll T\log T\asymp T^{1-c(\sigma-{1}/{2})}\log T$ .
To finish the proof, note that if $\theta=0$ , then for all $A>0$ , we have $c=\varpi$ . If $\theta>0$ , then fix $0<\varepsilon<\theta\varpi/(\theta+{1}/{2})$ and choose $A = {\theta\varpi}/{\varepsilon}-{1}/{2}$ . With these choices, we find that $c>\varpi-\varepsilon$ . Theorem 1·1 now follows.
4. Proof of Theorem 1·3
We begin with a few preliminary lemmas. Throughout the section, $\theta$ is an admissible exponent toward the generalised Ramanujan conjecture as in Theorem 1·1. Our first result is an n-dimensional version of the Riemann–Lebesgue lemma.
Lemma 4·1. Let $J \geq 1$ . Suppose that $h\in C^{n+2}(\mathbb{T}^n)$ has the Fourier expansion
We have $|c_{\boldsymbol{m}}|\ll_h \|\boldsymbol{m}\|_2^{-n-2}$ , and consequently, we have
Proof. We have
Let $\boldsymbol{m} = (m_1, m_2, \ldots, m_n)$ . Choose $j \in \{1, \ldots, n\}$ such that $|m_j|=\|\boldsymbol{m}\|_{\infty}$ . Integrate (4·1) by parts for $n+2$ times with respect to the coordinate $t_j$ of $\boldsymbol{t}$ so that
Since $\sqrt{n}\|\boldsymbol{m}\|_{\infty}\geq \|\boldsymbol{m}\|_2$ , the desired result follows from the triangle inequality.
Lemma 4·2. Let $x > 1$ and $T \ge 2$ , and let $\langle x\rangle$ being the closest integer to x. We have
Proof. This is [ Reference Ford, Soundararajan and ZaharescuFSZ09 , lemma 2] with $\varepsilon=\theta$ .
Using Theorem 1·1 and Lemma 4·2, we prove an analogue of [ Reference Ford and ZaharescuFZ05 , (3·8)].
Lemma 4·3. Let c be as in Theorem 1·1. If $1< x< \exp(({c}/{3})({\log T}/{\log\log T}))$ , then
Proof. Let $\delta = (({3}/{c})({\log\log T}/{\log T}))$ , so $0 < \delta \log x < 1$ . By Theorem 1·1, we have that
By the functional equation for $L(s,\pi)$ , $\beta+i\gamma$ is a nontrivial zero if and only if $1-\beta+i\gamma$ is a nontrivial zero. Therefore, we have
Note that $x^{\beta-1/2}+x^{-(\beta-1/2)}-2=(2\sinh({1}/{2}(\beta-{1}/{2})\log x))^2$ . Note that if $0<\beta-{1}/{2}<\delta$ , then $0<({1}/{2})(\beta-{1}/{2})\log x<{1}/{2}$ . Since $y<\sinh(y)<2y$ for $0<y<{1}/{2}$ , (4·3) is
The desired result follows.
Proof of Theorem 1·3. Let $h\in C^{n+2}(\mathbb{T}^n)$ , and let $\boldsymbol{\alpha}$ satisfy (1·12). Let $J\in[1, 100\log(eT)]$ . We begin with the expansion
By Lemma 4·1, we have
Write $x_{\boldsymbol{m}}=e^{2\pi(\boldsymbol{m}\cdot\boldsymbol{\alpha})}$ . Since $x_{-\boldsymbol{m}}^{i\gamma}=x_{\boldsymbol{m}}^{-i\gamma}$ and $c_{-\boldsymbol{m}}=-c_{\boldsymbol{m}}$ , we find that (4·4) equals
Choose J so that $\|\boldsymbol{m}\|_2\leq J$ implies $\log x_{\boldsymbol{m}}<(({c}/{3})({\log T}/{\log\log T}))$ (with c as in Theorem 1·1). By Lemma 4·3 and the above display, (4·4) equals
Since $\log x_{\boldsymbol{m}}\leq 2\pi\|\boldsymbol{m}\|_2 \|\boldsymbol{\alpha}\|_2$ , it follows from our preliminary bound for J that (4·4) equals
