1. Introduction
The theory of L-functions plays a crucial role in both number theory and arithmetic geometry. L-Functions exhibit natural connections with various mathematical subjects including number fields, automorphic forms, Artin representations, Shimura varieties, abelian varieties and intersection theory. The central values of L-functions and their derivatives reveal important connections to the geometric and arithmetic properties of Shimura varieties such as the Gross-Zagier formula, Colmez’s conjecture and the averaged Colmez formula. On the other hand, vector-valued modular forms are important generalizations of elliptic modular forms that arise naturally in the theory of Jacobi forms, Siegel modular forms and Moonshine. Important foundational results on vector-valued modular forms were establihsed by Knopp and Mason [Reference Knopp and Mason13, Reference Knopp and Mason14]. Being an important tool to tackle classical problems in the theory of modular forms, Selberg used these forms to give an estimation for the Fourier coefficients of the classical modular forms [Reference Selberg20]. Borcherds in [Reference Borcherds4] and [Reference Borcherds6] used vector-valued modular forms associated with Weil representations to provide a description of the Fourier expansion of various theta liftings. Some applications of vector-valued modular forms stand out in high-energy physics by mainly providing a method of differential equations in order to construct the modular multiplets and also revealing the simple structure of the modular invariant mass models [Reference Liu and Ding18]. Other applications concerning vector-valued modular forms of half-integer weight seem to provide a simple solution to the Riemann–Hilbert problem for representations of the modular group [Reference Bantay and Gannon2]. So it is only natural to study the L-functions of vector-valued modular forms and their properties as a buildup that aligns with the development of a Hecke theory to the space of vector-valued modular forms. L-Functions of vector-valued modular forms have been investigated in connection with sign changes of Fourier coefficients, Jacobi forms and Hecke theory (for examples, see [Reference Berndt3, Reference Bruinier and Stein7, Reference Jin and Lim11]). Moreover, they lead us to investigate a Gross–Kohnen–Zagier theorem in higher dimensions [Reference Borcherds5].
In [Reference Lim and Raji17], we show that averages of L-functions associated with vector-valued cusp forms do not vanish when the average is taken over the orthogonal basis of the space of vector-valued cusp forms. To illustrate, we let $\{f_{k,1}, \ldots, f_{k,d_k}\}$ be an orthogonal basis of the space $S_{k,\chi,\rho}$ of vector-valued cusp forms with Fourier coefficients $b_{k,l,j}(n)$, where χ is a multiplier system of weight $k\in\frac12 \mathbb{Z}$ on $\mathrm{SL}_2(\mathbb{Z})$ and $\rho:\mathrm{SL}_2(\mathbb{Z})\to \mathrm{GL}_m(\mathbb{C})$ is an m-dimensional unitary complex representation. We also let $t_0\in\mathbb{R}, \epsilon \gt 0$ and $1\leq i\leq m$. Then, there exists a constant $C(t_0, \epsilon, i) \gt 0$ such that for $k \gt C(t_0, \epsilon, i)$, the function
does not vanish at any point $s=\sigma + it_0$, with $\frac{k-1}{2} \lt \sigma \lt \frac k2 - \epsilon$, where $ \lt L^*(f_{k,l}, s),\mathbf{e}_i \gt $ denotes the ith component of $L^*(f_{k,l}, s)$ and $n_{i,0}$ is the number given in equation (3.2).
