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

An induction principle for the Bombieri-Vinogradov theorem over 𝔽q[t]{\mathbb{F}}_{q}[t] and a variant of the Titchmarsh divisor problem.

Sampa Dey Department of Mathematics, Indian Institute of Technology Gandhinagar, Gandhinagar, Gujarat 382355, India [email protected]  and  Aditi Savalia Department of Mathematics, Indian Institute of Technology Gandhinagar, Gandhinagar, Gujarat 382355, India [email protected]
Abstract.

Let 𝔽q[t]{\mathbb{F}}_{q}[t] be the polynomial ring over the finite field 𝔽q{\mathbb{F}}_{q}. For arithmetic functions ψ1,ψ2:𝔽q[t]\psi_{1},\psi_{2}:{\mathbb{F}}_{q}[t]\rightarrow\mathbb{C}, we establish that if a Bombieri-Vinogradov type equidistribution result holds for ψ1\psi_{1} and ψ2\psi_{2}, then it also holds for their Dirichlet convolution ψ1ψ2\psi_{1}\ast\psi_{2}. As an application of this, we resolve a version of the Titchmarsh divisor problem in 𝔽q[t]{\mathbb{F}}_{q}[t]. More precisely, we obtain an asymptotic for the average behaviour of the divisor function over shifted products of two primes in 𝔽q[t]{\mathbb{F}}_{q}[t].

2010 Mathematics Subject Classification:
Primary 11N37; Secondary 11T55, 11N36
Keywords and phrases. finite fields, function fields, divisor function, Bombieri-Vinogradov theorem, large sieve inequality, Titchmarsh divisor problem

1. Introduction

The Bombieri-Vinogradov theorem is one of the most celebrated theorems in analytic number theory, concerned with the equidistribution property of primes in arithmetic progressions. To articulate the theorem precisely, we first set up some notation. Let aa and dd be coprime integers. Denote

π(x;d,a):=#{px:pa(modd)},\pi(x;d,a):=\#\left\{p\leq x:p\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\right\},

where pp represents a prime number. Assuming the Generalized Riemann Hypothesis (GRH) for Dirichlet LL-functions, one can get that for any dxd\leq x,

π(x;d,a)=1ϕ(d)Li(x)+O(x1/2log(x)),\pi(x;d,a)=\frac{1}{\phi(d)}\operatorname{Li}(x)+O\left(x^{1/2}\log(x)\right),

where Li(x)\operatorname{Li}(x) is the usual logarithmic integral

Li(x):=2xdtlogt.\displaystyle\operatorname{Li}(x):=\int\limits_{2}^{x}\frac{dt}{\log t}.

An unconditional result in this context is the well-known Siegel-Walfisz theorem which asserts that for any N>0N>0, there exists c(N)>0c(N)>0 such that, if d(logx)Nd\leq(\log x)^{N},

π(x;d,a)=1ϕ(d)Li(x)+O(xexp(c(N)(logx)1/2)),\pi(x;d,a)=\frac{1}{\phi(d)}\operatorname{Li}(x)+O\left(x\exp\left(-c(N)(\log x)^{1/2}\right)\right),

uniformly in dd. In other words, a non-trivial upper bound on the error term

E(x;d,a):=π(x;d,a)1ϕ(d)Li(x)E(x;d,a):=\pi(x;d,a)-\frac{1}{\phi(d)}\operatorname{Li}(x)

is unconditionally known in the range d(logx)Nd\leq(\log x)^{N}, for any N>0N>0. In 1965, Bombieri and Vinogradov independently proved that for any A>0A>0, there exists B=B(A)>0B=B(A)>0 such that

dx1/2(logx)Bmax(a,d)=1maxyx|E(y;d,a)|Ax(logx)A.\sum\limits_{d\leq\frac{x^{1/2}}{(\log x)^{B}}}\max_{(a,d)=1}\max_{y\leq x}\big{|}E(y;d,a)\big{|}\ll_{A}\frac{x}{(\log x)^{A}}. (1.1)

More generally, an arithmetic function ff is said to have level of distribution θ\theta, if for any A>0A>0, there exists B=B(A)>0B=B(A)>0 such that

dxθ(logx)Bmax(a,d)=1maxyx|nyna(modd)f(n)1ϕ(d)ny(n,d)=1f(n)|Ax(logx)A.\sum\limits_{d\leq\frac{x^{\theta}}{(\log x)^{B}}}\max_{(a,d)=1}\max_{y\leq x}\bigg{|}\sum\limits_{\begin{subarray}{c}n\leq y\\ n\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}f(n)-\frac{1}{\phi(d)}\sum\limits_{\begin{subarray}{c}n\leq y\\ (n,d)=1\end{subarray}}f(n)\bigg{|}\ll_{A}\frac{x}{(\log x)^{A}}. (1.2)

The Bombieri- Vinogradov theorem asserts that the level of distribution for the prime indicator function is θ=12\theta=\frac{1}{2}. Further, it was conjectured by Elliott and Halberstam that the bound (1.2) holds for all 0<θ<10<\theta<1 for this function. It is worth noting that Bombieri-Vinogradov theorem yields a bound as strong as what would follow from GRH for Dirichlet LL-functions. Pushing the level of distribution beyond half has been an active area of research, resulting in important contributions due to Fouvry and Iwaniec [15] [16] , Bombieri, Friedlander and Iwaniec [6], [7], [8], Zhang [43], Maynard [31], [32], [33], Granville and Shao [19] and many others.

In 1976, Y. Motohashi [34] proved an interesting induction principle for the Bombieri-Vinogradov theorem. For any two arithmetic functions ff and gg, their multiplicative convolution or Dirichlet product is defined as

fg(n)=ab=nf(a)g(b).f\ast g(n)=\sum\limits_{ab=n}f(a)g(b).

For an arithmetic function ff, we consider the following three properties:

  • (a)

    f(n)=O(τ(n)C)f(n)=O(\tau(n)^{C}) for some fixed C>0C>0.

  • (b)

    Let χ\chi be a non-principal character modulo dd such that the conductor of χ\chi is of order O((logx)D)O((\log x)^{D}), for D>0D>0 suitably large. Then

    nxf(n)χ(n)=O(x(logx)3D).\sum\limits_{n\leq x}f(n)\chi(n)=O\left(\frac{x}{(\log x)^{3D}}\right).
  • (c)

    The function ff satisfies the Bombieri-Vinogradov type equidistribution property, that is, (1.2) holds with θ=12\theta=\frac{1}{2}.

Motohashi proved that if ff and gg satisfy properties (a)(a), (b)(b) and (c)(c), then their multiplicative convolution fgf\ast g also satisfies these three properties.

Recently, Darbar and Mukhopadhyay [10] generalized Motohashi’s result to imaginary quadratic fields. In this paper, we establish an analogous induction principle for equidistribution in arithmetic progressions, in the setting of 𝔽q[t]{\mathbb{F}}_{q}[t]. While it is true that the Riemann hypothesis is known over finite fields, theorems of Bombieri-Vinogradov type are relevant as they give information about equidistribution in arithmetic progressions for a variety of functions. For instance, our main result allows us to prove equidistribution in arithmetic progressions for almost primes as well as the kk-fold divisor function. Pushing the level of distribution beyond half remains an important question in the function field setting as well as demonstrated by recent work due to Sawin [37], [38, Theorem 1.2], as well as Sawin and Shusterman [39, Theorem 1.7].

We proceed to state our main result below after setting up relevant notation. Let 𝔽q{\mathbb{F}}_{q} be a finite field of order qq and 𝔽q[t]{\mathbb{F}}_{q}[t] be the polynomial ring defined over 𝔽q{\mathbb{F}}_{q}. We denote the degree of a polynomial ff in 𝔽q[t]{\mathbb{F}}_{q}[t] by deg(f)\text{deg}(f), and define the norm of a polynomial |f||f| as qdeg(f)q^{\text{deg}(f)}. Throughout the article, we consider f,g,hf,g,h to be monic polynomials in 𝔽q[t]{\mathbb{F}}_{q}[t]. Let τk(f)\tau_{k}(f) denote the kk-fold divisor function which counts the number of ways to write ff as a product of kk monic polynomials. When k=2k=2, the function τ2(f)\tau_{2}(f) is the usual divisor function τ(f)\tau(f) which counts the number of monic polynomials dividing ff.

Let mm be a non-constant polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t]. We may denote the ideal (m)(m) by mm without mentioning this explicitly if the usage is clear from the context. A multiplicative character χ\chi modulo mm is a complex valued multiplicative function on (𝔽q[t]/(m))\left({\mathbb{F}}_{q}[t]/(m)\right)^{\ast}. We extend χ\chi to 𝔽q[t]{\mathbb{F}}_{q}[t] by putting χ(f)=χ(f¯)\chi(f)=\chi(\overline{f}), where ff¯(modm)f\equiv\overline{f}\pmod{m}, and zero otherwise. A principal character χ0\chi_{0} mod mm is defined by the property that χ0(f)=1\chi_{0}(f)=1 if (f,m)=1(f,m)=1 and 0 otherwise. A character χ\chi mod mm is called primitive multiplicative character if there is no d|md|m such that (d)(m)(d)\neq(m) and χ\chi is induced by a character mod dd. For a character χ\chi modulo mm, a monic polynomial d|md|m is called the conductor of χ\chi if there is no d|dd^{\prime}|d such that χ\chi is induced by a primitive character mod dd^{\prime}. For more details we refer the reader to [36, ch. 4] and [27].

For arithmetic functions ψ1,ψ2:𝔽q[t]\psi_{1},\psi_{2}:{\mathbb{F}}_{q}[t]\rightarrow\mathbb{C}, their multiplicative convolution denoted by ψ1ψ2\psi_{1}\ast\psi_{2} is defined as

ψ1ψ2(f)=gh=fψ1(g)ψ2(h).\displaystyle\psi_{1}\ast\psi_{2}(f)=\sum_{gh=f}\psi_{1}(g)\psi_{2}(h).

For an arithmetic function ψ:𝔽q[t]\psi:{\mathbb{F}}_{q}[t]\rightarrow\mathbb{C}, consider the following properties:

  1. (1)

    (growth condition) ψ(f)=O(τ(f)C)\psi(f)=O(\tau(f)^{C}) for some C>0C>0.

  2. (2)

    (Siegel-Walfisz bound) For a non-principal character χ\chi with conductor of degree DlogN\leq D\log N, we have

    deg(f)=Nψ(f)χ(f)=O(qNN3D),\sum\limits_{\deg(f)=N}\psi(f)\chi(f)=O\left(\frac{q^{N}}{N^{3D}}\right),

    for D>0D>0 sufficiently large.

  3. (3)

    (Bombieri-Vinogradov type equidistribution property) For m,𝔽q[t]m,\ell\in{\mathbb{F}}_{q}[t] with mm monic, and (m,)=1(m,\ell)=1, let

    E(M;m,l;ψ):=fl(modm)deg(f)=Mψ(f)1ϕ(m)(f,m)=1deg(f)=Mψ(f).E(M;m,l;\psi):=\sum\limits_{\begin{subarray}{c}f\equiv l\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)\\ \deg(f)=M\end{subarray}}\psi(f)-\frac{1}{\phi(m)}\sum\limits_{\begin{subarray}{c}(f,m)=1\\ \deg(f)=M\end{subarray}}\psi(f). (1.3)

    Then for any A>0A>0, there exists B=B(A)>0B=B(A)>0 such that

    deg(m)N2BlogNmaxMNmax(m,l)=1|E(M;m,l;ψ)|AqNNA.\sum\limits_{\deg(m)\leq\frac{N}{2}-B\log N}\max\limits_{M\leq N}\max\limits_{(m,l)=1}|E(M;m,l;\psi)|\ll_{A}\frac{q^{N}}{N^{A}}.
Theorem 1.1.

Fix qq. Let ψ1\psi_{1}, ψ2\psi_{2} be arithmetic functions on 𝔽q[t]{\mathbb{F}}_{q}[t]. If ψ1\psi_{1} and ψ2\psi_{2} have the properties (1), (2) and (3), then the Dirichlet convolution ψ1ψ2\psi_{1}\ast\psi_{2} also satisfies (1), (2) and (3).

An important distinction that presents itself in our function-field analogue is that we derive the corresponding version of the large sieve inequality required for our proof. Classically, it is well-known that the large sieve inequality plays a pivotal role in proving results of this type. In the function field case, various versions of the large sieve inequality have been developed, for instance by Hsu [23] as well as Baier and Singh ([3], [4]). While versions of the large sieve inequality with additive characters are known in the function field case, for our purpose, we need a large sieve estimate involving multiplicative characters. A version of this has recently been given by Klurman, Mangerel and Teräväinen ([28, Lemma 4.2]). As we shall see in Section 3 (see the remark following Theorem 3.4), the estimate in [28] does not suffice for us and we require an upper bound which is better on average over Dirichlet characters χ\chi. We proceed to prove this in Section 3 (see Theorem 3.5).

Finally, as an application, we also obtain an asymptotic for the average behaviour of the number of divisors over shifts of products of two primes in 𝔽q[t]{\mathbb{F}}_{q}[t]. It is worth noting that this requires us to invoke a function field analogue of the Brun-Titchmarsh inequality, proved by Hsu in [23]. It is conceivable that the method should extend to shifts of products of kk-primes, where kk is fixed, though we do not do this here.

A direct consequence is the following corollary which we obtain upon using Theorem 1.1 iteratively.

Corollary 1.2.

Let ψi\psi_{i}, for i=1,2,,ni=1,2,...,n, be arithmetic functions on 𝔽q[t]{\mathbb{F}}_{q}[t] such that each of them satisfies properties (1), (2) and (3). Then the Dirichlet convolution ψ1ψ2ψn\psi_{1}\ast\psi_{2}\ast...\ast\psi_{n} also does so.

Motohashi’s generalization is significant in terms of yielding a family of arithmetic functions for which equidistribution results now become available. In the setting of 𝔽q[t]{\mathbb{F}}_{q}[t] as well, we find such interesting applications.

Let π(N)\pi(N) denote the number of monic irreducible polynomials of degree NN in 𝔽q[t]{\mathbb{F}}_{q}[t]. The prime number theorem (cf. [36], Theorem 2.2) in 𝔽q[t]{\mathbb{F}}_{q}[t] gives

π(N)=qNN+O(qN/2N).\displaystyle\pi(N)=\frac{q^{N}}{N}+O\left(\frac{q^{N/2}}{N}\right). (1.4)

We also have the prime number theorem for arithmetic progressions (cf. [36], Theorem 4.8) stated as follows. Let d,a𝔽q[t]d,a\in{\mathbb{F}}_{q}[t], dd has positive degree and (a,d)=1(a,d)=1. Then the number of monic irreducible polynomials of degree NN in 𝔽q[t]{\mathbb{F}}_{q}[t] in the arithmetic progression a(modd)a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud) is given by

π(N;d,a)=qNϕ(d)N+O(qN/2N).\displaystyle\pi(N;d,a)=\frac{q^{N}}{\phi(d)N}+O\left(\frac{q^{N/2}}{N}\right). (1.5)

Thus, the primes are equidistributed in arithmetic progressions with a level of distribution θ=12\theta=\frac{1}{2}. A natural question that arises is about the level of distribution of products of two primes in arithmetic progressions. Let 𝟙𝒫\mathbbm{1}_{\mathcal{P}} denote the prime indicator function. Applying Theorem 1.1 on 𝟙𝒫\mathbbm{1}_{\mathcal{P}}, we have the following result for the indicator function of the product of two primes.

Corollary 1.3.

Over the polynomial ring 𝔽q[t]{\mathbb{F}}_{q}[t], the Dirichlet convolution 𝟙𝒫𝟙𝒫\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}} satisfies all the three properties (1), (2) and (3).