We apply Lemma 4·2 to conclude that (4·5) equals
where $\mathcal{E}$ satisfies (note that $\log x_{\boldsymbol{m}}=2\pi(\boldsymbol{m}\cdot\boldsymbol{\alpha})$ )
Our choice of J ensures that $\log(2x_{\boldsymbol{m}})\ll \log T$ , so it follows from (1·12) that
Since $x_{\boldsymbol{m}}^{1+\theta}=e^{2\pi(1+\theta)(\boldsymbol{m}\cdot\boldsymbol{\alpha})}\leq e^{4\pi\|\boldsymbol{m}\|_2 \|\boldsymbol{\alpha}\|_2}$ , it follows that
We choose $J = (\!\log T)^{2/3}$ . Since $N_{\pi}(T)\ll T\log T$ , (4·6) equals
Observe that if $x_{\boldsymbol{m}}\neq\langle x_{\boldsymbol{m}}\rangle$ , then $|e^{iT\log\frac{x_{\boldsymbol{m}}}{\langle x_{\boldsymbol{m}}\rangle}}-1|\leq |iT\log({x_{\boldsymbol{m}}}/{\langle x_{\boldsymbol{m}}\rangle})|$ . Also, since $0\leq\theta<{1}/{2}$ , it follows that $|\Lambda_{\pi}(\langle x_{\boldsymbol{m}}\rangle)|\ll \sqrt{x_{\boldsymbol{m}}}$ . The proof of Lemma 4·1 ensures that $|c_{\boldsymbol{m}}|\ll \|\boldsymbol{m}\|_{2}^{-n-2}$ , so the sum over $\boldsymbol{m}$ converges absolutely. By the decay of $|c_{\boldsymbol{m}}|$ , our choice of J, and (4·7), (4·8) equals
In particular, each sum over $\boldsymbol{m}$ converges absolutely.
To handle the sum over $\boldsymbol{m}$ such that $x_{\boldsymbol{m}}\neq\langle x_{\boldsymbol{m}}\rangle$ , we note (by absolute convergence) that for all $\varepsilon>0$ , there exists $M_{\varepsilon}=M_{\varepsilon}({\boldsymbol \alpha},h)>0$ such that
Consequently, we have
As we let $\varepsilon\to 0$ sufficiently slowly, we conclude that (4·6) equals o(T), as desired.
For the sum over $\boldsymbol{m}$ such that $x_{\boldsymbol{m}} = \langle x_{\boldsymbol{m}} \rangle$ , which means $x_{\boldsymbol{m}} \in \mathbb{Z}$ , the terms which are not prime powers will vanish due to the presence of the von Mangoldt function. For the other terms which are prime powers, we have $\boldsymbol{m} \cdot \boldsymbol{\alpha} = {k \log p}/{2 \pi}$ for some $k \in \mathbb{N}$ by the definition of $x_{\boldsymbol{m}}$ . This will only happen when $\boldsymbol{m}$ is a multiple of $q_j \boldsymbol{b}_j$ for some $j \in \{1, \ldots, r\}$ due to our choice of the vector $\boldsymbol{\alpha}$ in (1·13), so
5. Proof of Corollary 1·4
Let $\mathbb{B}\subseteq\mathbb{T}^n$ be a product of n subintervals of $\mathbb{T}$ for which $|\int_{\mathbb{B}}g_{f,\boldsymbol{\alpha}}(\boldsymbol{t}){d}\boldsymbol{t}|$ attains its maximum. For $\varepsilon>0$ , let $\varphi_{\varepsilon}\;:\;\mathbb{T}^n\to\mathbb{R}$ satisfy the following conditions:
-
(i) $\varphi_{\varepsilon}$ is nonnegative and infinitely differentiable;
-
(ii) $\varphi_{\varepsilon}$ is supported on a compact subset of $U_{\varepsilon}\;:\!=\;\{\boldsymbol{t}\in \mathbb{T}^n\colon \|\boldsymbol{t}\|_2<\varepsilon\}$ ; and
-
(iii) $\int_{\mathbb{T}^n}\varphi_{\varepsilon}(\boldsymbol{t}){d}\boldsymbol{t}=1$ .