Kohnen, Sengupta and Weigel in [Reference Kohnen, Sengupta and Weigel16] proved a nonvanishing result for the derivatives of L-functions in the critical strip for elliptic modular forms on the full group. In [Reference Raji19], the second author generalized their result to modular forms of half-integer weight on the plus space. In this paper, we show analogous results for the averages of the derivatives of L-functions for the orthogonal basis of the space of vector-valued cusp forms in the critical strip. In particular, given $k\in\frac12 \mathbb{Z}$, χ a multiplier system of weight k on $\mathrm{SL}_2(\mathbb{Z})$, $t_0\in\mathbb{R}, \epsilon \gt 0, 1\leq i\leq m$ and n a positive integer, we show that there exists a constant $C(t_0, \epsilon,i, n) \gt 0$ such that for $k \gt C(t_0, \epsilon, i, n)$, the function
does not vanish at any point $s = \sigma + it$, with $t = t_0, \frac{k-1}{2} \lt \sigma \lt \frac k2 -\epsilon$. From this, we show that there exists a constant $C(t_0, \epsilon, n)$ such that for $k \gt C(t_0, \epsilon, n)$ and any $s = \sigma + it$, with $t = t_0,\ \frac{k-1}2 \lt \sigma \lt \frac k2 - \epsilon,\ \frac k2+\epsilon \lt \sigma \lt \frac{k+1}2$, there exists $f\in S_{k,\chi,\rho}$ such that $\frac{d^n}{ds^n} L^*(f, s) \neq 0$.
The isomorphism between the space of Jacobi forms of weight k and index m on $\mathrm{SL}_2(\mathbb{Z})$ and the space of vector-valued modular cusp forms with a specific multiplier system and a given Weil representation depending on m leads to an analogous result for Jacobi forms. We also give a similar result for cusp forms in the plus space.
2. The Kernel function
In this section, we define the kernel function $R_{k,s,i}$ and determine its Fourier expansion. The kernel function being a cusp form will play an important role in determining the coefficients of a given cusp form in terms of L-functions when the given cusp form is written in terms of the orthogonal basis. So, let $\Gamma = \mathrm{SL}_2(\mathbb{Z})$, $k\in \frac12 \mathbb{Z}$ and χ a unitary multiplier system of weight k on Γ, i.e. $\chi:\Gamma\to\mathbb{C}$ satisfies the following conditions:
(1) $|\chi(\gamma)| = 1$ for all $\gamma\in \Gamma$.
(2) χ satisfies the consistency condition
\begin{align*} \chi(\gamma_3) (c_3\tau + d_3)^k = \chi(\gamma_1)\chi(\gamma_2) (c_1\gamma_2\tau + d_1)^k (c_2\tau+d_2)^k, \end{align*}where $\gamma_3 = \gamma_1\gamma_2$ and $\gamma_i = \left(\begin{smallmatrix} a_i&b_i\\ c_i&d_i\end{smallmatrix}\right)\in \Gamma$ for $i=1,2$ and 3.
Let m be a positive integer and $\rho:\Gamma\to \mathrm{GL}(m, \mathbb{C})$ an m-dimensional unitary complex representation. Let $\{\mathbf{e}_1, \ldots, \mathbf{e}_m\}$ denote the standard basis of $\mathbb{C}^m$. For a vector-valued function $f = \sum_{j=1}^m f_j \mathbf{e}_j$ on $\mathbb{H}$ and $\gamma\in \Gamma$, define a slash operator by
The definition of the vector-valued modular forms is given as follows.
Definition 2.1. A vector-valued modular form of weight $k\in\frac12\mathbb{Z}$, multiplier system χ and type ρ on Γ is a sum $f = \sum_{j=1}^m f_j \mathbf{e}_j$ of functions holomorphic in $\mathbb{H}$ satisfying the following conditions:
(1) $f|_{k,\chi,\rho}\gamma = f$ for all $\gamma\in \Gamma$.
(2) For each $1\leq j\leq m$, each function fj has a Fourier expansion of the form
\begin{align*} f_i(\tau) = \sum_{n+\kappa_j\geq 0} a_j(n)\text{e}^{2\pi i(n+\kappa_j)\tau}. \end{align*}Here and throughout the paper, κj is a certain positive number with $0\leq \kappa_j \lt 1$.
The space of all vector-valued modular forms of weight k, multiplier system χ and type ρ on Γ is denoted by $M_{k,\chi,\rho}$. There is a subspace $S_{k,\chi,\rho}$ of vector-valued cusp forms for which we require that each $a_j(n) = 0$ when $n+\kappa_j$ is non-positive.