Now, taking ψ1\psi\equiv 1, it is easy to see that ψ\psi has level of distribution 11 (see Section 5), and satisfies properties (1), (2). So, our Theorem 1.1 gives that τk\tau_{k} has Bombieri-Vinogradov type inequality, for each kk. We state this as the following corollary.

Corollary 1.4.

For each kk\in\mathbb{N}, the kk-fold divisor function τk\tau_{k} satisfies properties 1, 2 and 3.

One of the significant applications of the classical Bombieri-Vinogradov theorem is to the celebrated Titchmarsh divisor problem. In 1930, Titchmarsh [41] proved that

pxτ(pa)C1x\sum\limits_{p\leq x}\tau(p-a)\sim C_{1}x

for a fixed integer aa and some constant C1>1C_{1}>1 under the assumption of GRH. It was only after more than three decades that an unconditional proof of this result was obtained by Linnik [30]. In [6], Bombieri, Friedlander and Iwaniec proved a version of the theorem with arbitrary logx\log x savings. The error term in the dispersion method was further improved by Drappeau [11] to obtain power savings under GRH. Several variants of the above sum have been studied by Rodriques [35], Halberstam [20], Fouvry [14], Akbary and Ghioca [1], Felix [13], Vatwani and Wong [42] and others. Using the Bombieri-Vinogradov theorem, one can show the following (cf. [9], Theorem 9.3.19.3.1). For a fixed aa, there exists a positive constant cc such that

pxτ(pa)=cx+O(xloglogxlogx).\sum\limits_{p\leq x}\tau(p-a)=cx+O\left(\frac{x\log\log x}{\log x}\right).

In [34], Motohashi generalized this divisor problem to products of kk-primes. In the same year, Fuji [17] proved that

p1p2xτ(p1p21)=2ζ(2)ζ(3)ζ(6)xloglogx+O(x),\sum\limits_{p_{1}p_{2}\leq x}\tau(p_{1}p_{2}-1)=2\frac{\zeta(2)\zeta(3)}{\zeta(6)}x\log\log x+O(x), (1.6)

where ζ(s)\zeta(s) denotes the Riemann zeta function. Drappeau and Topacogullari [12] studied analogous sums over integers with a fixed number of distinct prime divisors. More precisely, letting N1N\geq 1 and ϵ>0\epsilon>0, they showed that there exists a constant δ>0\delta>0 and polynomials Ph,k(X)P_{h,\ell}^{k}(X) of degree k1k-1 such that, for 1kloglogx1\leq k\ll\log\log x and |h|xδ|h|\leq x^{\delta},

|h|<nxω(n)=kτ(nh)=x0NPh.k(loglogx)(logx)+O(x(loglogx)kk!(logx)N+1ϵ).\displaystyle\sum_{\begin{subarray}{c}|h|<n\leq x\\ \omega(n)=k\end{subarray}}\tau(n-h)=x\sum_{0\leq\ell\leq N}\frac{P_{h.\ell}^{k}(\log\log x)}{(\log x)^{\ell}}+O\left(\frac{x(\log\log x)^{k}}{k!(\log x)^{N+1-\epsilon}}\right).

Here ω(n)\omega(n) denotes the number of distinct prime divisors of an integer nn and the implicit constants depend only on NN and ϵ\epsilon. Taking k=2k=2 in their result gives a close analogue of (1.6). Darbar and Mukhopadhyay extended (1.6) to imaginary quadratic number fields (cf. [10], Theorem 1.6). Analogues of this for function fields have been studied extensively in the literature. Let PP denote a monic irreducible polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t] and let aa be a fixed non-zero polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t]. Then for a fixed qq, as NN\rightarrow\infty, Hsu [24] proved that

deg(P)=Nτ(Pa)=P|a(11|P|)(1+1|P|(|P|1))1ζq(2)ζq(3)ζq(6)qN+O(qNlogNN),\sum\limits_{\begin{subarray}{c}\deg(P)=N\end{subarray}}\tau(P-a)=\prod\limits_{P|a}\left(1-\frac{1}{|P|}\right)\left(1+\frac{1}{|P|(|P|-1)}\right)^{-1}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}q^{N}+O\left(\frac{q^{N}\log N}{N}\right),

where the implied constant depends only on aa. Here ζq(s)\zeta_{q}(s) is the zeta function over 𝔽q[t]{\mathbb{F}}_{q}[t], defined by

ζq(s)=P(11|P|s)1\displaystyle\zeta_{q}(s)=\prod_{P}\left(1-\frac{1}{|P|^{s}}\right)^{-1} (1.7)

for Re(s)>1\mathrm{Re}(s)>1, where the product runs over all the monic irreducible polynomials in 𝔽q[t]{\mathbb{F}}_{q}[t].

If we keep NN fixed and let qq\rightarrow\infty, then Andrade, Bary-Soroker and Rudnick [2] obtained the asymptotic formula

deg(P)=Nτ(Pa)=qN+qNN+ON(qN12).\sum\limits_{\begin{subarray}{c}\deg(P)=N\end{subarray}}\tau(P-a)=q^{N}+\frac{q^{N}}{N}+O_{N}\left(q^{N-\frac{1}{2}}\right).

As an application of the induction principle in the setting of 𝔽q[t]{\mathbb{F}}_{q}[t], we can obtain a generalization of Hsu’s result to products of mm-primes where mm is fixed. In particular, we obtain the following analogue of the Titchmarsh divisor problem over 𝔽q[t]{\mathbb{F}}_{q}[t].

Theorem 1.5.

Fix qq. Let aa be a fixed non zero polynomial and PiP_{i} denote a monic irreducible polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t]. Then as NN\rightarrow\infty, we have

deg(P1P2)=Nτ(P1P2a)=2Caζq(2)ζq(3)ζq(6)qNlogN+O(qNloglogN),\sum\limits_{\deg(P_{1}P_{2})=N}\tau(P_{1}P_{2}-a)=2C_{a}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}q^{N}\log N+O\left(q^{N}\log\log N\right),

where ζq(s)\zeta_{q}(s) is the zeta function over 𝔽q[t]{\mathbb{F}}_{q}[t] and the constant CaC_{a} is given by the product

Ca=P|a(11|P|)(1+1|P|(|P|1))1.\displaystyle C_{a}=\prod\limits_{P|a}\left(1-\frac{1}{|P|}\right)\left(1+\frac{1}{|P|(|P|-1)}\right)^{-1}. (1.8)

In case we allow both NN and qq to go to infinity, we get the following version of the above result.

Theorem 1.6.

Let aa be a fixed non zero polynomial and PiP_{i} denote a monic irreducible polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t]. Then as N,qN,q\rightarrow\infty, we have

deg(P1P2)=Nτ(P1P2a)=2qN(logN+γ)+O(qNN(logN)2)+O(qN12logN).\sum\limits_{\deg(P_{1}P_{2})=N}\tau(P_{1}P_{2}-a)=2q^{N}(\log N+\gamma)+O\left(\frac{q^{N}}{N}(\log N)^{2}\right)+O\left(q^{N-\frac{1}{2}}\log N\right).

Moreover, letting qq be fixed and NN\to\infty, it is possible to extend the above formulas for products of mm primes to get

deg(P1P2Pm)=Nτ(P1P2Pma)c(m)qN(logN)m1,\displaystyle\sum_{\deg(P_{1}P_{2}...P_{m})=N}\tau(P_{1}P_{2}...P_{m}-a)\sim c(m)q^{N}(\log N)^{m-1},

where c(m)c(m) is a constant independent of NN. This is a tedious but straightforward modification of the proof of Theorem 1.5 which we leave to the reader.

The paper is organized as follows. In Section 2, we prove some basic results on the divisor function and state a version of Perron’s formula over 𝔽q[t]{\mathbb{F}}_{q}[t]. In Section 3, we establish a large sieve inequality for multiplicative characters which will be needed to prove our main theorem. The proofs of Theorem 1.1 and Corollary 1.4 are contained in Sections 4 and 5 respectively. By invoking an 𝔽q[t]{\mathbb{F}}_{q}[t]-analogue of the Brun-Titchmarsh inequality, we prove Theorem 1.5 and Theorem 1.6 in Section 6.

2. Preliminaries

In this section, we state some lemmas which will be useful to prove the main theorems in this paper.

Let k2k\geq 2. For the kk-fold divisor function, we have ([2, Lemma 2.2])

deg(f)=Nf-monicτk(f)=(N+k1k1)qN.\displaystyle\sum_{\begin{subarray}{c}\text{deg}(f)=N\\ f\text{-monic}\end{subarray}}\tau_{k}(f)={N+k-1\choose k-1}q^{N}.

For our purpose we will use the following inequality which is a direct consequence of the above result.

Lemma 2.1.

Let k2k\geq 2. Then for the kk-fold divisor function we have

  1. (i)
    deg(f)=Nf-monicτk(f)Nk1qN,\displaystyle\sum_{\begin{subarray}{c}\text{deg}(f)=N\\ f\text{-monic}\end{subarray}}\tau_{k}(f)\ll N^{k-1}q^{N},
  2. (ii)
    deg(f)Nτk(f)qdeg(f)Nk.\displaystyle\sum_{\text{deg}(f)\leq N}\frac{\tau_{k}(f)}{q^{\text{deg}(f)}}\ll N^{k}.

Further, the following lemma allows us to bound (τ(f))d(\tau(f))^{d} by τk(f)\tau_{k}(f) for suitably large values of kk.

Lemma 2.2.

For any fixed positive integer dd, there exists k=k(d)>0k=k(d)>0 such that

(τ(f))dτk(f),(\tau(f))^{d}\leq\tau_{k}(f),

for all f𝔽q[t]f\in{\mathbb{F}}_{q}[t]. In fact this holds for any k(d+1)!k\geq(d+1)!.

Proof.

Since τk\tau_{k} is multiplicative, it is enough to prove the inequality for powers of monic irreducible polynomials. Let PP be a monic irreducible polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t], and k(d+1)!k\geq(d+1)! . We have

τk(Pa)\displaystyle\tau_{k}(P^{a}) =\displaystyle= (a+k1k1)\displaystyle{a+k-1\choose k-1}
\displaystyle\geq ((a+k1)(a+k2)(a+d+1)(k1)!)(a+1)d\displaystyle\bigg{(}\frac{(a+k-1)(a+k-2)...(a+d+1)}{(k-1)!}\bigg{)}(a+1)^{d}
=\displaystyle= C(a,k)(a+1)d(say).\displaystyle C(a,k)(a+1)^{d}\qquad\text{(say).}

Note that, (a+1)d=(τ(Pa))d(a+1)^{d}=(\tau(P^{a}))^{d}. We will be done if we can show that the factor C(a,k)C(a,k) is at least 11. For a given kk, it is easy to see that C(a,k)C(1,k)C(a,k)\geq C(1,k) and that

C(1,k)\displaystyle C(1,k) =k(d+1)!1,\displaystyle=\frac{k}{(d+1)!}\geq 1,

for our choice of kk. ∎

In [24], Hsu proved an asymptotic bound on the sum of 1ϕ(f)\frac{1}{\phi(f)} over monic polynomials. Let ζq(s)\zeta_{q}(s) be the zeta function over 𝔽q[t]{\mathbb{F}}_{q}[t] as in (1.7). Using the identity

P𝔽q[t],monic, irreducible(1+1|P|(|P|1))=ζq(2)ζq(3)ζq(6),\prod\limits_{\begin{subarray}{c}P\in{\mathbb{F}}_{q}[t],\\ \text{monic, irreducible}\end{subarray}}\left(1+\frac{1}{|P|(|P|-1)}\right)=\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)},

we record Hsu’s result here.

Lemma 2.3.

[24, Lemma 3.1] Let g𝔽q[t]g\in{\mathbb{F}}_{q}[t] be non-zero. We have

deg(f)N(f,g)=11ϕ(f)=Cgζq(2)ζq(3)ζq(6)N+O(1),\displaystyle\sum_{\begin{subarray}{c}\text{deg}(f)\leq N\\ (f,g)=1\end{subarray}}\frac{1}{\phi(f)}=C_{g}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}N+O(1),

where CgC_{g} is as defined in (1.8) and the implicit constant depends only on gg.

We will be using this lemma with g=1g=1 or with gg being a fixed polynomial throughout this paper. For the sake of convenience of the reader, we remark that there is a minor typo in Lemma 3.1 of [24]. In the main term on the right-hand side, the product should be over pI,pgp\in I,p\nmid g, where II denotes the set of monic irreducible polynomials in 𝔽q[t]{\mathbb{F}}_{q}[t].

We will also use the following 𝔽q[t]{\mathbb{F}}_{q}[t] analogue of the classical Brun-Titchmarsh inequality concerning the number of primes in an arithmetic progression, derived by Hsu in [23].

Lemma 2.4.

[23, Lemma 4.3] Let aa, bb be non-zero polynomials in 𝔽q[t]{\mathbb{F}}_{q}[t] with a monic, (a,b)=1(a,b)=1 and deg(a)>deg(b)0\deg(a)>\deg(b)\geq 0. Then for any positive integer N>deg(a)N>\deg(a),

π(N;a,b)2qNϕ(a)(Ndeg(a)+1),\displaystyle\pi(N;a,b)\leq 2\frac{q^{N}}{\phi(a)(N-\deg(a)+1)},

where π(N;a,b)\pi(N;a,b) denotes the number of monic irreducible polynomials f𝔽q[t]f\in{\mathbb{F}}_{q}[t] such that deg(f)=N\deg(f)=N and fb(moda)f\equiv b\mkern 4.0mu({\operator@font mod}\mkern 6.0mua).

Next, we derive what can be thought of as some version of Perron’s formula over 𝔽q[t]{\mathbb{F}}_{q}[t]. This is derived using Cauchy’s residue theorem and will play a crucial role in proving our main theorem.

Lemma 2.5.

Let fmonica(f)|f|s\sum\limits_{f\,\text{monic}}\frac{a(f)}{|f|^{s}} be a Dirichlet series with a(f)|f|ϵa(f)\ll|f|^{\epsilon}, for any ϵ>0\epsilon>0. Let NN\in\mathbb{R}\setminus\mathbb{Z}. Then for σ>1\sigma>1, and any MNM\geq N,

deg(f)Na(f)=12πiσiTσ+iTdeg(f)Ma(f)|f|sqNssds+O(qσNT).\displaystyle\sum_{\text{deg}(f)\leq N}a(f)=\frac{1}{2\pi i}\int_{\sigma-iT}^{\sigma+iT}\sum_{\text{deg}(f)\leq M}\frac{a(f)}{|f|^{s}}\frac{q^{Ns}}{s}ds+O\left(\frac{q^{\sigma N}}{T}\right).
Proof.

Let N=N0\lfloor N\rfloor=N_{0}. Note that

deg(f)Na(f)=N0+12a(f).\displaystyle\sum_{\deg(f)\leq N}a(f)=\sum_{N_{0}+\frac{1}{2}}a(f).

Hence without loss of generality we assume that N=N0+12N=N_{0}+\frac{1}{2}. We have

σiTσ+iTdeg(f)Ma(f)|f|sqNssds\displaystyle\int_{\sigma-iT}^{\sigma+iT}\sum_{\text{deg}(f)\leq M}\frac{a(f)}{|f|^{s}}\frac{q^{Ns}}{s}ds =deg(f)Ma(f)σiTσ+iT(qN/|f|)ss𝑑s.\displaystyle=\sum_{\text{deg}(f)\leq M}a(f)\int_{\sigma-iT}^{\sigma+iT}\frac{(q^{N}/|f|)^{s}}{s}ds. (2.1)

Writing the above integral as σiTσ+iTxss𝑑s\displaystyle\int_{\sigma-iT}^{\sigma+iT}\frac{x^{s}}{s}ds, we have from (2.7)(2.7) on p. 219219 of Tenenbaum [40],

12πiσiTσ+iT(qN/|f|)ss𝑑s=h(qN|f|)+O(qσN|f|σ(1+T|log(qN/|f|)|)),\displaystyle\frac{1}{2\pi i}\int_{\sigma-iT}^{\sigma+iT}\frac{(q^{N}/|f|)^{s}}{s}ds=h\left(\frac{q^{N}}{|f|}\right)+O\left(\frac{q^{\sigma N}}{|f|^{\sigma}\big{(}1+T|\log(q^{N}/|f|)|\big{)}}\right),

where,

h(x)={1 if x>11/2 if x=10 if 0<x<1.\displaystyle h(x)=\left\{\begin{array}[]{cc}1&\text{ if }x>1\\ 1/2&\text{ if }x=1\\ 0&\text{ if }0<x<1.\end{array}\right.