Let $\textbf{1}_{\mathbb{B}}$ be the indicator function of $\mathbb{B}$ , and define $h_{\varepsilon}(\boldsymbol{t})=\int_{\mathbb{T}^n}\varphi_{\varepsilon}(\boldsymbol{t})\textbf{1}_{\mathbb{B}}(\boldsymbol{x}-\boldsymbol{t}){d}\boldsymbol{t}$ . Then $h_{\varepsilon}$ is infinitely differentiable, and thus Theorem 1·3 holds with r arbitrarily large for $h=h_{\varepsilon}$ . Consequently, for any fixed $r\geq n+2$ , we have
It follows from our definition of $g_{f,\boldsymbol{\alpha}}(\boldsymbol{t})$ in (1·14) that $g_{f,\boldsymbol{\alpha}}(\boldsymbol{t})\ll 1$ , hence
for all $\boldsymbol{y}\in U_{\varepsilon}$ . Thus, we have
By the mean value theorem, there exists $\boldsymbol{y}\in U_{\varepsilon}$ such that
The proof follows once we let $\varepsilon \to 0$ sufficiently slowly as a function of T.
6. Proofs of Corollaries 1·5-1·8
Throughout Sections 6·3-6·5, all levels are assumed to be squarefree.
6·1. An estimate for the density function
We begin with a useful estimate for the density function $g_{f,\boldsymbol{\alpha}}$ associated to a holomorphic cuspidal newform $f\in S_k^{\textrm{new}}(\Gamma_0(q))$ as in Section 2.2.
Lemma 6·1. Let $f \in S_k (\Gamma_0(q))$ be a newform and let $\alpha = {a\log p}/{2 \pi q}$ . Then we have
Proof. In this case, Deligne’s bound implies that
6·2. Proof of Corollary 1·5
First, we prove a simple criterion for $f_1$ winning the $({\log p}/{2\pi}, h)$ -race. From Lemma 6·1, we obtain
Consequently, the inequality
implies that
which by Theorem 1·3 tells us that $f_1$ wins the $({\log p}/{2 \pi}, h)$ -race.
Throughout the proof, we let $ k_h \;:\!=\; \int_\mathbb{T} h(t) \cos(2 \pi t)dt$ . The number of p for which $f_1$ wins the race is equal to $T_1 - T_2 + T_3$ , where
From the symmetry of $\mu_{ST,2}$ , we see that $\nu(\{ (x,y) \in [\!-\!2,2]^2\;:\; x-y > 0\}) = {1}/{2}$ . By Corollary 2·2 with $\mathcal{I} = (0,4)$ , we have
We have to show that $T_2$ and $T_3$ are both $O( \pi(X) ({\sqrt{\log\log\log X}}/{(\!\log\log X)^{1/4}}))$ . Note that
If $p \ge \sqrt{X}$ and $f_1$ loses the $(\alpha, h)$ -race and $k_h(\lambda_{f_2}(p) - \lambda_{f_1}(p)) > 0$ , then by (6·1), we have
Denoting by $J_X$ the rightmost interval in the preceding display, we have
By Corollary 2·2, this is at most
Since $\nu(|k_h|^{-1} J_X) = O( | J_X | ) = O(X^{-1/4})$ , it follows that
If the conditions for $T_3$ are true, then
Therefore, $T_3\ll \pi(X) ({\sqrt{\log\log\log X}}/{(\!\log\log X)^{1/4}})$ by the argument used to bound $T_2$ .
The result follows from the estimate shown for $T_1$ and the bounds for $T_2$ and $T_3$ .
6·3. Proof of Corollary 1·6
Proof. From Theorem 1·3, Lemma 6·1, and (2·8) we have
Consider the statements
and
Defining
we have
By Corollary 2·2 we have:
We proceed to show that $T_2$ and $T_3$ are $O\Big( \varepsilon_X^2\pi(X) \Big)$ as $X \to \infty$ . We first examine $T_2$ . Set $\mathcal{I} = (\delta_1, \delta_2)$ . If the condition in (6·4) is false, then
Applying (6·2), we deduce that
where C is an implied constant in (6·2). Then (6·3) gives us
Since $p \in [(1-\varepsilon_X)X, X]$ , it follows that $(\lambda_{f_2}(p) - \lambda_{f_1} (p)) \in I_X$ , where $I_X$ is
By Corollary 2·2, we have
From the prime number theorem, we obtain
Combining this with the fact that $\nu(I_X) = O(\varepsilon_X)$ , we conclude the following:
Therefore, $T_2 = O( \varepsilon_X^2 \pi(X))$ . A very similar argument can be used to bound $T_3$ . More specifically, if (6·4) holds, then we have
If (6·3) fails and (6·4) holds, then, much like (6·5), we obtain
By Corollary 2·2, the number of such $p \in ((1-\varepsilon_X)X, X)$ is $O\Big( \varepsilon_X^2\pi(X) \Big)$ .