Following [Reference Kim and Lim12], we now define the L-function of a vector-valued cusp form. For a vector-valued cusp form $f(\tau) = \sum_{j=1}^m \sum_{n+\kappa_j \gt 0} a_j(n)\text{e}^{2\pi i(n+\kappa_j)\tau}\mathbf{\text{e}}_j\in S_{k,\chi,\rho}$, we see that $a_j(n) = O(n^{k/2})$ for every $1\leq j\leq m$ as $n\to\infty$ by the same argument for elliptic modular forms. Then, the vector-valued L-function defined by
converges absolutely for $\mathrm{Re}(s)\gg0$. This has an integral representation
From this, we see that it has an analytic continuation to $\mathbb{C}$ and a functional equation given by
where $L^*(f,s) = \frac{\Gamma(s)}{(2\pi)^s} L(f,s)$ and $S = \left(\begin{smallmatrix} 0&-1\\1&0\end{smallmatrix}\right)$.
Let i be an integer with $1\leq i\leq m$. Define
For $s\in\mathbb{C}$ with $1 \lt \mathrm{Re}(s) \lt k-1$, we define the kernel function by
where $\gamma_{k}(s) := \frac12 \text{e}^{\pi is/2} \Gamma(s) \Gamma(k-s)$. Then, this series converges absolutely uniformly whenever $\tau = u+iv$ satisfies $v\geq \epsilon, v\leq 1/\epsilon$ for a given ϵ > 0 and s varies over a compact set. Moreover, it is a vector-valued cusp form in $S_{k,\chi,\rho}$.
We write $ \lt \cdot, \cdot \gt $ for the standard scalar product on $\mathbb{C}^m$, i.e.
Then, for $f,g\in S_{k,\chi,\rho}$, we define the Petersson scalar product of f and g by
where $\mathcal{F}$ is the standard fundamental domain for the action of Γ on $\mathbb{H}$. Then, by [Reference Lim and Raji17, Lemma 3.1], we have
where $c_k := \frac{(-1)^{k/2} \pi (k-2)!}{2^{k-2}}$.
We can also compute the Fourier expansion of $R_{k,s,i}$.
Lemma 2.2. [Reference Lim and Raji17, Lemma 3.2] The function $R_{k,s,i}$ has the Fourier expansion
where $r_{k,s,i,j}(n)$ is given by
where $_1 F_1(\alpha, \beta;z)$ is Kummer’s degenerate hypergeometric function.
3. The main theorem
In this section, we give the main theorem for the existence of at least one L-function whose derivative does not vanish.
Theorem 3.1. Let $k\in\frac12 \mathbb{Z}$, and let χ be a multiplier system of weight k on $\mathrm{SL}_2(\mathbb{Z})$. Let $t_0\in \mathbb{R}, \epsilon \gt 0$ and n a positive integer. Then, there exists a constant $C(t_0, \epsilon, n)$ such that for $k \gt C(t_0, \epsilon, n)$ and any $s = \sigma + it$ with $t = t_0,\ \frac{k-1}2 \lt \sigma \lt \frac k2 - \epsilon,\ \frac k2+\epsilon \lt \sigma \lt \frac{k+1}2$, there exists $f\in S_{k,\chi,\rho}$ such that $\frac{d^n}{\text{d}s^n} L^*(f, s) \neq 0$.
Proof. We follow the argument in the proof of Theorem 3.1 in [Reference Kohnen, Sengupta and Weigel16]. Suppose that $\{f_{k,1}, \ldots, f_{k,d_k}\}$ is an orthogonal basis of $S_{k,\chi,\rho}$ with Fourier expansions
and dk is the dimension of $\dim S_{k,\chi,\rho}$.