Since qN0+12|f|\frac{q^{N_{0}+\frac{1}{2}}}{|f|} is never 11, we obtain

12πideg(f)Ma(f)σiTσ+iT(qN/|f|)ss𝑑s\displaystyle\frac{1}{2\pi i}\sum_{\text{deg}(f)\leq M}a(f)\int_{\sigma-iT}^{\sigma+iT}\frac{(q^{N}/|f|)^{s}}{s}ds =deg(f)Na(f)\displaystyle=\sum_{\text{deg}(f)\leq N}a(f)
+O(qσNdeg(f)M|a(f)||f|σ(1+T|log(qN/|f|))).\displaystyle+O\left(q^{\sigma N}\sum_{\text{deg}(f)\leq M}\frac{|a(f)|}{|f|^{\sigma}\big{(}1+T|\log(q^{N}/|f|)\big{)}}\right). (2.2)

We now analyze the error term above in more detail. If |f|qN0|f|\leq q^{N_{0}}, we find that the logarithm in (2.2) is bounded below as follows.

log(qN|f|)=log(qN0+12|f|)12logq.\log\left(\frac{q^{N}}{|f|}\right)=\log\left(\frac{q^{N_{0}+\frac{1}{2}}}{|f|}\right)\geq\frac{1}{2}\log q.

Similarly if |f|>qN0+1|f|>q^{N_{0}}+1, we can again see that

log(qN|f|)12logq.\log\left(\frac{q^{N}}{|f|}\right)\geq\frac{1}{2}\log q.

Thus, the logarithm term in (2.2) is always 1\gg 1, yielding

degfNa(f)=σiTσ+iTdeg(f)Ma(f)|f|sqNssds+O(qσNdegfM|a(f)||f|σ(11+T)),\displaystyle\sum_{\deg f\leq N}a(f)=\int_{\sigma-iT}^{\sigma+iT}\sum_{\text{deg}(f)\leq M}\frac{a(f)}{|f|^{s}}\frac{q^{Ns}}{s}ds+O\left(q^{\sigma N}\sum_{\deg f\leq M}\frac{|a(f)|}{|f|^{\sigma}}\left(\frac{1}{1+T}\right)\right),

upon combining (2.1) and (2.2). As fa(f)|f|s\sum_{f}\frac{a(f)}{|f|^{s}} converges absolutely for Re(s)=σ\mathrm{Re}(s)=\sigma, we obtain the desired result.

We end the current section by stating the orthogonality property of multiplicative characters. For more details, the reader may refer to [36, ch. 4].

Lemma 2.6.

For any two characters χ1\chi_{1} and χ2\chi_{2} modulo mm, we have

  1. (i)
    1ϕ(m)g(modm)χ1¯(g)χ2(g)={1,ifχ1=χ2;0,otherwise.\frac{1}{\phi(m)}\sum\limits_{g\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)}\overline{\chi_{1}}(g)\chi_{2}(g)=\begin{cases}1,\,\text{if}\,\chi_{1}=\chi_{2};\\ 0,\,\text{otherwise}.\end{cases}
  2. (ii)
    1ϕ(m)χ(modm)χ(f)χ(g)¯={1,iffg(modm);0,otherwise.\frac{1}{\phi(m)}\sum\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)}\chi(f)\overline{\chi(g)}=\begin{cases}1,\,\text{if}\,f\equiv g\mkern 4.0mu({\operator@font mod}\mkern 6.0mum);\\ 0,\,\text{otherwise}.\end{cases}

where χ¯\overline{\chi} is defined by χ1¯(h)=χ1(h)¯\overline{\chi_{1}}(h)=\overline{\chi_{1}(h)}, for all h𝔽q[t]h\in{\mathbb{F}}_{q}[t] and f,g𝔽q[t]f,g\in{\mathbb{F}}_{q}[t] are coprime to mm.

3. The Large sieve inequality over 𝔽q[t]{\mathbb{F}}_{q}[t]

The large sieve inequality has proved to be a versatile and powerful tool in number theory. The Bombieri-Vinogradov theorem can be considered as one of the finest applications of the large sieve method. The classical large sieve was first introduced by Linnik [29] around 1941 in the context of solving Vinogradov’s hypothesis related to the size of the least quadratic non-residue modulo a prime. It was further developed by contributions of Bombieri, Davenport, Halberstam, Gallagher, and many others. Analogous results on the large sieve inequality over number fields have been proved by Huxley ([25], [26]) and Hinz ([21], [22]). In 1971, Johnsen [27] established an analogue of the large sieve inequality for additive characters, and using similar techniques as introduced by Gallagher in [18], extended the theory to multiplicative characters as well. However, it appears that one requires to impose additional conditions on the set of moduli (see [27], p. 173). More generally, Hsu [23] proved a function field analogue of the large sieve inequality in arbitrary dimension. Recently Baier and Singh ([3], [5]) extended the large sieve inequality to square moduli and power moduli. In particular, [3] yields results on the large sieve inequality with additive characters in arbitrary dimension with a restricted set of moduli. For our purpose, we concentrate on the dimension one case (cf. [3], Corollary 6.5).

Continuing with the same notation and definitions as in [23] and [3], let 𝔽q(t){\mathbb{F}}_{q}(t) be the rational function field and 𝔽q(t){\mathbb{F}}_{q}(t)_{\infty} be the completion of 𝔽q(t){\mathbb{F}}_{q}(t) at the prime at infinity denoted by \infty. The absolute norm denoted by |.||.|_{\infty} is defined as

|i=naiti|=qn,\bigg{|}~{}\sum\limits_{i=-\infty}^{n}a_{i}t^{i}\bigg{|}_{\infty}=q^{n},

when 0an𝔽q0\neq a_{n}\in{\mathbb{F}}_{q}. Over 𝔽q[t]{\mathbb{F}}_{q}[t], this defines the usual norm of a polynomial. Let Tr:𝔽q𝔽p\mathrm{Tr}:\,{\mathbb{F}}_{q}\rightarrow{\mathbb{F}}_{p} be the usual trace map, where pp is the characteristic of the field 𝔽q{\mathbb{F}}_{q}. Consider the non-trivial group homomorphism E:𝔽qE:{\mathbb{F}}_{q}\rightarrow\mathbb{C}^{\ast} defined as

E(x)=exp(2πipTr(x)),\displaystyle E(x)=\exp\left(\frac{2\pi i}{p}\mathrm{Tr}(x)\right),

and define a map e:𝔽q(t)e:{\mathbb{F}}_{q}(t)_{\infty}\rightarrow\mathbb{C}^{\ast} as

e(i=naiti)=E(a1).\displaystyle e\left(\sum\limits_{i=-\infty}^{n}a_{i}t^{i}\right)=E(a_{-1}).

Using the additivity property of the trace function we see that, EE is a nontrivial additive character on 𝔽q{\mathbb{F}}_{q} and consequently ee becomes an additive character on 𝔽q[t]{\mathbb{F}}_{q}[t]. In particular, for some fixed f𝔽q[t]f\in{\mathbb{F}}_{q}[t], let σf:𝔽q[t]\sigma_{f}:\,{\mathbb{F}}_{q}[t]\rightarrow\mathbb{C}^{\ast}, be the additive character

σf(g)=e(gf).\displaystyle\sigma_{f}(g)=e\bigg{(}\frac{g}{f}\bigg{)}.

With the above notation in mind, we record the following result by Baier and Singh for dimension one. Considering the n=1n=1 case of Corollary 6.16.1 of [3], we are able to obtain a version of Corollary 6.56.5 of [3] with q+1q+1 replaced by qq as noted below. This is significant to us since we will keep qq fixed and NN\rightarrow\infty.

Theorem 3.1.

Let Q1Q\geq 1 be any natural number. Then

deg(f)Qfmonich(modf)(h,f)=1|deg(g)Nage(ghf)|2(qN+q2Q)deg(g)N|ag|2,\sum\limits_{\begin{subarray}{c}\deg(f)\leq Q\\ f\,\text{monic}\end{subarray}}\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\bigg{|}\sum\limits_{\deg(g)\leq N}a_{g}e\bigg{(}g\frac{h}{f}\bigg{)}\bigg{|}^{2}\ll\left(q^{N}+q^{2Q}\right)\sum\limits_{\deg(g)\leq N}|a_{g}|^{2}, (3.1)

where aga_{g}\in\mathbb{C} for all gg.

Lemma 3.2.

[27, Lemma 3] Let χ\chi be a primitive character modulo ff. Let

τ(χ¯)=h(modf)χ¯(h)e(hf).\tau(\overline{\chi})=\sum\limits_{h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\overline{\chi}(h)e\left(\frac{h}{f}\right). (3.2)

For any g𝔽q[t]g\in{\mathbb{F}}_{q}[t], we have

χ(g)τ(χ¯)=h(modf)(h,f)=1χ¯(h)e(ghf).\displaystyle\chi(g)\tau(\overline{\chi})=\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\overline{\chi}(h)e\left(\frac{gh}{f}\right). (3.3)
Proof.

Consider the case (g,f)=1(g,f)=1. Let g¯𝔽q[t]\overline{g}\in{\mathbb{F}}_{q}[t] satisfy gg¯1(modf)g\overline{g}\equiv 1\mkern 4.0mu({\operator@font mod}\mkern 6.0muf). Thus, in this case, we have

χ(g)τ(χ¯)=h(modf)χ¯(hg¯)e(hf)=m(modf)χ¯(m)e(mgf).\displaystyle\chi(g)\tau(\overline{\chi})=\sum\limits_{h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\overline{\chi}(h\overline{g})e\left(\frac{h}{f}\right)=\sum\limits_{m\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\overline{\chi}(m)e\left(\frac{mg}{f}\right).

We now consider (g,f)=d(g,f)=d; d>1d>1. Clearly the left-hand side of (3.3) is zero. Let f=f1df=f_{1}d and g=g1dg=g_{1}d. Then, applying the division algorithm on hh modulo f1f_{1} to write h=f1+ch=f_{1}\ell+c, we have that the right-hand side of (3.3) is

h(modf)(h,f)=1χ¯(h)e(ghf)=c(modf1)e(g1cf1)(modd)χ¯(f1+c).\displaystyle\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\overline{\chi}(h)e\left(\frac{gh}{f}\right)=\sum_{c\mkern 4.0mu({\operator@font mod}\mkern 6.0muf_{1})}e\left(\frac{g_{1}c}{f_{1}}\right)\sum_{\ell\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)}\overline{\chi}(f_{1}\ell+c).

Let S(c):=(modd)χ¯(f1+c)S(c):=\sum_{\ell\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)}\overline{\chi}(f_{1}\ell+c). Note that S(c+kf1)=S(c)S(c+kf_{1})=S(c), for any k𝔽q[t]k\in{\mathbb{F}}_{q}[t]. For a𝔽q[t]a\in{\mathbb{F}}_{q}[t] with (a,f)=1(a,f)=1 and a1(modf1)a\equiv 1\mkern 4.0mu({\operator@font mod}\mkern 6.0muf_{1}),

χ¯(a)S(c)\displaystyle\overline{\chi}(a)S(c) =χ¯(a)(modd)χ¯(f1+c)\displaystyle=\overline{\chi}(a)\sum_{\ell\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)}\overline{\chi}(f_{1}\ell+c)
=a(modd)χ¯(f1a+ac)\displaystyle=\sum_{a\ell\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)}\overline{\chi}(f_{1}a\ell+ac)
=(modd)χ¯(f1+c)=S(c).\displaystyle=\sum_{\ell^{\prime}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)}\overline{\chi}(f_{1}\ell^{\prime}+c)=S(c).

But, χ¯(a)1\overline{\chi}(a)\neq 1 for all such aa, as χ\chi is primitive. Thus, S(c)=0S(c)=0. This completes the proof. ∎

Lemma 3.3.

[27, Lemma 2] Let χ\chi be a primitive character modulo ff. Then |τ(χ)|2=|f||\tau(\chi)|^{2}=|f|, where τ\tau is defined as in Lemma 3.2.

Proof.

Let g𝔽q[t]g\in{\mathbb{F}}_{q}[t]. We have from (3.3)

|χ(g)τ(χ¯)|2\displaystyle\left|\chi(g)\tau(\overline{\chi})\right|^{2} =h(modf)(h,f)=1χ¯(h)e(ghf)h(modf)(h,f)=1χ(h)e(ghf)\displaystyle=\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\overline{\chi}(h)e\left(\frac{gh}{f}\right)\sum\limits_{\begin{subarray}{c}h^{\prime}\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h^{\prime},f)=1\end{subarray}}\chi(h^{\prime})e\left(-\frac{gh^{\prime}}{f}\right)
=h,h(modf)(hh,f)=1χ¯(h)χ(h)e(g(hh)f).\displaystyle=\sum_{\begin{subarray}{c}h,h^{\prime}\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (hh^{\prime},f)=1\end{subarray}}\overline{\chi}(h)\chi(h^{\prime})e\left(\frac{g(h-h^{\prime})}{f}\right).

Summing over gg, we have

|τ(χ¯)|2g(modf)|χ(g)|2\displaystyle\left|\tau(\overline{\chi})\right|^{2}\sum_{g\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\left|\chi(g)\right|^{2} =h,h(modf)(hh,f)=1χ¯(h)χ(h)g(modf)e(g(hh)f)\displaystyle=\sum_{\begin{subarray}{c}h,h^{\prime}\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (hh^{\prime},f)=1\end{subarray}}\overline{\chi}(h)\chi(h^{\prime})\sum_{g\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}e\left(\frac{g(h-h^{\prime})}{f}\right)
=|f|h(modf)|χ(h)|2.\displaystyle=|f|\sum_{h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}|\chi(h)|^{2}.

This proves our claim. ∎

Using Theorem 3.1, and Lemmas 3.2 and 3.3, we now obtain the following function-field analogue of the large sieve inequality with multiplicative characters, which we later make use of.

Theorem 3.4.

let (ag)(a_{g}) be a sequence of complex numbers and Q,NQ,N\in\mathbb{N}. Then

deg(f)Q,fmonic|f|ϕ(f)χ|deg(g)Nagχ(g)|2(qN+q2Q)deg(g)N|ag|2,\sum\limits_{\begin{subarray}{c}\deg(f)\leq Q,\\ f\,\text{monic}\end{subarray}}\frac{|f|}{\phi(f)}\sum\limits_{\chi}{\vphantom{\sum\limits_{\chi}}}^{\ast}\bigg{|}\sum\limits_{\deg(g)\leq N}a_{g}\chi(g)\bigg{|}^{2}\ll(q^{N}+q^{2Q})\sum\limits_{\deg(g)\leq N}|a_{g}|^{2}, (3.4)

where \sum{\vphantom{\sum\limits_{\chi}}}^{\ast} represents that the sum runs over primitive characters modulo ff.

Proof.