6·4. Proof of Corollary 1·7
By Theorem 1·3, if $f_2$ wins the $(\alpha, h)$ -race, where $\alpha = {\log p}/{2\pi}$ , then we must have
By Lemma 6·1, we have
It follows that $\log(q_1/q_2) \le 2 p^{-1/2}( 1 + (p^{ \frac{1}{2} } - 1 )^{-1})\log{p}$ . The left hand side is independent of $\alpha$ and positive, while the right hand side tends to zero as p grows. Thus, this inequality holds for only finitely many primes p.
6·5. Proof of Corollary 1·8
Fix $t_0 \in [0,1)$ , and $k_1, k_2 \in \mathbb{Z}$ . By Theorem 1·3 and the same reasoning as in [ Reference Liu and ZaharescuLZ21 , theorem 1·2], that $f_1$ wins the local $(\alpha, t_0)$ -race against $f_2$ if
or, equivalently, if $({1}/{2\pi}) \log(q_1/q_2) > g_{f_2, \alpha}(t_0) - g_{f_1, \alpha}(t_0)$ .
We first assume $t_0 \neq {1}/{4}, {3}/{4}$ . For $\alpha = {\log p}/{2 \pi}$ and $q_2 \in (q_1, q_1 + \sqrt{q_1})$ , by Lemma 6·1, the following is sufficient to guarantee that $f_1$ wins the $(\alpha,t_0)$ race:
This inequality is automatically true if both
and
are satisfied.
Now set $X = q_1^{1/4} $ and $Y = q_1^{1/6} $ . For $p \in [Y,X)$ , if $q_1$ is sufficiently large, then (6·8) is satisfied. Choose $I_1, I_2 \subseteq [\!-\!2, 2]$ such that $(\lambda_{f_1}(p) ,\lambda_{f_2}(p)) \in I_1 \times I_2$ implies (6·7). Suppose $f_1 \in S_{k_1}^{new}(\Gamma_0(q_1)), f_2 \in S_{k_2}^{new} (\Gamma_0(q_2))$ are non-CM newforms, where $q_2 \in [q_1, q_1 + \sqrt{q_1}]$ , and assume, as in the statement of the Corollary, that $f_2 \neq f_1 \otimes \chi$ for any primitive Dirichlet character $\chi$ . Let
Then by Theorem 2·1 we have
So if $q_1$ is sufficiently large, then $\pi_{f_1, f_2, I_1, I_2}(X) > \pi_{f_1, f_2, I_1, I_2}(Y)$ . So there exists p between Y and X such that (6·6) is satisfied and therefore $f_1$ wins the local $(t_0, {\log p}/{2\pi})$ race.
Finally, if $t_0 = {1}/{4}, {3}/{4}$ , then instead of (6·6), we wish to find p such that
We obtain this the same way as the first case.
7. Example
We conclude with a numerical example to illustrate (1·15). For our example, we consider the L-function $L(s,\Delta)$ associated to the discriminant modular form
where $\tau(n)$ denotes the Ramanujan tau function. We use Rubinstein’s lcalc package [ Reference RubinsteinRub14 ] to calculate the $2\cdot 10^5$ nontrivial zeros $L(s, \Delta)$ up to height $T = 74920.77$ .
Let M and $\boldsymbol{\alpha}$ satisfy the following relation for (1·13),
so that (1·14) will define our density function $g_{\Delta, \boldsymbol{\alpha}}(x,y)$ . We graph $g_{\Delta, \boldsymbol{\alpha}}(x,y)$ in Figure 1(a) below. Next, we partition the unit square $[0, 1) \times [0, 1)$ as
Given $(x,y)\in[0,1)\times[0,1)$ , there exists a unique pair of integers a and b with $0\leq a,b\leq 29$ such that $(x,y)\in S_{a,b}$ . Denoting this unique square as S(x,y), we define
This gives us a discretised approximation to $g_{\Delta,\boldsymbol{\alpha}}(x,y)$ , which we plot in Figure 1(b) below.