For each $1\leq i\leq m$, by equation (2.1), we have
Let
If we take the first Fourier coefficients of ith component function on both sides of equation (3.1), then by Lemma 2.2 we have
where
We assume that
is zero. If we take the nth derivative with respect to s on both sides in equation (3.3), then we have
Then, the left-hand side of equation (3.5) is equal to
Then, we have
Let $\psi(s) := \frac{\Gamma^\prime(s)} {\Gamma(s)}$. Then, one can see that $\frac{\Gamma^{(n)}(s)}{\Gamma(s)}$ is a polynomial $P(\psi, \psi^{(1)}, \ldots, \psi^{(n-1)})$ with integral coefficients and it contains the term ψ n, which is the highest power of ψ occurring in P. It is known that ψ satisfies the following asymptotic formulas:
and
for $s\to\infty$ in $|\arg(s)| \lt \pi$, where Bn denotes the nth Bernoulli number (for example, see [Reference Abramowitz and Stegun1, 6.3.18 and 6.4.11]). Let $s = \frac{k}{2} - \delta + it_0\ (\epsilon \lt \delta \lt \frac12)$. Then, the leading term of $\frac{\Gamma^{(n-\nu)}(k-s)} {\Gamma(k-s)}$ for $0\leq \nu\leq n-1$ is $(\log(\frac k2 +\delta - it_0))^{n-\nu}$ as $k\to\infty$ and $\psi^{(n)}(s) = o\left( \frac{1}{|s|^n}\right)$ as $|s|\to\infty$ in $|\arg(s)| \lt \pi$ for $n\in\mathbb{N}$. Therefore, we have
as $k\to\infty$, where Q is a polynomial of degree n and its highest coefficient is $(-1)^n$.
For the first term on the right-hand side of equation (3.5), we have
If we divide this by $(2\pi)^s (n_{i,0}+\kappa_i)^{s-1} \Gamma(k-s)$, then we have
Let $s = \frac{k}{2} - \delta + it_0\ (\epsilon \lt \delta \lt \frac12)$. Then, by [Reference Abramowitz and Stegun1, 6.1.23 and 6.1.47], we have
where the O constant is absolute, uniformly in $\epsilon \lt \delta \lt \frac12$. On the other hand, the highest-order term in $\frac{\Gamma^{(n-\nu)}(s)}{\Gamma(s)}$ is $\left( \psi \left(\frac k2 -\delta + it_0\right)\right)^{n-\nu}$. This behaves like $(\log(\frac k2 - \delta + it_0))^{n-\nu}$ for $0\leq \nu \lt n$ as $k\to\infty$. Thus, we can see that all terms in the sum in equation (3.6) go to zero as $k\to\infty$.
The second term on the right-hand side of equation (3.5) is equal to
In the above equation, the derivative in the last two lines is equal to
By [Reference Abramowitz and Stegun1, 13.2.1], for $\mathrm{Re}(\beta) \gt \mathrm{Re}(\alpha) \gt 0$, we have
Therefore, for any $n\in\mathbb{Z}_{\geq0}$, we obtain
Since $\log(u)= o(u^{-\epsilon^\prime})$ for any $\epsilon^\prime \gt 0$ as u → 0, we see that
where Kn is a constant depending only on n.
Let $s = \frac k2 - \delta + it_0\ (\epsilon \lt \delta \lt \frac12)$. Then, the series in equation (3.7) is
This can be estimated in terms of the Riemann zeta function and a positive constant factor $B(t_0, n)$ depending only on t 0 and n. If we divide the second term on the right-hand side of equation (3.5) by $(2\pi)^s (n_{i,0}+\kappa_i)^{s-1} \Gamma(k-s)$, then the absolute value is
and this goes to 0 as $k\to\infty$ uniformly in $\delta\in (\epsilon, \frac12)$ by Stirling’s formula.
In conclusion, if we divide both sides of equation (3.5) by $(2\pi)^s (n_{i,0}+\kappa_i)^{s-1} \Gamma(k-s)$, then the right-hand side goes to zero as $k\to\infty$, but the absolute value of the left-hand side is
as $k\to\infty$. This is a contradiction.
Therefore, there exists a constant $C(t_0, \epsilon,i, n) \gt 0$ such that for $k \gt C(t_0, \epsilon, i, n)$, the function in equation (3.4) does not vanish at any point $s = \sigma + it$, with $t = t_0, \frac{k-1}{2} \lt \sigma \lt \frac k2 -\epsilon$. From this, we get the desired result by using the functional equation of $L^*(f, s)$ ($f\in S_{k,\chi,\rho}$).