Let us write

S(χ)=deg(g)Nagχ(g).S(\chi)=\sum\limits_{\deg(g)\leq N}a_{g}\chi(g). (3.5)

Recall the definition (3.2) of τ\tau. We have

τ(χ¯)S(χ)\displaystyle\tau(\overline{\chi})S(\chi) =τ(χ¯)deg(g)N.agχ(g)\displaystyle=\tau(\overline{\chi})\sum\limits_{\deg(g)\leq N.}a_{g}\chi(g)
=deg(g)N.agχ(g)τ(χ¯)\displaystyle=\sum\limits_{\deg(g)\leq N.}a_{g}\chi(g)\tau(\overline{\chi})
=deg(g)Nagh(modf)(h,f)=1χ¯(h)e(ghf),\displaystyle=\sum\limits_{\deg(g)\leq N}a_{g}\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\overline{\chi}(h)e\bigg{(}g\frac{h}{f}\bigg{)},

using (3.3). Therefore,

τ(χ¯)S(χ)\displaystyle\tau(\overline{\chi})S(\chi) =h(modf)(h,f)=1χ¯(h)U(h/f),\displaystyle=\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\overline{\chi}(h)U(h/f),

where

U(x):=deg(g)Nage(gx).U(x):=\sum\limits_{\deg(g)\leq N}a_{g}e\big{(}gx\big{)}.

This gives

χ(modf)|τ(χ¯)S(χ)|2\displaystyle\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}|\tau(\overline{\chi})S(\chi)|^{2} =χ(modf)|h(modf)(h,f)=1χ¯(h)U(h/f)|2\displaystyle=\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\bigg{|}\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}\overline{\chi}(h)U(h/f)\bigg{|}^{2}
=h(modf)(h,f)=1(h,f)=1U(h/f)U(h/f)¯χ(modf)χ¯(h)χ(h).\displaystyle=\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\\ (h^{\prime},f)=1\end{subarray}}U(h/f)\overline{U(h^{\prime}/f)}\sum\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\overline{\chi}(h)\chi(h^{\prime}).

Using Lemma 3.3 for the left-hand side and the orthogonality property of multiplicative characters (Lemma 2.6) on the right-hand side, we have

|f|χ(modf)|S(χ)|2\displaystyle|f|\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}|S(\chi)|^{2} ϕ(f)h(modf)(h,f)=1|U(h/f)|2.\displaystyle\leq\phi(f)\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}|U(h/f)|^{2}.

Therefore,

|f|ϕ(f)χ(modf)|S(χ)|2\displaystyle\frac{|f|}{\phi(f)}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}|S(\chi)|^{2} h(modf)(h,f)=1|U(h/f)|2.\displaystyle\leq\sum\limits_{\begin{subarray}{c}h\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)\\ (h,f)=1\end{subarray}}|U(h/f)|^{2}.

Finally, summing over all monic polynomials ff having degree Q\leq Q and using Theorem 3.1, we obtain the result. ∎

Observe that from Lemma 4.2 of Klurman, Mangerel and Teräväinen [28], one can obtain the bound

|f|ϕ(f)χ(modf)|deg(g)Nagχ(g)|2(qN+2qdeg(f))deg(g)N|ag|2.\displaystyle\frac{|f|}{\phi(f)}\sum\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\bigg{|}\sum_{\deg(g)\leq N}a_{g}\chi(g)\bigg{|}^{2}\leq(q^{N}+2q^{\text{deg}(f)})\sum\limits_{\deg(g)\leq N}|a_{g}|^{2}. (3.6)

Summing this over monic polynomials ff of degree tt and then letting tt run from DD to QQ, we obtain

D<deg(f)Qfmonic|f|ϕ(f)χ(modf)|deg(g)Nagχ(g)|2(qN+Q+q2Q)deg(g)N|ag|2.\displaystyle\sum\limits_{\begin{subarray}{c}D<\deg(f)\leq Q\\ f\,\text{monic}\end{subarray}}\frac{|f|}{\phi(f)}\sum\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}\bigg{|}\sum_{\deg(g)\leq N}a_{g}\chi(g)\bigg{|}^{2}\ll(q^{N+Q}+q^{2Q})\sum\limits_{\deg(g)\leq N}|a_{g}|^{2}. (3.7)

Our bound (3.4) gives a savings of qQq^{Q} in the first main term by comparison. This is necessary to obtain the factor of qDq^{-D} in the modified form of the large sieve inequality stated in the next theorem. This plays a crucial role in the proof of Theorem 1.1.

Theorem 3.5.

let (ag)(a_{g}) be a sequence of complex numbers and D,Q,ND,Q,N\in\mathbb{N} . Then

D<deg(f)Qfmonic1ϕ(f)χ|deg(g)Nagχ(g)|2(qND+qQ)deg(g)N|ag|2,\sum\limits_{\begin{subarray}{c}D<\deg(f)\leq Q\\ f\,\text{monic}\end{subarray}}\frac{1}{\phi(f)}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi}\bigg{|}\sum_{\deg(g)\leq N}a_{g}\chi(g)\bigg{|}^{2}\ll(q^{N-D}+q^{Q})\sum\limits_{\deg(g)\leq N}|a_{g}|^{2}, (3.8)

where \sideset{}{{}^{\ast}}{\sum} represents that the sum runs over primitive characters modulo ff.

Proof.

Let S(χ)S(\chi) be as defined in (3.5). Let ψ:\psi:\mathbb{R}\rightarrow\mathbb{R} be the positive-valued decreasing continuous function defined as ψ(t)=qt\psi(t)={q^{-t}}. For any polynomial f𝔽q[t]f\in{\mathbb{F}}_{q}[t] and tt\in\mathbb{N}, we define

af:=|f|ϕ(f)χ(modf)|S(χ)|2,af(t)=deg(f)=t,fmonicaf and A(x):=txaf(t).a_{f}:=\frac{|f|}{\phi(f)}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}|S(\chi)|^{2},\qquad a_{f}(t)=\sum\limits_{\begin{subarray}{c}\deg(f)=t,\\ f\,\text{monic}\end{subarray}}a_{f}\qquad\text{ and }\qquad A(x):=\sum\limits_{\begin{subarray}{c}t\leq x\end{subarray}}a_{f}(t).

Using partial summation we have,

D<tQψ(t)af(t)=ψ(Q)A(Q)ψ(D)A(D)DQA(t)ψ(t)𝑑t.\sum\limits_{\begin{subarray}{c}D<t\leq Q\end{subarray}}\psi(t)a_{f}(t)=\psi(Q)A(Q)-\psi(D)A(D)-\int\limits_{D}^{Q}A(t)\psi^{\prime}(t)\,dt. (3.9)

Note that the left-hand side of (3.9) is

D<deg(f)Qfmonic1ϕ(f)χ(modf)|S(χ)|2.\sum\limits_{\begin{subarray}{c}D<\deg(f)\leq Q\\ f\,\text{monic}\end{subarray}}\frac{1}{\phi(f)}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0muf)}|S(\chi)|^{2}. (3.10)

From Theorem 3.4 we have,

A(x)(qN+q2x)Z,A(x)\ll\left(q^{N}+q^{2x}\right)Z,

where Z:=deg(g)N|ag|2Z:=\sum\limits_{\deg(g)\leq N}|a_{g}|^{2}. Therefore, the right-hand side of equation (3.9) is

(qND+qQ)Z.\displaystyle\ll\left(q^{N-D}+q^{Q}\right)Z. (3.11)

The result thus follows from (3.10) and (3.11).∎

4. Proof of Theorem 1.1

Let ψ1\psi_{1} and ψ2\psi_{2} satisfy properties (1), (2) and (3). Using Lemma 2.2, we have a constant kk such that ψ1(f)τk(f)\psi_{1}(f)\ll\tau_{k}(f) and ψ2(f)τk(f)\psi_{2}(f)\ll\tau_{k}(f). It is easy to derive property (1) for the convolution ψ1ψ2\psi_{1}\ast\psi_{2}. In the classical case, it is well-known that the Bombieri-Vinogradov type bound implies that of Siegel-Walfisz type. For the sake of completeness, we present the 𝔽q[t]{\mathbb{F}}_{q}[t] analogue of this fact. Let us first assume that ψ1ψ2\psi_{1}\ast\psi_{2} has property (3) and derive property (2) from that.

Let χ\chi be a non-principal character modulo dd having conductor dd^{\prime} of degree DlogN\leq D\log N, for sufficiently large DD. Let χ1\chi_{1} be a primitive character modulo dd^{\prime} such that χ(f)=χ0(f)χ1(f)\chi(f)=\chi_{0}(f)\chi_{1}(f) for all f𝔽q[t]f\in{\mathbb{F}}_{q}[t]. Write d=drd=d^{\prime}r. We see that

deg(f)=Nψ1ψ2(f)χ(f)\displaystyle\sum\limits_{\deg(f)=N}\psi_{1}\ast\psi_{2}(f)\chi(f) =deg(f)=Nψ1ψ2(f)χ1(f)χ0(f)\displaystyle=\sum_{\deg(f)=N}\psi_{1}\ast\psi_{2}(f)\chi_{1}(f)\chi_{0}(f)
=β(modd)(β,r)=1χ1(β)fβ(modd)deg(f)=Nψ1ψ2(f).\displaystyle=\sum_{\begin{subarray}{c}\beta\mkern 4.0mu({\operator@font mod}\mkern 6.0mud^{\prime})\\ (\beta,r)=1\end{subarray}}\chi_{1}(\beta)\sum_{\begin{subarray}{c}f\equiv\beta\mkern 4.0mu({\operator@font mod}\mkern 6.0mud^{\prime})\\ \deg(f)=N\end{subarray}}\psi_{1}\ast\psi_{2}(f).

Using property (3) with A=4DA=4D for the inner sum and fundamental property of μ\mu, we derive that the above sum is,

1ϕ(d)deg(f)=N(f,d)=1ψ1ψ2(f)β(modd)χ1(β)k|βk|rμ(k)+O(qN|d|N4D)\displaystyle\frac{1}{\phi(d^{\prime})}\sum_{\begin{subarray}{c}\deg(f)=N\\ (f,d^{\prime})=1\end{subarray}}\psi_{1}\ast\psi_{2}(f)\sum_{\begin{subarray}{c}\beta\mkern 4.0mu({\operator@font mod}\mkern 6.0mud^{\prime})\end{subarray}}\chi_{1}(\beta)\sum_{\begin{subarray}{c}k|\beta\\ k|r\end{subarray}}\mu(k)+O\left(\frac{q^{N}|d^{\prime}|}{N^{4D}}\right)
=1ϕ(d)deg(f)=N(f,d)=1ψ1ψ2(f)k|r(k,d)=1μ(k)kβ(modd)χ1(kβ)+O(qNN3D)\displaystyle=\frac{1}{\phi(d^{\prime})}\sum_{\begin{subarray}{c}\deg(f)=N\\ (f,d^{\prime})=1\end{subarray}}\psi_{1}\ast\psi_{2}(f)\sum_{\begin{subarray}{c}k|r\\ (k,d^{\prime})=1\end{subarray}}\mu(k)\sum_{\begin{subarray}{c}k\beta^{\prime}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud^{\prime})\end{subarray}}\chi_{1}(k\beta^{\prime})+O\left(\frac{q^{N}}{N^{3D}}\right)
qNN3D,\displaystyle\ll\frac{q^{N}}{N^{3D}},

as the sum kβ(modd)χ(kβ)\displaystyle\sum_{k\beta\mkern 4.0mu({\operator@font mod}\mkern 6.0mud^{\prime})}\chi(k\beta) vanishes due to orthogonality property of Dirichlet characters.

The remainder of this section is devoted to proving property (3). The crucial part here is judicious use of the large sieve inequality. Let mm-monic and l𝔽q[t]l\in{\mathbb{F}}_{q}[t] with (m,l)=1(m,l)=1.

From property (3) of ψ1\psi_{1} and ψ2\psi_{2}, for any fixed constant A>0A>0, we have a constant B=B(A)>0B=B(A)>0 such that

deg(m)QmaxMNmax(m,l)=1|E(M;m,l;ψ1)|qNNA,\displaystyle\sum\limits_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}|E(M;m,l;\psi_{1})|\ll\frac{q^{N}}{N^{A}}, (4.1)

and

deg(m)QmaxMNmax(m,l)=1|E(M;m,l;ψ2)|qNNA;\displaystyle\sum\limits_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}|E(M;m,l;\psi_{2})|\ll\frac{q^{N}}{N^{A}}; (4.2)

where Q=N2BlogNQ=\frac{N}{2}-B\log N.

As ψ1(f),ψ2(f)τk(f)\psi_{1}(f),\psi_{2}(f)\ll\tau_{k}(f), we have E(M;m,l;ψ1ψ2)qN/NBk+1E(M;m,l;\psi_{1}\ast\psi_{2})\ll q^{N}/N^{B-k+1} trivially for MN2M\leq\frac{N}{2}. Thus, we only need to deal with N2<MN\frac{N}{2}<M\leq N. Let N2<MN\frac{N}{2}<M\leq N. By the definition in (1.3), we have

E(M;m,l;ψ1ψ2)=ghl(modm)deg(gh)=Mψ1(g)ψ2(h)1ϕ(m)(gh,m)=1deg(gh)=Mψ1(g)ψ2(h).\displaystyle E(M;m,l;\psi_{1}\ast\psi_{2})=\sum\limits_{\begin{subarray}{c}gh\equiv l\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)\\ \text{deg}(gh)=M\end{subarray}}\psi_{1}(g)\psi_{2}(h)-\frac{1}{\phi(m)}\sum_{\begin{subarray}{c}(gh,m)=1\\ \text{deg}(gh)=M\end{subarray}}\psi_{1}(g)\psi_{2}(h). (4.3)

We divide the range of summation over deg(g)\text{deg}(g) in both sums in (4.3) into three sub-ranges as follows: (i) deg(g)AlogN\text{deg}(g)\leq A^{\prime}\log N, (ii) AlogN<deg(g)MBlogNA^{\prime}\log N<\text{deg}(g)\leq M-B^{\prime}\log N and (iii) MBlogN<deg(g)MM-B^{\prime}\log N<\text{deg}(g)\leq M; where A,B>0A^{\prime},B^{\prime}>0 are suitable constants to be chosen later. Since (gh,m)=1(gh,m)=1, the contribution to (4.3) from the range (i) is

deg(g)AlogN(g,m)=1ψ1(g)deg(h)Mdeg(g)hg1l(modm)ψ2(h)1ϕ(m)deg(g)AlogN(g,m)=1ψ1(g)deg(h)Mdeg(g)(h,m)=1ψ2(h),\displaystyle\sum_{\begin{subarray}{c}\text{deg}(g)\leq A^{\prime}\log N\\ (g,m)=1\end{subarray}}\psi_{1}(g)\sum_{\begin{subarray}{c}\text{deg}(h)\leq M-\text{deg}(g)\\ h\equiv g^{-1}l\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)\end{subarray}}\psi_{2}(h)-\frac{1}{\phi(m)}\sum_{\begin{subarray}{c}\text{deg}(g)\leq A^{\prime}\log N\\ (g,m)=1\end{subarray}}\psi_{1}(g)\sum_{\begin{subarray}{c}\text{deg}(h)\leq M-\text{deg}(g)\\ (h,m)=1\end{subarray}}\psi_{2}(h),

which can be re-written as

deg(g)AlogN(g,m)=1ψ1(g)E(Mdeg(g);m,g1l;ψ2).\displaystyle\sum_{\begin{subarray}{c}\text{deg}(g)\leq A^{\prime}\log N\\ (g,m)=1\end{subarray}}\psi_{1}(g)E(M-\text{deg}(g);m,g^{-1}l;\psi_{2}).

Expressing the two other sums arising from the ranges (ii) and (iii) in the same way, (4.3) can be written as

E(M;m,l;ψ1ψ2)\displaystyle E(M;m,l;\psi_{1}\ast\psi_{2}) =deg(g)AlogN(g,m)=1ψ1(g)E(Mdeg(g);m,g1l;ψ2)\displaystyle=\sum_{\begin{subarray}{c}\text{deg}(g)\leq A^{\prime}\log N\\ (g,m)=1\end{subarray}}\psi_{1}(g)E(M-\text{deg}(g);m,g^{-1}l;\psi_{2})
+AlogN<deg(g)MBlogN(g,m)=1ψ1(g)E(Mdeg(g);m,g1l;ψ2)\displaystyle+\sum_{\begin{subarray}{c}A^{\prime}\log N<\text{deg}(g)\leq M-B^{\prime}\log N\\ (g,m)=1\end{subarray}}\psi_{1}(g)E(M-\text{deg}(g);m,g^{-1}l;\psi_{2})
+deg(h)BlogN(h,m)=1ψ2(h)E(Mdeg(h);m,h1l;ψ1)\displaystyle+\sum_{\begin{subarray}{c}\text{deg}(h)\leq B^{\prime}\log N\\ (h,m)=1\end{subarray}}\psi_{2}(h)E(M-\text{deg}(h);m,h^{-1}l;\psi_{1})
:=S1+S2+S3(say).\displaystyle:=S_{1}+S_{2}+S_{3}\qquad\text{(say)}.

On the sum S1S_{1}, using (4.2) and Lemma 2.1(i), we obtain

deg(m)QmaxMNmax(m,l)=1S1\displaystyle\sum_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}S_{1} qNNAdeg(g)AlogNτk(g)\displaystyle\ll\frac{q^{N}}{N^{A}}\sum_{\text{deg}(g)\leq A^{\prime}\log N}\tau_{k}(g)
qNNA(AlogN)k1qAlogN.\displaystyle\ll\frac{q^{N}}{N^{A}}(A^{\prime}\log N)^{k-1}q^{A^{\prime}\log N}.

Now, considering the logarithm base qq, we conclude that

deg(m)QmaxMNmax(m,l)=1S1\displaystyle\sum_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}S_{1} qNNAAϵ,\displaystyle\ll\frac{q^{N}}{N^{A-A^{\prime}-\epsilon}}, (4.4)

for any ϵ>0\epsilon>0. Similarly, using (4.1) and Lemma 2.1(i) we obtain for any ϵ>0\epsilon>0,

deg(m)QmaxMNmax(m,l)=1S3\displaystyle\sum_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}S_{3} qNNAAϵ.\displaystyle\ll\frac{q^{N}}{N^{A-A^{\prime}-\epsilon}}. (4.5)

Next, we turn our attention to

S4:=deg(m)QmaxMNmax(m,l)=1S2.\displaystyle S_{4}:=\sum_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}S_{2}.

Choosing D1=A2logND_{1}=\frac{A^{\prime}}{2}\log N, we split the range of deg(m)\deg(m) and write S4S_{4} into two subsums

S4=deg(m)D1maxMNmax(m,l)=1S2+D1<deg(m)QmaxMNmax(m,l)=1S2=:T1+T2.S_{4}=\sum_{\deg(m)\leq D_{1}}\max\limits_{M\leq N}\max\limits_{(m,l)=1}S_{2}+\sum_{D_{1}<\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}S_{2}\\ =:T_{1}+T_{2}. (4.6)

Now onwards, we use the notation A1:=AlogNA_{1}:=A^{\prime}\log N and A2:=MBlogNA_{2}:=M-B^{\prime}\log N. Using Lemma 2.6, we express S2S_{2} as

S2\displaystyle S_{2} =1ϕ(m)χ(modm)χχ0χ¯(l)A1<deg(g)A2ψ1(g)χ(g)deg(h)Mdeg(g)ψ2(h)χ(h),\displaystyle=\frac{1}{\phi(m)}\sum_{\begin{subarray}{c}\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)\\ \chi\neq\chi_{0}\end{subarray}}\overline{\chi}(l)\sum_{\begin{subarray}{c}A_{1}<\text{deg}(g)\leq A_{2}\end{subarray}}\psi_{1}(g)\chi(g)\sum_{\begin{subarray}{c}\text{deg}(h)\leq M-\text{deg}(g)\end{subarray}}\psi_{2}(h)\chi(h), (4.7)

where χ0\chi_{0} denotes the principal character modulo mm. Putting (4.7) in T1T_{1}, we see that

T1deg(m)D11ϕ(m)maxMNχ(modm)χχ0|A1<deg(g)A2ψ1(g)χ(g)deg(h)Mdeg(g)ψ2(h)χ(h)|.\displaystyle T_{1}\leq\sum_{\deg(m)\leq D_{1}}\frac{1}{\phi(m)}\max\limits_{M\leq N}\sum_{\begin{subarray}{c}\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum)\\ \chi\neq\chi_{0}\end{subarray}}\Bigg{|}~{}\sum_{\begin{subarray}{c}A_{1}<\text{deg}(g)\leq A_{2}\end{subarray}}\psi_{1}(g)\chi(g)\sum_{\begin{subarray}{c}\text{deg}(h)\leq M-\text{deg}(g)\end{subarray}}\psi_{2}(h)\chi(h)\Bigg{|}.

Now, for some δ(0,1)\delta\in(0,1), we can write the inner sum over gg and hh in (4.7), say T1(χ)T_{1}^{\prime}(\chi), as

T1(χ)\displaystyle T_{1}^{\prime}(\chi) =A1<deg(g)(1δ)Mψ1(g)χ(g)deg(h)Mdeg(g)ψ2(h)χ(h)\displaystyle=\sum_{A_{1}<\text{deg}(g)\leq(1-\delta)M}\psi_{1}(g)\chi(g)\sum_{\text{deg}(h)\leq M-\text{deg}(g)}\psi_{2}(h)\chi(h)
+deg(h)δMψ2(h)χ(h)(1δ)M<deg(g)min(Mdeg(h),A2)ψ1(g)χ(g).\displaystyle\qquad\qquad\qquad\qquad+\sum_{\text{deg}(h)\leq\delta M}\psi_{2}(h)\chi(h)\sum_{(1-\delta)M<\text{deg}(g)\leq\min{(M-\text{deg}(h),A_{2})}}\psi_{1}(g)\chi(g).

Using property (1) and (2) for ψ1\psi_{1} and ψ2\psi_{2} and the bound from Lemma 2.1(ii), we have

T1(χ)\displaystyle T_{1}^{\prime}(\chi) qMM3A/2(A1<deg(g)(1δ)Mτk(g)qdeg(g)+deg(h)δMτk(h)qdeg(h))qMM3A/2k.\displaystyle\ll\frac{q^{M}}{M^{3A^{\prime}/2}}\left(\sum_{A_{1}<\text{deg}(g)\leq(1-\delta)M}\frac{\tau_{k}(g)}{q^{\text{deg}(g)}}+\sum_{\text{deg}(h)\leq\delta M}\frac{\tau_{k}(h)}{q^{\text{deg}(h)}}\right)\ll\frac{q^{M}}{M^{3A^{\prime}/2-k}}. (4.8)

Thus,

T1qNN3A/2kdeg(m)D11qNNAk.\displaystyle T_{1}\ll\frac{q^{N}}{N^{3A^{\prime}/2-k}}\sum_{\deg(m)\leq D_{1}}1\ll\frac{q^{N}}{N^{A^{\prime}-k}}. (4.9)

Now, to estimate T2T_{2}, we first write the sum (4.7) in terms of primitive characters and separate the terms coming from characters having small conductor. Note that, any character χ\chi modulo mm having conductor m1m_{1} can be expressed as χ(g)=χ1(g)χ0(g)\chi(g)=\chi_{1}(g)\chi_{0}(g) for all g𝔽q[t]g\in{\mathbb{F}}_{q}[t], where χ0\chi_{0} is the principal character modulo mm and χ1\chi_{1} is a primitive character modulo m1m_{1}. Let m=m1m2m=m_{1}m_{2}. Using the inequality ϕ(m1m2)ϕ(m1)ϕ(m2)\phi(m_{1}m_{2})\geq\phi(m_{1})\phi(m_{2}), we obtain

T2\displaystyle T_{2}\leq deg(m2)QmaxMNmax(m2,l)=11ϕ(m2)(deg(m1)D1max(m1,l)=11ϕ(m1)|χ(modm1)χ¯(l)T1(χ)|\displaystyle\sum\limits_{\deg(m_{2})\leq Q}\max\limits_{M\leq N}\max\limits_{(m_{2},l)=1}\frac{1}{\phi(m_{2})}\left(\sum_{\deg(m_{1})\leq D_{1}}\max\limits_{(m_{1},l)=1}\frac{1}{\phi(m_{1})}\Bigg{|}~{}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)T_{1}^{\prime}(\chi)\Bigg{|}\right.
+D1<deg(m1)Qmax(m1,l)=11ϕ(m1)|χ(modm1)χ¯(l)T1(χ)|)\displaystyle\left.+\sum_{D_{1}<\deg(m_{1})\leq Q}\max\limits_{(m_{1},l)=1}\frac{1}{\phi(m_{1})}\Bigg{|}~{}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)T_{1}^{\prime}(\chi)\Bigg{|}\right)
=:U1+U2(say),\displaystyle=:U_{1}+U_{2}\qquad\text{(say)}, (4.10)

where \sum{\vphantom{\sum\limits_{\chi}}}^{\ast} denotes the sum is over primitive characters. The first sum U1U_{1} can be bounded using the above upper bound for T1(χ)T_{1}^{\prime}(\chi) in (4.8) and Lemma 2.3 as follows.

U1\displaystyle U_{1} deg(m2)QmaxMN1ϕ(m2)deg(m1)D1qMM3A/2k\displaystyle\ll\sum_{\deg(m_{2})\leq Q}\max\limits_{M\leq N}\frac{1}{\phi(m_{2})}\sum_{\deg(m_{1})\leq D_{1}}\frac{q^{M}}{M^{3A^{\prime}/2-k}}
qNNAk1.\displaystyle\ll\frac{q^{N}}{N^{A^{\prime}-k-1}}. (4.11)

Next, we define ψ~1(f)=ψ1(f)\widetilde{\psi}_{1}(f)=\psi_{1}(f) and ψ~2(f)=ψ2(f)\widetilde{\psi}_{2}(f)=\psi_{2}(f) if (f,m2)=1(f,m_{2})=1, and 0 else. Thus,

U2=\displaystyle U_{2}= deg(m2)QmaxMNmax(m2,l)=11ϕ(m2)D1<deg(m1)Qmax(m1,l)=11ϕ(m1)\displaystyle\sum\limits_{\deg(m_{2})\leq Q}\max\limits_{M\leq N}\max\limits_{(m_{2},l)=1}\frac{1}{\phi(m_{2})}\sum\limits_{D_{1}<\deg(m_{1})\leq Q}\max\limits_{(m_{1},l)=1}\frac{1}{\phi(m_{1})}
|χ(modm1)χ¯(l)A1<deg(g)A2ψ~1(g)χ(g)deg(h)Mdeg(g)ψ~2(h)χ(h)|.\displaystyle\qquad\qquad\qquad\Bigg{|}~{}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)\sum\limits_{A_{1}<\text{deg}(g)\leq A_{2}}\widetilde{\psi}_{1}(g)\chi(g)\sum\limits_{\text{deg}(h)\leq M-\text{deg}(g)}\widetilde{\psi}_{2}(h)\chi(h)\Bigg{|}.

Now, let M0=M+12M_{0}=M+\frac{1}{2}, σ=1+12N\sigma=1+\frac{1}{2N} and T>MT>M to be specified later. Note that τk(f)|f|ϵ\tau_{k}(f)\ll|f|^{\epsilon}, for any ϵ>0\epsilon>0. Thus, applying the function field analogue of Perron’s formula given in Lemma 2.5 on U2U_{2}, using Lemma 2.3 to simplify the error, we obtain

U2\displaystyle U_{2}\leq deg(m2)QmaxMNmax(m2,l)=11ϕ(m2)D1<deg(m1)Qmax(m1,l)=11ϕ(m1)\displaystyle\sum\limits_{\deg(m_{2})\leq Q}\max\limits_{M\leq N}\max\limits_{(m_{2},l)=1}\frac{1}{\phi(m_{2})}\sum\limits_{D_{1}<\deg(m_{1})\leq Q}\max\limits_{(m_{1},l)=1}\frac{1}{\phi(m_{1})}
|χ(modm1)χ¯(l)σiqTσ+iqTΦ~1(χ,s)Φ~2(χ,s)qM0ss𝑑s|+O(q3N2qTNB1),\displaystyle\qquad\qquad\Bigg{|}~{}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)\int_{\sigma-iq^{T}}^{\sigma+iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2}(\chi,s)\frac{q^{M_{0}s}}{s}ds\Bigg{|}+O\left(\frac{q^{\frac{3N}{2}}}{q^{T}N^{B-1}}\right), (4.12)

where

Φ~1(χ,s):=A1<deg(g)A2ψ~1(g)χ(g)|g|s and Φ~2(χ,s):=deg(h)Tψ~2(h)χ(h)|h|s.\widetilde{\Phi}_{1}(\chi,s):=\sum_{A_{1}<\text{deg}(g)\leq A_{2}}\frac{\widetilde{\psi}_{1}(g)\chi(g)}{|g|^{s}}\qquad\text{ and }\qquad\widetilde{\Phi}_{2}(\chi,s):=\sum_{\text{deg}(h)\leq T}\frac{\widetilde{\psi}_{2}(h)\chi(h)}{|h|^{s}}.

We define

IM,m2(l)\displaystyle I_{M,m_{2}}(l) :=D1<deg(m1)Qmax(m1,l)=11ϕ(m1)|χ(modm1)χ¯(l)σiqTσ+iqTΦ~1(χ,s)Φ~2(χ,s)qM0ss𝑑s|.\displaystyle:=\sum\limits_{D_{1}<\deg(m_{1})\leq Q}\max\limits_{(m_{1},l)=1}\frac{1}{\phi(m_{1})}\Bigg{|}~{}\sideset{}{{}^{\ast}}{\sum}\limits_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)\int_{\sigma-iq^{T}}^{\sigma+iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2}(\chi,s)\frac{q^{M_{0}s}}{s}ds\Bigg{|}.

Now, we divide the sum over deg(m1)\deg(m_{1}) in the form

IM,m2(l)=j=0JIM,m2(j)(l),\displaystyle I_{M,m_{2}}(l)=\sum\limits_{j=0}^{J}I_{M,m_{2}}^{(j)}(l),

where J+D1<QJ+1+D1J+D_{1}<Q\leq J+1+D_{1} and

IM,m2(j)(l)=deg(m1)Rjmax(m1,l)=11ϕ(m1)|χ(modm1)χ¯(l)σiqTσ+iqTΦ~1(χ,s)Φ~2(χ,s)qM0ss𝑑s|.\displaystyle I_{M,m_{2}}^{(j)}(l)=\sum\limits_{\deg(m_{1})\in R_{j}}\max\limits_{(m_{1},l)=1}\frac{1}{\phi(m_{1})}\Bigg{|}~{}\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)\int_{\sigma-iq^{T}}^{\sigma+iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2}(\chi,s)\frac{q^{M_{0}s}}{s}ds\Bigg{|}.