(1) In [Reference Kohnen15, Reference Kohnen, Sengupta and Weigel16], one takes the average over the basis consisting of normalized Hecke eigenforms for the case of classical modular forms. In fact, the basis is an orthogonal basis, and the results in [Reference Kohnen15, Reference Kohnen, Sengupta and Weigel16] hold for any orthogonal basis of the space of cusp forms.
(2) The vector-valued cusp form f in Theorem 3.1 depends on n in general.
4. The case of $\Gamma_0(N)$
Now, we consider the case of an elliptic modular form of integral weight on the congruence subgroup $\Gamma_0(N)$. By using Theorem 3.1, we can extend a result in [Reference Kohnen, Sengupta and Weigel16] to the case of $\Gamma_0(N)$. To illustrate, let N be a positive integer and let k be a positive even integer. Let $\Gamma = \Gamma_0(N)$, and let $S_k(\Gamma)$ be the space of cusp forms of weight k on Γ. Let $\{\gamma_1, \ldots, \gamma_m\}$ be the set of representatives of $\Gamma \setminus \mathrm{SL}_2(\mathbb{Z})$, with $\gamma_1 = I$. For $f\in S_k(\Gamma)$, we define a vector-valued function $\tilde{f}:\mathbb{H}\to \mathbb{C}^m$ by $\tilde{f} = \sum_{j=1}^m f_j\mathbf{e}_j$ and
where $(f|_k\left(\begin{smallmatrix} a&b\\ c&d\end{smallmatrix}\right))(z) := (cz+d)^{-k} f(\gamma z)$. Then, $\tilde{f}$ is a vector-valued modular form of weight k and the trivial multiplier system with respect to ρ on $\mathrm{SL}_2(\mathbb{Z})$, where ρ is a certain m-dimensional unitary complex representation such that $\rho(\gamma)$ is a permutation matrix for each $\gamma\in \mathrm{SL}_2(\mathbb{Z})$ and is an identity matrix if $\gamma\in\Gamma$. Then, the map $f\mapsto \tilde{f}$ induces an isomorphism between $S_k(\Gamma)$ and $S_{k,\rho}$, where $S_{k,\rho}$ denotes the space of vector-valued cusp forms of weight k and trivial multiplier system with respect to ρ on $\mathrm{SL}_2(\mathbb{Z})$.
For $\tilde{f}, \tilde{g}\in S_{k,\rho}$, we define a Petersson inner product by
Note that if $f, g\in S_{k}(\Gamma)$ such that f and g are orthogonal, then $\tilde{f}$ and $\tilde{g}$ is also orthogonal.
Corollary 4.1. Let k be a positive even integer with k > 2. Let N and n be positive integers and $\Gamma = \Gamma_0(N)$. Let $t_0\in\mathbb{R}, \epsilon \gt 0$. Then, there exists a constant $C(t_0, \epsilon, n) \gt 0$ such that for $k \gt C(t_0, \epsilon,n )$, there exists a basis element $f\in S_{k}(\Gamma)$ satisfying
at any point $s=\sigma + it_0$, with
5. The case of Jacobi forms
We now consider the case of Jacobi forms. Let k be a positive even integer and m be a positive integer. From now, we use the notation $\tau = u+iv\in\mathbb{H}$ and $z = x+iy\in\mathbb{C}$. We review basic notions of Jacobi forms (for more details, see [Reference Choi and Lim9, Section 3.1] and [Reference Eichler and Zagier10, Section 5]). Let F be a complex-valued function on $\mathbb{H}\times \mathbb{C}$. For $\gamma=\left(\begin{smallmatrix} a&b\\ c&d\end{smallmatrix}\right) \in\mathrm{SL}_2(\mathbb{Z}) , X = (\lambda, \mu)\in \mathbb{Z}^2$, we define
and
where $\gamma(\tau,z) = (\frac{a\tau+b}{c\tau+d},\frac{z}{c\tau+d})$.
We now give the definition of a Jacobi form.