Here RjR_{j} denotes the range D1+j<deg(m1)j+1+D1D_{1}+j<\deg(m_{1})\leq j+1+D_{1} for each j=0,..,(J1)j=0,..,(J-1), and J+D1<deg(m1)QJ+D_{1}<\deg(m_{1})\leq Q at j=Jj=J. Also, we divide the integral IM,m2I_{M,m_{2}} into two sub parts by writing Φ~2(χ,s)\widetilde{\Phi}_{2}(\chi,s) as

Φ~2(χ,s)\displaystyle\widetilde{\Phi}_{2}{(\chi,s)} =hYjψ2(h)χ(h)|h|s+Yj<hTψ2(h)χ(h)|h|s\displaystyle=\sum\limits_{h\leq Y_{j}}\frac{\psi_{2}(h)\chi(h)}{|h|^{s}}+\sum\limits_{Y_{j}<h\leq T}\frac{\psi_{2}(h)\chi(h)}{|h|^{s}}
:=Φ~2j(1)(χ,s)+Φ~2j(2)(χ,s),\displaystyle:=\widetilde{\Phi}_{2j}^{(1)}{(\chi,s)}+\widetilde{\Phi}_{2j}^{(2)}{(\chi,s)},

where Yj<TY_{j}<T will be chosen later. Now, for each jj, we have

σiqTσ+iqTΦ~1(χ,s)Φ~2j(2)(χ,s)qM0ss𝑑s1/2iqT1/2+iqTΦ~1(χ,s)Φ~2j(2)(χ,s)qM0ss𝑑s\displaystyle\int_{\sigma-iq^{T}}^{\sigma+iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2j}^{(2)}(\chi,s)\frac{q^{M_{0}s}}{s}ds-\int_{1/2-iq^{T}}^{1/2+iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2j}^{(2)}(\chi,s)\frac{q^{M_{0}s}}{s}ds
=σiqT1/2iqTΦ~1(χ,s)Φ~2j(2)(χ,s)qM0ss𝑑s+1/2+iqTσ+iqTΦ~1(χ,s)Φ~2j(2)(χ,s)qM0ss𝑑s\displaystyle=\int_{\sigma-iq^{T}}^{1/2-iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2j}^{(2)}(\chi,s)\frac{q^{M_{0}s}}{s}ds+\int_{1/2+iq^{T}}^{\sigma+iq^{T}}\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2j}^{(2)}(\chi,s)\frac{q^{M_{0}s}}{s}ds
qM0qT(A1<γA2qγ/2degg=γτk(g))(γYjqγ/2degh=γτk(h))\displaystyle\ll\frac{q^{M_{0}}}{q^{T}}\left(\sum_{A_{1}<\gamma\leq A_{2}}q^{-\gamma/2}\sum\limits_{\text{deg}g=\gamma}\tau_{k}(g)\right)\left(\sum_{\gamma\leq Y_{j}}q^{-\gamma/2}\sum_{\text{deg}h=\gamma}\tau_{k}(h)\right)
q3N/2qYj/2(Yj)k(A2)kqTNB/2,\displaystyle\ll\frac{q^{3N/2}q^{Y_{j}/2}(Y_{j})^{k}(A_{2})^{k}}{q^{T}N^{B^{\prime}/2}},

using Lemma 2.1(i) and keeping in mind that A2=MBlogNNA_{2}=M-B^{\prime}\log N\leq N. Therefore, for each jj we can write

IM,m2(j)(l)=1/2iqT1/2+iqTSj(1)(s)qM0ss𝑑s+σiqTσ+iqTSj(2)(s)qM0ss𝑑s+O(q3N/2qYj/2(Yj)k(A2)kqTNB/2),\displaystyle I_{M,m_{2}}^{(j)}(l)=\int_{1/2-iq^{T}}^{1/2+iq^{T}}S_{j}^{(1)}(s)\frac{q^{M_{0}s}}{s}ds+\int_{\sigma-iq^{T}}^{\sigma+iq^{T}}S_{j}^{(2)}(s)\frac{q^{M_{0}s}}{s}ds+O\left(\frac{q^{3N/2}q^{Y_{j}/2}(Y_{j})^{k}(A_{2})^{k}}{q^{T}N^{B^{\prime}/2}}\right), (4.13)

where

Sj(α):=deg(m1)Rjmax(m1,l)=1cϕ(m1)χ(modm1)χ¯(l)Φ~1(χ,s)Φ~2j(k)(χ,s),\displaystyle S_{j}^{(\alpha)}:=\sum_{\deg(m_{1})\in R_{j}}\max\limits_{(m_{1},l)=1}\frac{c}{\phi(m_{1})}\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi}(l)\widetilde{\Phi}_{1}(\chi,s)\widetilde{\Phi}_{2j}^{(k)}(\chi,s), (4.14)

for α=1,2\alpha=1,2 and some complex number cc with |c|=1|c|=1. Now, using the Cauchy-Schwarz inequality first for the sum over the primitive characters χ\chi and then for the sum over m1m_{1}, we obtain

Sj(1)(s)\displaystyle S_{j}^{(1)}(s)\leq (deg(m1)Rj1ϕ(m1)χ(modm1)|Φ~1(χ,s)|2)1/2\displaystyle\left(\sum_{\deg(m_{1})\in R_{j}}\frac{1}{\phi(m_{1})}\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}|\widetilde{\Phi}_{1}(\chi,s)|^{2}\right)^{1/2}
×(deg(m1)Rj1ϕ(m1)χ(modm1)|Φ~2j(1)(χ,s)|2)1/2.\displaystyle\qquad\qquad\qquad\qquad\times\left(\sum_{\deg(m_{1})\in R_{j}}\frac{1}{\phi(m_{1})}\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}|\widetilde{\Phi}_{2j}^{(1)}(\chi,s)|^{2}\right)^{1/2}.

Next, applying the large sieve inequality as in Theorem 3.5 for s=1/2+its=1/2+it such that qTtqT-q^{T}\leq t\leq q^{T}, and Lemma 2.1(ii), we have for j<Jj<J,

Sj(1)(s)\displaystyle S_{j}^{(1)}(s) ((qD1+j+1+qA2D1j)deg(g)A2τk(g)|g|)1/2((qD1+j+1+qYjD1j)deg(h)Yjτk(h)|h|)1/2\displaystyle\ll\left(\left(q^{D_{1}+j+1}+q^{A_{2}-D_{1}-j}\right)\sum_{\deg(g)\leq A_{2}}\frac{\tau_{k}(g)}{|g|}\right)^{1/2}\left(\left(q^{D_{1}+j+1}+q^{Y_{j}-D_{1}-j}\right)\sum_{\deg(h)\leq Y_{j}}\frac{\tau_{k}(h)}{|h|}\right)^{1/2}
(q2Q+qYj+1+qA2+1+qA2+Yj2D12j)1/2A2k/2Yjk/2.\displaystyle\ll\left(q^{2Q}+q^{Y_{j}+1}+q^{A_{2}+1}+q^{A_{2}+Y_{j}-2D_{1}-2j}\right)^{1/2}A_{2}^{k/2}Y_{j}^{k/2}.

Similarly, for j=Jj=J also, we can get the same upper bound. Choose Yj=2D1+2jY_{j}=2D_{1}+2j. Thus,

Sj(1)(1/2+it)qN/2N12min{2B,B}k.\displaystyle S_{j}^{(1)}(1/2+it)\ll\frac{q^{N/2}}{N^{\frac{1}{2}\min\{2B,B^{\prime}\}-k}}. (4.15)

Now, we turn to Sj(2)(s)S_{j}^{(2)}(s) at s=σ+its=\sigma+it, such that qTtqT-q^{T}\leq t\leq q^{T}. We divide the sums Φ~1(χ,s)\widetilde{\Phi}_{1}(\chi,s) and Φ~2j(2)(χ,s)\widetilde{\Phi}_{2j}^{(2)}(\chi,s) as follows. Define for A1+I<A2A1+I+1A_{1}+I<A_{2}\leq A_{1}+I+1,

Φ~1(χ,s)=i=0IΦ~1i(χ,s),\displaystyle\widetilde{\Phi}_{1}(\chi,s)=\sum_{i=0}^{I}\widetilde{\Phi}_{1i}(\chi,s),

where Φ~1i(χ,s):=A1+i<deg(g)A1+i+1ψ~1(g)χ(g)|g|s\displaystyle\widetilde{\Phi}_{1i}(\chi,s):=\sum_{A_{1}+i<\deg(g)\leq A_{1}+i+1}\frac{\widetilde{\psi}_{1}(g)\chi(g)}{|g|^{s}}; and for Yj+R<TYj+R+1Y_{j}+R<T\leq Y_{j}+R+1,

Φ~2j(2)(χ,s)=r=0RΦ~2jr(2)(χ,s),\displaystyle\widetilde{\Phi}_{2j}^{(2)}(\chi,s)=\sum_{r=0}^{R}\widetilde{\Phi}_{2jr}^{(2)}(\chi,s),

where Φ~2jr(2)(χ,s):=Yj+r<deg(h)Yj+r+1ψ~2(g)χ(g)|g|s\displaystyle\widetilde{\Phi}_{2jr}^{(2)}(\chi,s):=\sum_{Y_{j}+r<\deg(h)\leq Y_{j}+r+1}\frac{\widetilde{\psi}_{2}(g)\chi(g)}{|g|^{s}}.
Thus, we can write Sj(2)(s)S_{j}^{(2)}(s) as

Sj(2)(s)=i=0Ir=0Rdeg(m1)Rjcϕ(m1)χ(modm1)χ1¯(l)Φ~1i(χ,s)Φ~2jr(2)(χ,s).\displaystyle S_{j}^{(2)}(s)=\sum_{i=0}^{I}\sum_{r=0}^{R}\sum_{\deg(m_{1})\in R_{j}}\frac{c}{\phi(m_{1})}\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\overline{\chi_{1}}(l)\widetilde{\Phi}_{1i}(\chi,s)\widetilde{\Phi}_{2jr}^{(2)}(\chi,s).

Applying the Cauchy-Schwarz inequality twice as in Sj(1)(s)S_{j}^{(1)}(s), we obtain that Sj(2)(s)S_{j}^{(2)}(s) is bounded above by

i=0Ir=0R(deg(m1)Rj1ϕ(m1)χ(modm1)|Φ~1i(χ,s)|)1/2(deg(m1)Rj1ϕ(m1)χ(modm1)|Φ~2jr(2)(χ,s)|)1/2.\displaystyle\sum_{i=0}^{I}\sum_{r=0}^{R}\left(\!\sum_{\deg(m_{1})\in R_{j}}\!\frac{1}{\phi(m_{1})}\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\left|\widetilde{\Phi}_{1i}(\chi,s)\right|\right)^{1/2}\left(\!\sum_{\deg(m_{1})\in R_{j}}\frac{1}{\phi(m_{1})}\!\sideset{}{{}^{\ast}}{\sum}_{\chi\mkern 4.0mu({\operator@font mod}\mkern 6.0mum_{1})}\left|\widetilde{\Phi}_{2jr}^{(2)}(\chi,s)\right|\right)^{1/2}\!.

Again applying the large sieve inequality from Theorem 3.5 for s=σ+its=\sigma+it and Lemma 2.1(ii), we have for j<Jj<J and for some k>0k^{\prime}>0,

Sj(2)(s)i=0Ir=0R\displaystyle S_{j}^{(2)}(s)\ll\sum_{i=0}^{I}\sum_{r=0}^{R} ((q(A1+i+1)(D1+j)+qD1+j+1)A1+i<deg(g)A1+i+1τk(g)|g|2)1/2\displaystyle\left(\left(q^{(A_{1}+i+1)-(D_{1}+j)}+q^{D_{1}+j+1}\right)\sum_{A_{1}+i<\deg(g)\leq A_{1}+i+1}\frac{\tau_{k^{\prime}}(g)}{|g|^{2}}\right)^{1/2}
((q(Yj+r+1)(D1+j)+qD1+j+1)Yj+r<deg(h)Yj+r+1τk(h)|g|2)1/2\displaystyle\left(\left(q^{(Y_{j}+r+1)-(D_{1}+j)}+q^{D_{1}+j+1}\right)\sum_{Y_{j}+r<\deg(h)\leq Y_{j}+r+1}\frac{\tau_{k^{\prime}}(h)}{|g|^{2}}\right)^{1/2}
i=0Ir=0R\displaystyle\ll\sum_{i=0}^{I}\sum_{r=0}^{R} ((q(A1+i+1)(D1+j)+qD1+j+1)(A1+i+1)kqA1+i)1/2\displaystyle\left(\left(q^{(A_{1}+i+1)-(D_{1}+j)}+q^{D_{1}+j+1}\right)\frac{(A_{1}+i+1)^{k^{\prime}}}{q^{A_{1}+i}}\right)^{1/2}
((q(Yj+r+1)(D1+j)+qD1+j+1)(Yj+r+1)kqYj+r)1/2\displaystyle\left(\left(q^{(Y_{j}+r+1)-(D_{1}+j)}+q^{D_{1}+j+1}\right)\frac{(Y_{j}+r+1)^{k^{\prime}}}{q^{Y_{j}+r}}\right)^{1/2}
i=0Ir=0R\displaystyle\ll\sum_{i=0}^{I}\sum_{r=0}^{R} (qD1j+qjD1i)1/2(qD1j+qD1jr)1/2(A2)k/2(T)k/2,\displaystyle\left(q^{-D_{1}-j}+q^{j-D_{1}-i}\right)^{1/2}\left(q^{-D_{1}-j}+q^{-D_{1}-j-r}\right)^{1/2}(A_{2})^{k^{\prime}/2}(T)^{k^{\prime}/2}, (4.16)

because of our choices A1=2D1A_{1}=2D_{1} and Yj=2D1+2jY_{j}=2D_{1}+2j. We choose T=2NT=2N. Since the lengths of the sums over ii and rr are O(A2)O(A_{2}) and O(T)O(T) respectively, and the summand is maximum at i=0i=0 and r=0r=0, from (4.16), we obtain

Sj(2)(s)qD1Nk+2=Nk+2A/2,S_{j}^{(2)}(s)\ll q^{-D_{1}}N^{k^{\prime}+2}=N^{k^{\prime}+2-A^{\prime}/2}, (4.17)

putting the value D1=A2logND_{1}=\frac{A^{\prime}}{2}\log N. Therefore, using (4.15) and (4.17) in (4.13), for each jj we have

IM,m2(j)(l)qNN12min{2B,B}k+qNNA/2k2,\displaystyle I_{M,m_{2}}^{(j)}(l)\ll\frac{q^{N}}{N^{\frac{1}{2}\min\{2B,B^{\prime}\}-k}}+\frac{q^{N}}{N^{A^{\prime}/2-k^{\prime}-2}},

and hence

IM,m2(l)qNN12min{2B,B}k1+qNNA/2k3,\displaystyle I_{M,m_{2}}(l)\ll\frac{q^{N}}{N^{\frac{1}{2}\min\{2B,B^{\prime}\}-k-1}}+\frac{q^{N}}{N^{A^{\prime}/2-k^{\prime}-3}}, (4.18)

since the length of the sum over jj is QN\ll Q\ll N. Using Lemma 2.3, and putting (4.18) in (4.12), we obtain

U2qNN12min{2B,B}k2+qNNA/2k4.\displaystyle U_{2}\ll\frac{q^{N}}{N^{\frac{1}{2}\min\{2B,B^{\prime}\}-k-2}}+\frac{q^{N}}{N^{A^{\prime}/2-k^{\prime}-4}}. (4.19)

Therefore, using (4.11) and (4.19) in (4.10), we have

T2qNNAk1+qNN12min{2B,B}k2+qNNA/2k4.T_{2}\ll\frac{q^{N}}{N^{A^{\prime}-k-1}}+\frac{q^{N}}{N^{\frac{1}{2}\min\{2B,B^{\prime}\}-k-2}}+\frac{q^{N}}{N^{A^{\prime}/2-k^{\prime}-4}}. (4.20)

Combining (4.9) and (4.20) in (4.6), we get

S4qNNAk+qNNAk1+qNN12min{2B,B}k2+qNNA/2k4.S_{4}\ll\frac{q^{N}}{N^{A^{\prime}-k}}+\frac{q^{N}}{N^{A^{\prime}-k-1}}+\frac{q^{N}}{N^{\frac{1}{2}\min\{2B,B^{\prime}\}-k-2}}+\frac{q^{N}}{N^{A^{\prime}/2-k^{\prime}-4}}. (4.21)

Suppose k′′=max{k,k}k^{\prime\prime}=\max\left\{k,k^{\prime}\right\} and we choose A=B=A/2A^{\prime}=B^{\prime}=A/2. Combining all the bounds (4.4), (4.5) and (4.21), we have for A>8k′′+32A>8k^{\prime\prime}+32,

deg(m)QmaxMNmax(m,l)=1|E(M;m,l;ψ1ψ2)|qNNA/8.\displaystyle\sum\limits_{\deg(m)\leq Q}\max\limits_{M\leq N}\max\limits_{(m,l)=1}|E(M;m,l;\psi_{1}\ast\psi_{2})|\ll\frac{q^{N}}{N^{A/8}}.

Choosing B=B(8A)B=B(8A), we get the required form. Thus we have derived property (3) for ψ1ψ2\psi_{1}\ast\psi_{2}.

5. Proof of Corollary 1.4

Consider ψ1\psi\equiv 1. Fix some m,𝔽q[t]m,\ell\in{\mathbb{F}}_{q}[t] with (m,)=1(m,\ell)=1, deg()<deg(m)\deg(\ell)<\deg(m). We have

f(modm)deg(f)=Mψ(f)=f=md+deg(f)=M1=deg(d)=Mdeg(m)1=qMdeg(m),\displaystyle\sum_{\begin{subarray}{c}f\equiv\ell\pmod{m}\\ \deg(f)=M\end{subarray}}\psi(f)=\sum_{\begin{subarray}{c}f=md+\ell\\ \deg(f)=M\end{subarray}}1=\sum_{\deg(d)=M-\deg(m)}1=q^{M-\deg(m)}, (5.1)

and

(f,m)=1deg(f)=Mψ(f)\displaystyle\sum_{\begin{subarray}{c}(f,m)=1\\ \deg(f)=M\end{subarray}}\psi(f) =deg(f)=Md|f,d|mμ(d)\displaystyle=\sum_{\deg(f)=M}\sum_{d|f,d|m}\mu(d)
=d|mμ(d)d|fdeg(f)=M1\displaystyle=\sum_{d|m}\mu(d)\sum_{\begin{subarray}{c}d|f\\ \deg(f)=M\end{subarray}}1
=d|mμ(d)qMdeg(d)=qMP|m(11|P|)=ϕ(m)qMdeg(m)\displaystyle=\sum_{d|m}\mu(d)q^{M-\deg(d)}=q^{M}\prod_{P|m}\left(1-\frac{1}{|P|}\right)=\phi(m)q^{M-\deg(m)} (5.2)

from (5.1) and (5.2), we have that ψ\psi has property (3), and (1), (2) trivially follows. Since τk=ψψψ\tau_{k}=\psi\ast\psi\ast...\psi (kk times), applying Corollary 1.2, we get our result.

6. Application to the Titchmarsh divisor problem

As an application of our main result for the prime indicator function, we first prove Theorem 1.5 in this section.

Proof on Theorem 1.5.

For a fixed a𝔽q[t]a\in{\mathbb{F}}_{q}[t], consider the sum

S:=P1,P2deg(P1P2)=Nτ(P1P2a),S:=\sum\limits_{\begin{subarray}{c}P_{1},P_{2}\\ \deg(P_{1}P_{2})=N\end{subarray}}\tau(P_{1}P_{2}-a), (6.1)

where PiP_{i} represents a monic irreducible polynomial in 𝔽q[t]{\mathbb{F}}_{q}[t]. Let NN be large enough so that deg(a)<N\deg(a)<N. We split the sum SS as follows.

S\displaystyle S =deg(P1P2)=Nd|(P1P2a)1\displaystyle=\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sum\limits_{d|(P_{1}P_{2}-a)}1
=deg(d)N(d,a)=1deg(P1P2)=NP1P2a(modd)1+deg(d)N(d,a)>1deg(P1P2)=NP1P2a(modd)1\displaystyle=\sum\limits_{\begin{subarray}{c}\deg(d)\leq N\\ (d,a)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1+\sum\limits_{\begin{subarray}{c}\deg(d)\leq N\\ (d,a)>1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1
=:S+S′′(say).\displaystyle=:S^{\prime}+S^{\prime\prime}\qquad\text{(say)}.

Let us first consider the sum S′′S^{\prime\prime}. We can reduce the sum over dd as follows.

S′′=\displaystyle S^{\prime\prime}= deg(P1P2)=Nd|(P1P2a)deg(d)N2,(d,a)>11+deg(P1P2)=Nd|(P1P2a)deg(d)>N2,(d,a)>11\displaystyle\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sum\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)\leq\frac{N}{2},(d,a)>1\end{subarray}}1+\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)>\frac{N}{2},(d,a)>1\end{subarray}}1
\displaystyle\leq 2deg(d)N2(d,a)>1deg(P1P2)=NP1P2a(modd)1\displaystyle 2\sum\limits_{\begin{subarray}{c}\deg(d)\leq\frac{N}{2}\\ (d,a)>1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1

Observe that the inner sum is non trivial only when the gcd (d,a)(d,a) is a prime. Therefore, this can be bounded using (1.5) and Lemma 2.3 as

S′′\displaystyle S^{\prime\prime}\leq 2P|adeg(d)N2(d,a)=Pdeg(P1)=Ndeg(P)P1aP(moddP)1\displaystyle 2\sum_{P|a}\sum_{\begin{subarray}{c}\deg(d)\leq\frac{N}{2}\\ (d,a)=P\end{subarray}}~{}\sum_{\begin{subarray}{c}\deg(P_{1})=N-\deg(P)\\ P_{1}\equiv\frac{a}{P}\mkern 4.0mu({\operator@font mod}\mkern 6.0mu\frac{d}{P})\end{subarray}}1
\displaystyle\ll P|aqNdeg(P)Ndeg(P)deg(dP)N2deg(P)1ϕ(dP)qN.\displaystyle\sum_{P|a}\frac{q^{N-\deg(P)}}{N-\deg(P)}\sum_{\deg\left(\frac{d}{P}\right)\leq\frac{N}{2}-\deg(P)}\frac{1}{\phi\left(\frac{d}{P}\right)}\ll q^{N}.

Therefore,

S=S+O(qN).\displaystyle S=S^{\prime}+O\left(q^{N}\right). (6.2)

We now consider SS^{\prime} and split the sum in SS^{\prime} as follows.

S=\displaystyle S^{\prime}= deg(P1P2)=Nd|(P1P2a)deg(d)<N21+deg(P1P2)=Nd|(P1P2a)deg(d)>N21+deg(P1P2)=Nd|(P1P2a)deg(d)=N21\displaystyle\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)<\frac{N}{2}\end{subarray}}1+\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)>\frac{N}{2}\end{subarray}}1+\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)=\frac{N}{2}\end{subarray}}1
=\displaystyle= 2deg(P1P2)=Nd|(P1P2a)deg(d)<N21+deg(P1P2)=Nd|(P1P2a)deg(d)=N21\displaystyle 2\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)<\frac{N}{2}\end{subarray}}1+\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}d|(P_{1}P_{2}-a)\\ \deg(d)=\frac{N}{2}\end{subarray}}1
=:\displaystyle=: 2S1+S2(say).\displaystyle 2S_{1}+S_{2}\qquad\text{(say)}. (6.3)

Here and now onwards, \sideset{}{{}^{\prime}}{\sum} represents that the sum is restricted to dd with (d,a)=1(d,a)=1. Let Q:=N2BlogNQ:=\frac{N}{2}-B\log N, for some B>0B>0 to be chosen later. We split the range of deg(d)\deg(d) in S1S_{1} into two parts, which gives

S1=\displaystyle S_{1}= deg(d)Q(d,a)=1deg(P1P2)=NP1P2a(modd)1+Q<deg(d)<N2(d,a)=1deg(P1P2)=NP1P2a(modd)1\displaystyle\sum\limits_{\begin{subarray}{c}\deg(d)\leq Q\\ (d,a)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1+\sum\limits_{\begin{subarray}{c}Q<\deg(d)<\frac{N}{2}\\ (d,a)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1
=:\displaystyle=: S11+S12(say).\displaystyle\,S_{11}+S_{12}\qquad\text{(say)}. (6.4)

First, we concentrate on the sum S11S_{11}. We have

S11=\displaystyle S_{11}= deg(d)Qdeg(P1P2)=NP1P2a(modd)1\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1
=\displaystyle= deg(d)Qdeg(f)=Nfa(modd)𝟙𝒫𝟙𝒫(f)\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(f)=N\\ f\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}}(f)
=\displaystyle= deg(d)Q(1ϕ(d)deg(f)=N(f,d)=1𝟙𝒫𝟙𝒫(f)+E(N;d,a;𝟙𝒫𝟙𝒫)),\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\Bigg{(}\frac{1}{\phi(d)}\sum\limits_{\begin{subarray}{c}\deg(f)=N\\ (f,\,d)=1\end{subarray}}\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}}(f)+E(N;d,a;\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}})\Bigg{)},

where E(N;d,a;𝟙𝒫𝟙𝒫)E(N;d,a;\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}}) is as defined in (1.3). Therefore,

S11=\displaystyle S_{11}= deg(d)Q1ϕ(d)deg(P1P2)=N(P1P2,d)=11+deg(d)QE(N;d,a;𝟙𝒫𝟙𝒫).\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\frac{1}{\phi(d)}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ (P_{1}P_{2},\,d)=1\end{subarray}}1+\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}E(N;d,a;\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}}).

Let A=1A=1 in Corollary 1.3 and choose B=B(1)B=B(1). Thus we have

deg(d)QE(N;d,1;𝟙𝒫𝟙𝒫)=\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}E(N;d,1;\mathbbm{1}_{\mathcal{P}}\ast\mathbbm{1}_{\mathcal{P}})= O(qNN),\displaystyle O\left(\frac{q^{N}}{N}\right),

Using this bound we further write the sum as

S11=\displaystyle S_{11}= S111S112+O(qNN),\displaystyle S_{11}^{1}-S_{11}^{2}+O\left(\frac{q^{N}}{N}\right),

where

S111:=\displaystyle S_{11}^{1}:= deg(d)Q1ϕ(d)deg(P1P2)=N1andS112:=\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\frac{1}{\phi(d)}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\end{subarray}}1\qquad\text{and}\qquad S_{11}^{2}:= deg(d)Q1ϕ(d)deg(P1P2)=N(P1P2,d)11.\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\frac{1}{\phi(d)}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ (P_{1}P_{2},\,d)\neq 1\end{subarray}}1.

The main contribution to our sum comes from S111S_{11}^{1}. We determine this as follows.

S111=\displaystyle S_{11}^{1}= deg(d)Q1ϕ(d)k=1N1deg(P1)=k1deg(P2)=Nk1.\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}\deg(d)\leq Q\end{subarray}}\frac{1}{\phi(d)}\sum\limits_{k=1}^{N-1}\sum\limits_{\deg(P_{1})=k}1\sum\limits_{\deg(P_{2})=N-k}1. (6.5)

Using (1.4) to compute the inner sum, we get,

k=1N1deg(P1)=k1deg(P2)=Nk1=\displaystyle\sum\limits_{k=1}^{N-1}\sum\limits_{\deg(P_{1})=k}1\sum\limits_{\deg(P_{2})=N-k}1= k>logNNlogNdeg(P1)=Nk1deg(P2)=k1+2k=1logNdeg(P1)=Nk1deg(P2)=k1\displaystyle\sum\limits_{k>\log N}^{N-\log N}\sum\limits_{\deg(P_{1})=N-k}1\sum\limits_{\deg(P_{2})=k}1+2\sum\limits_{k=1}^{\log N}\sum\limits_{\deg(P_{1})=N-k}1\sum\limits_{\deg(P_{2})=k}1
=\displaystyle= k>logNNlogNqkk(1+O(qk2))qNkNk(1+O(qNk2))\displaystyle\sum\limits_{k>\log N}^{N-\log N}\frac{q^{k}}{k}\left(1+O\left(q^{-\frac{k}{2}}\right)\right)\frac{q^{N-k}}{N-k}\left(1+O\left(q^{-\frac{N-k}{2}}\right)\right)
+\displaystyle+ O(qNNloglogN)\displaystyle O\left(\frac{q^{N}}{N}\log\log N\right)
=\displaystyle= qNNk>logNNlogN(1k+1Nk)(1+O(qk2)+O(qNk2))\displaystyle\frac{q^{N}}{N}\sum\limits_{k>\log N}^{N-\log N}\left(\frac{1}{k}+\frac{1}{N-k}\right)\left(1+O\left(q^{-\frac{k}{2}}\right)+O\left(q^{-\frac{N-k}{2}}\right)\right)
+\displaystyle+ O(qNNloglogN)\displaystyle O\left(\frac{q^{N}}{N}\log\log N\right)
=\displaystyle= 2qNN(logN+O(loglogN))(1+O(qlogN2))+O(qNNloglogN)\displaystyle\frac{2q^{N}}{N}\left(\log N+O\left(\log\log N\right)\right)\left(1+O\left(q^{-\frac{\log N}{2}}\right)\right)+O\left(\frac{q^{N}}{N}\log\log N\right)
=\displaystyle= 2qNNlogN+O(qNlogNN32)+O(qNloglogNN).\displaystyle\frac{2q^{N}}{N}\log N+O\left(\frac{q^{N}\log N}{N^{\frac{3}{2}}}\right)+O\left(\frac{q^{N}\log\log N}{N}\right). (6.6)

Substituting (6.6) into (6.5) and using Lemma 2.3 for the fixed polynomial aa, we obtain

S111=\displaystyle S_{11}^{1}= (Caζq(2)ζq(3)ζq(6)Q+O(1))(2qNNlogN+O(qNlogNN32)+O(qNloglogNN))\displaystyle\left(C_{a}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}Q+O(1)\right)\left(\frac{2q^{N}}{N}\log N+O\left(\frac{q^{N}\log N}{N^{\frac{3}{2}}}\right)+O\left(\frac{q^{N}\log\log N}{N}\right)\right)
=\displaystyle= 2Caζq(2)ζq(3)ζq(6)qNNQ(logN)+O(qNloglogN),\displaystyle 2C_{a}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}\frac{q^{N}}{N}Q(\log N)+O\left(q^{N}\log\log N\right),

where CaC_{a} is as in (1.8). Now, putting Q=N2BlogNQ=\frac{N}{2}-B\log N, we have

S111=\displaystyle S_{11}^{1}= Caζq(2)ζq(3)ζq(6)qNlogN+O(qNloglogN).\displaystyle C_{a}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}q^{N}\log N+O\left(q^{N}\log\log N\right). (6.7)

We will now show that the contribution from S112S_{11}^{2} is negligible. If (P1P2,d)1(P_{1}P_{2},\,d)\neq 1, we have P1|dP_{1}|d or P2|dP_{2}|d. Since both the cases are symmetric, we have

S112\displaystyle S_{11}^{2}\ll deg(d)Q1ϕ(d)deg(P1P2)=NP1|d1.\displaystyle\sum\limits_{\deg(d)\leq Q}\frac{1}{\phi(d)}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}|d\end{subarray}}1.

By writing d=P1mkd=P_{1}^{m}k such that (P1,k)=1(P_{1},k)=1, we see that

S112\displaystyle S_{11}^{2}\ll deg(P1P2)=Nm11ϕ(P1m)deg(k)Q1ϕ(k).\displaystyle\sum\limits_{\deg(P_{1}P_{2})=N}\sum\limits_{m\geq 1}\frac{1}{\phi(P_{1}^{m})}\sum\limits_{\deg(k)\leq Q}\frac{1}{\phi(k)}.