Definition 5.1. A Jacobi form of weight k and index m on $\mathrm{SL}_2(\mathbb{Z})$ is a holomorphic function F on $\mathbb{H}\times\mathbb{C}$, satisfying
(1) $F|_{k,m} \gamma =F$ for every $\gamma\in\mathrm{SL}_2(\mathbb{Z})$,
(2) $F|_m X = F$ for every $X\in \mathbb{Z}^2$,
(3) F has the Fourier expansion of the form
(5.1)\begin{equation} F(\tau,z) = \sum_{\substack{l, r\in\mathbb{Z}\\ 4ml - r^2 \geq0}}a(l,r)\text{e}^{2\pi il\tau}\text{e}^{2\pi irz}. \end{equation}
We denote by $J_{k,m}$ the space of all Jacobi forms of weight k and index m on $\mathrm{SL}_2(\mathbb{Z})$. If a Jacobi form satisfies the condition $a(l,r)\neq 0$ only if $4ml - r^2 \gt 0$, then it is called a Jacobi cusp form. We denote by $S_{k,m}$ the space of all Jacobi cusp forms of weight k and index m on $\mathrm{SL}_2(\mathbb{Z})$.
Let F be a Jacobi cusp form $F\in S_{k,m}$ with its Fourier expansion equation (5.1). We define the partial L-functions of F by
for $1\leq j\leq 2m$. This L-function was studied in [Reference Berndt3, Reference Choi and Lim8]. Moreover, F can be written as
with uniquely determined holomorphic functions $F_j:\mathbb{H}\to\mathbb{C}$ and functions in $\{F_j|\ 1\leq j\leq 2m\}$ have the Fourier expansions
where the theta series $\theta_{m, j}$ is defined by
for $1\leq j\leq 2m$.
We write $\mathrm{Mp}_2(\mathbb{R})$ for the metaplectic group. The elements of $\mathrm{Mp}_2(\mathbb{R})$ are pairs $(\gamma,\phi(\tau))$, where $\gamma = \left(\begin{smallmatrix} a&b\\ c&d\end{smallmatrix}\right)\in\mathrm{SL}_2(\mathbb{R})$ and ϕ denotes a holomorphic function on $\mathbb{H}$, with $\phi(\tau)^2= c\tau+d$. Throughout this paper, following [Reference Shimura21], we use the convention that $\sqrt{\tau}$ is chosen so that $\arg(\sqrt{\tau})\in (-\pi/2, \pi/2]$. The map
defines a locally isomorphic embedding of $\mathrm{SL}_2(\mathbb{R})$ into $\mathrm{Mp}_2(\mathbb{R})$. Let $\mathrm{Mp}_2(\mathbb{Z})$ be the inverse image of $\mathrm{SL}_2(\mathbb{Z})$ under the covering map $\mathrm{Mp}_2(\mathbb{R})\to\mathrm{SL}_2(\mathbb{R})$. It is well-known that $\mathrm{Mp}_2(\mathbb{Z})$ is generated by $\widetilde{T}$ and $\widetilde{S}$, where $\widetilde{T}$ and $\widetilde{S}$ are the lifts of the standard generators T and S of $\mathrm{SL}_2(\mathbb{Z})$, respectively. We define a 2m-dimensional unitary complex representation $\widetilde{\rho}_m$ of $\mathrm{Mp}_2(\mathbb{Z})$ by
and
Let χ be a multiplier system of weight $\frac12$ on $\mathrm{SL}_2(\mathbb{Z})$. We define a map $\rho_m : \mathrm{SL}_2(\mathbb{Z}) \to \mathrm{GL}_{2m}(\mathbb{C})$ by
for $\gamma\in\mathrm{SL}_2(\mathbb{Z})$. The map ρm gives a 2m-dimensional unitary representation of $\mathrm{SL}_2(\mathbb{Z})$.
Let $\{\mathbf{e}_1, \ldots, \mathbf{e}_{2m}\}$ denote the standard basis of $\mathbb{C}^{2m}$. For $F\in S_{k,m}$, we define a vector-valued function $\tilde{F}:\mathbb{H}\to \mathbb{C}^{2m}$ by $\tilde{F} = \sum_{j=1}^{2m} F_j\mathbf{e}_j$, where Fj is defined by the theta expansion in equation (5.2). Then, the map $F\mapsto \tilde{F}$ induces an isomorphism between $S_{k,m}$ and $S_{k-\frac12, \bar{\chi}, \rho_m}$.