Further, using Lemma 2.3 with g=1g=1, the innermost sum is Q\ll Q. Using the bound m11ϕ(P1m)1|P1|\sum\limits_{m\geq 1}\frac{1}{\phi(P_{1}^{m})}\ll\frac{1}{|P_{1}|} and (1.4), we have

S112\displaystyle S_{11}^{2}\ll 2Qk=1logNdeg(P1)=k1|P1|deg(P2)=Nk1+Qk>logNNlogNdeg(P1)=Nk1|P1|deg(P2)=k1\displaystyle 2Q\sum\limits_{k=1}^{\log N}\sum\limits_{\deg(P_{1})=k}\frac{1}{|P_{1}|}\sum\limits_{\deg(P_{2})=N-k}1+Q\sum\limits_{k>\log N}^{N-\log N}\sum\limits_{\deg(P_{1})=N-k}\frac{1}{|P_{1}|}\sum\limits_{\deg(P_{2})=k}1
\displaystyle\ll NloglogN+qNNlogNqNNlogN.\displaystyle N\log\log N+\frac{q^{N}}{N}\log N\ll\frac{q^{N}}{N}\log N. (6.8)

Therefore, from (6.7) and (6.8), we have

S11=\displaystyle S_{11}= Caζq(2)ζq(3)ζq(6)qNlogN+O(qNloglogN).\displaystyle C_{a}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}q^{N}\log N+O\left(q^{N}\log\log N\right). (6.9)

Next, we estimate the sum S12S_{12} by using the analogue of the Brun-Titchmarsh inequality over 𝔽q[t]{\mathbb{F}}_{q}[t] due to Hsu as stated in Lemma 2.4. Recall that

S12=Q<deg(d)<N2deg(P1P2)=NP1P2a(modd)1.\displaystyle S_{12}=\sideset{}{{}^{\prime}}{\sum}\limits_{Q<\deg(d)<\frac{N}{2}}\sum\limits_{\begin{subarray}{c}\deg(P_{1}P_{2})=N\\ P_{1}P_{2}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1.

From Lemma 2.4 and (1.4), we see that

S12=\displaystyle S_{12}= Q<deg(d)<N2(kN2deg(P1)=k(P1,d)=1deg(P2)=NkP2aP11(modd)1+k<N2deg(P2)=k(P2,d)=1deg(P1)=NkP1aP21(modd)1)\displaystyle\sideset{}{{}^{\prime}}{\sum}\limits_{\begin{subarray}{c}Q<\deg(d)<\frac{N}{2}\end{subarray}}\left(\sum\limits_{k\leq\frac{N}{2}}\sum\limits_{\begin{subarray}{c}\deg(P_{1})=k\\ (P_{1},d)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{2})=N-k\\ P_{2}\equiv aP_{1}^{-1}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1+\sum\limits_{k<\frac{N}{2}}\sum\limits_{\begin{subarray}{c}\deg(P_{2})=k\\ (P_{2},d)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1})=N-k\\ P_{1}\equiv aP_{2}^{-1}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1\right)
\displaystyle\ll Q<deg(d)<N2kN2(qkk)(qNkϕ(d)(Nk+degd+1))BqNN(logN)2,\displaystyle\sum\limits_{\begin{subarray}{c}Q<\deg(d)<\frac{N}{2}\end{subarray}}\sum\limits_{k\leq\frac{N}{2}}\left(\frac{q^{k}}{k}\right)\left(\frac{q^{N-k}}{\phi(d)(N-k+\deg d+1)}\right)\ll_{B}\frac{q^{N}}{N}(\log N)^{2}, (6.10)

where aPi1aP_{i}^{-1} is the residue class modulo dd such that PiPi1a(modd)P_{i}P_{i}^{-1}\equiv a\mkern 4.0mu({\operator@font mod}\mkern 6.0mud). Here we have used Lemma 2.3 to write the final bound. Using (6.9) and (6.10), we have from (6.4) that

S1=\displaystyle S_{1}= Caζq(2)ζq(3)ζq(6)qNlogN+O(qNloglogN).\displaystyle C_{a}\frac{\zeta_{q}(2)\zeta_{q}(3)}{\zeta_{q}(6)}q^{N}\log N+O\left(q^{N}\log\log N\right). (6.11)

It remains to estimate S2S_{2}. Again we split the sum as

S2=\displaystyle S_{2}= 2degd=N2(d,a)=1k<N2deg(P1)=k(P1,d)=1deg(P2)=NkP2aP11(modd)1+degd=N2(d,a)=1deg(P2)=N2(P2,d)=1deg(P1)=N2P1aP21(modd)1.\displaystyle 2\sum_{\begin{subarray}{c}\deg d=\frac{N}{2}\\ (d,a)=1\end{subarray}}\sum\limits_{k<\frac{N}{2}}\sum\limits_{\begin{subarray}{c}\deg(P_{1})=k\\ (P_{1},d)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{2})=N-k\\ P_{2}\equiv aP_{1}^{-1}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1+\sum_{\begin{subarray}{c}\deg d=\frac{N}{2}\\ (d,a)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{2})=\frac{N}{2}\\ (P_{2},d)=1\end{subarray}}\sum\limits_{\begin{subarray}{c}\deg(P_{1})=\frac{N}{2}\\ P_{1}\equiv aP_{2}^{-1}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud)\end{subarray}}1.

The first term in the above expression is BqN(logN)2/N\ll_{B}q^{N}(\log N)^{2}/N as done in (6.10). For the second term, observe that the arithmetic progression aP21(modd)aP_{2}^{-1}\mkern 4.0mu({\operator@font mod}\mkern 6.0mud) can contain atmost one prime of degree N2\frac{N}{2}. Indeed, for a general term aP21+fdaP_{2}^{-1}+fd in the arithmetic progression, ff must have degree zero as deg(d)=N2\deg(d)=\frac{N}{2}. Since we are counting monics, ff must be 11. Thus, using (1.4),we have that the second term above is qN/N\ll q^{N}/N. Thus, we have

S2BqNN(logN)2.\displaystyle S_{2}\ll_{B}\frac{q^{N}}{N}(\log N)^{2}. (6.12)

Using (6.11) and (6.12) in (6.3) and (6.2) completes the proof of Theorem 1.5. ∎

Proof of Theorem 1.6.

Proof follows by taking qq\rightarrow\infty Theorem 1.5 with slight modification. Note that for qq\rightarrow\infty, we have S"qN1S"\ll q^{N-1}. Also, from equation (6.6) we have

k=1N1degP1=k1degP2=Nk1\displaystyle\sum_{k=1}^{N-1}\sum_{\deg P_{1}=k}1\sum_{\deg P_{2}=N-k}1 =k=1N1qkk(1+O(qk/2))qNkNk(1+O(q(Nk)/2))\displaystyle=\sum_{k=1}^{N-1}\frac{q^{k}}{k}\left(1+O\left(q^{-k/2}\right)\right)\frac{q^{N-k}}{N-k}\left(1+O\left(q^{-(N-k)/2}\right)\right)
=2qNNk=1N11k(1+O(q1/2))\displaystyle=\frac{2q^{N}}{N}\sum_{k=1}^{N-1}\frac{1}{k}\left(1+O\left(q^{-1/2}\right)\right)
=2qNN(logN+γ)+O(qNN2)+O(qN12NlogN),\displaystyle=\frac{2q^{N}}{N}(\log N+\gamma)+O\left(\frac{q^{N}}{N^{2}}\right)+O\left(\frac{q^{N-\frac{1}{2}}}{N}\log N\right),

as qq\rightarrow\infty. Therefore,

S111=qN(logN+γ)+O(qNN(logN)2)+O(qN12logN),\displaystyle S_{11}^{1}=q^{N}(\log N+\gamma)+O\left(\frac{q^{N}}{N}(\log N)^{2}\right)+O\left(q^{N-\frac{1}{2}}\log N\right), (6.13)

as the constant in Lemma 2.3 tends to 0 as qq\rightarrow\infty. Considering qq\rightarrow\infty in rest of the argument, we derive the result. ∎

Acknowledgments. Both the authors express their sincere gratitude to Prof. Akshaa Vatwani for suggesting the problem that led to this work, as well as for her insightful discussions and suggestions on the relevant references. The second author is very thankful to Prof. M. Ram Murty for his lectures on “Arithmetic in function fields” at IIT Gandhinagar which were instrumental in motivating this project. The funding from the MHRD SPARC project SPARC/2018 -2019/P567/SL, under which these lectures were organized is gratefully acknowledged.

References

  • [1] A. Akbary and D. Ghioca, A geometric variant of Titchmarsh divisor problem, Int. J. Number Theory 8 (2012), no. 1, 53–69. MR 2887882
  • [2] J. C. Andrade, L. Bary-Soroker, and Z. Rudnick, Shifted convolution and the Titchmarsh divisor problem over 𝔽q[t]\mathbb{F}_{q}[t], Philos. Trans. Roy. Soc. A 373 (2015), no. 2040, 20140308, 18. MR 3338116
  • [3] S. Baier and R. K. Singh, Large sieve inequality with power moduli for function fields, J. Number Theory 196 (2019), 1–13. MR 3906465
  • [4] by same author, Erratum to “Large sieve inequality with power moduli for function fields” [J. Number Theory 196 (2019) 1–13], J. Number Theory 210 (2020), 431–432. MR 4057535
  • [5] by same author, The large sieve inequality with square moduli for quadratic extensions of function fields, Int. J. Number Theory 16 (2020), no. 9, 1907–1922. MR 4153360
  • [6] E. Bombieri, J. B. Friedlander, and H. Iwaniec, Primes in arithmetic progressions to large moduli, Acta Math. 156 (1986), no. 3-4, 203–251. MR 834613
  • [7] by same author, Primes in arithmetic progressions to large moduli. II, Math. Ann. 277 (1987), no. 3, 361–393. MR 891581
  • [8] by same author, Primes in arithmetic progressions to large moduli. III, J. Amer. Math. Soc. 2 (1989), no. 2, 215–224. MR 976723
  • [9] A. C. Cojocaru and M. R. Murty, An introduction to sieve methods and their applications, London Mathematical Society Student Texts, vol. 66, Cambridge University Press, Cambridge, 2006. MR 2200366
  • [10] P. Darbar and A. Mukhopadhyay, A Bombieri-type theorem for convolution with application on number field, Acta Math. Hungar. 163 (2021), no. 1, 37–61. MR 4217957
  • [11] S. Drappeau, Sums of Kloosterman sums in arithmetic progressions, and the error term in the dispersion method, Proc. Lond. Math. Soc. (3) 114 (2017), no. 4, 684–732. MR 3653244
  • [12] S. Drappeau and B. Topacogullari, Combinatorial identities and Titchmarsh’s divisor problem for multiplicative functions, Algebra Number Theory 13 (2019), no. 10, 2383–2425. MR 4047638
  • [13] A. T. Felix, Generalizing the Titchmarsh divisor problem, Int. J. Number Theory 8 (2012), no. 3, 613–629. MR 2904920
  • [14] É. Fouvry, Sur le problème des diviseurs de Titchmarsh, J. Reine Angew. Math. 357 (1985), 51–76. MR 783533
  • [15] É. Fouvry and H. Iwaniec, On a theorem of Bombieri-Vinogradov type, Mathematika 27 (1980), no. 2, 135–152 (1981). MR 610700
  • [16] by same author, Primes in arithmetic progressions, Acta Arith. 42 (1983), no. 2, 197–218. MR 719249
  • [17] A. Fujii, A local study of some additive problems in the theory of numbers, Proc. Japan Acad. 52 (1976), no. 3, 113–115. MR 399021
  • [18] P. X. Gallagher, The large sieve, Mathematika 14 (1967), 14–20. MR 214562
  • [19] A. Granville and X. Shao, Bombieri-Vinogradov for multiplicative functions, and beyond the x1/2x^{1/2}-barrier, Adv. Math. 350 (2019), 304–358. MR 3947647
  • [20] H. Halberstam, Footnote to the Titchmarsh-Linnik divisor problem, Proc. Amer. Math. Soc. 18 (1967), 187–188. MR 204379
  • [21] J. G. Hinz, Methoden des grossen Siebes in algebraischen Zahlkörpern, Manuscripta Math. 57 (1987), no. 2, 181–194. MR 871630
  • [22] by same author, A generalization of Bombieri’s prime number theorem to algebraic number fields, Acta Arith. 51 (1988), no. 2, 173–193. MR 975109
  • [23] C. Hsu, A large sieve inequality for rational function fields, J. Number Theory 58 (1996), no. 2, 267–287. MR 1393616
  • [24] by same author, Applications of the large sieve inequality for 𝐅q[T]\mathbf{F}_{q}[T], Finite Fields Appl. 4 (1998), no. 3, 275–281. MR 1640777
  • [25] M. N. Huxley, The large sieve inequality for algebraic number fields, Mathematika 15 (1968), 178–187. MR 237455
  • [26] by same author, The large sieve inequality for algebraic number fields. III. Zero-density results, J. London Math. Soc. (2) 3 (1971), 233–240. MR 276196
  • [27] J. Johnsen, On the large sieve method in GF[q,x]{\rm GF}[q,\,x], Mathematika 18 (1971), 172–184. MR 302617
  • [28] O. Klurman, A. P. Mangerel, and J. Teräväinen, Correlations of multiplicative functions in function fields, arXiv preprint arXiv:2009.13497 (2020).
  • [29] U. V. Linnik, The large sieve., C. R. (Doklady) Acad. Sci. URSS (N.S.) 30 (1941), 292–294. MR 0004266
  • [30] by same author, The dispersion method in binary additive problems, American Mathematical Society, Providence, R.I., 1963, Translated by S. Schuur. MR 0168543
  • [31] J. Maynard, Primes in arithmetic progressions to large moduli I: fixed residue classes, arXiv preprint arXiv:2006.06572 (2020).
  • [32] by same author, Primes in arithmetic progressions to large moduli II: Well-factorable estimates, arXiv preprint arXiv:2006.07088 (2020).
  • [33] by same author, Primes in arithmetic progressions to large moduli III: Uniform residue classes, arXiv preprint arXiv:2006.08250 (2020).
  • [34] Y. Motohashi, An induction principle for the generalization of Bombieri’s prime number theorem, Proc. Japan Acad. 52 (1976), no. 6, 273–275. MR 422179
  • [35] G. Rodriquez, Sul problema dei divisori di Titchmarsh, Boll. Un. Mat. Ital. (3) 20 (1965), 358–366. MR 0197409
  • [36] M. Rosen, Number theory in function fields, Graduate Texts in Mathematics, vol. 210, Springer-Verlag, New York, 2002. MR 1876657
  • [37] W. Sawin, Square-root cancellation for sums of factorization functions over short intervals in function fields, Duke Mathematical Journal 170 (2021), no. 5, 997–1026.
  • [38] by same author, Square-root cancellation for sums of factorization functions over squarefree progressions in 𝔽q[t]\mathbb{F}_{q}[t], arXiv preprint arXiv:2102.09730 (2021).
  • [39] W. Sawin and M. Shusterman, On the Chowla and twin primes conjectures over 𝔽q[T]\mathbb{F}_{q}[T], arXiv preprint arXiv:1808.04001 (2018).
  • [40] G. Tenenbaum, Introduction to analytic and probabilistic number theory, third ed., Graduate Studies in Mathematics, vol. 163, American Mathematical Society, Providence, RI, 2015, Translated from the 2008 French edition by Patrick D. F. Ion. MR 3363366
  • [41] C. E. Titchmarsh, A divisor problem, Rend. Circ. Mat. Palermo 54 (1930), 414–429.
  • [42] A. Vatwani and P. Wong, On generalizations of the Titchmarsh divisor problem, Acta Arith. 193 (2020), no. 4, 321–337. MR 4074394
  • [43] Y. Zhang, Bounded gaps between primes, Ann. of Math. (2) 179 (2014), no. 3, 1121–1174. MR 3171761