Let $L^*(F,j,s) := \frac{\Gamma(s)}{(2\pi)^s} L(F,j,s)$. Then, we have the following corollary.
Corollary 5.2. Let k be a positive even integer with k > 2. Let m and n be positive integers. Let $t_0\in\mathbb{R}$ and ϵ > 0.
(1) Let j be a positive integer with $1\leq j\leq 2m$. Then, there exists a constant $C(t_0, \epsilon,j,n) \gt 0$ such that for any $k \gt C(t_0, \epsilon, j,n)$ and any $s=\sigma + it_0$ with
\begin{equation*}\frac{2k-3}{4} \lt \sigma \lt \frac {2k-1}{4} - \epsilon,\end{equation*}there exists a Jacobi cusp form $F\in S_{k,m}$ such that
\begin{align*} \frac{d^n}{ds^n} L^*(F,j,s)\neq 0. \end{align*}(2) There exists a constant $C(t_0, \epsilon,n) \gt 0$ such that for any $k \gt C(t_0, \epsilon,n)$ and any $s=\sigma + it_0$ with
\begin{equation*}\frac{2k-3}{4} \lt \sigma \lt \frac {2k-1}{4} - \epsilon \ \ and \ \ \frac {2k-1}{4} + \epsilon \lt \sigma \lt \frac{2k+1}4,\end{equation*}there exist a Jacobi cusp form $F\in S_{k,m}$ and $j\in\{1,\ldots, 2m\}$ such that
\begin{align*} \frac{d^n}{ds^n} L^*(F,j,s) \neq 0. \end{align*}
Remark 5.3. Note that $\rho_m(-I)$ is not equal to the identity matrix in $\mathrm{GL}_{2m}(\mathbb{C})$. Instead, we have
By a similar argument, we prove the same result as in Theorem 3.1 for the representation ρm.
6. The case of Kohnen plus space
Let k be a positive even integer. By [Reference Eichler and Zagier10, Theorem 5.4], there is an isomorphism ϕ between $S_{k,1}$ and $S^+_{k-\frac12}$, where $S^+_{k-\frac12}$ denotes the space of cusp forms in the plus space of weight $k-\frac12$ on $\Gamma_0(4)$. Moreover, this isomorphism is compatible with the Petersson scalar products.
Let f be a cusp form in $S^+_{k-\frac12}$ with Fourier expansion $f(\tau) = \sum_{\substack{n \gt 0\\ n\equiv 0,3 \pmod{4}}} c(n)\text{e}^{2\pi in\tau}$. Then, the L-function of f is defined by
For $1\leq j\leq 2$, let cj be defined by
Then, $c(n) = c_1(n) + c_2(n)$ for all n. With this, we consider partial sums of $L(f,s)$ by
for $1\leq j\leq 2$.
Suppose that F is a Jacobi cusp form in $S_{k,1}$. By the theta expansion in equation (5.2), we have a corresponding vector-valued modular form $(F_1(\tau), F_2(\tau))$. Then, the isomorphism ϕ from $S_{k,1}$ to $S^+_{k-\frac12}$ is given by
From this, we see that
We have the following corollary regarding the partial sums of $L(f,s)$ for $f\in S^+_{k-\frac12}$.
Corollary 6.1. Let k be a positive even integer with k > 2. Let n be a positive integer. Let $t_0\in\mathbb{R}$ and ϵ > 0.
(1) Let j be a positive integer with $1\leq j\leq 2$. Then, there exists a constant $C(t_0, \epsilon,j,n) \gt 0$ such that for any $k \gt C(t_0, \epsilon,j,n)$ and any $s=\sigma + it_0$ with
\begin{equation*}\frac{2k-3}{4} \lt \sigma \lt \frac {2k-1}{4} - \epsilon,\end{equation*}there exists a cusp form $f\in S^+_{k-\frac12}$ such that
\begin{align*} \frac{d^n}{ds^n} \left[ 4^s L^*(f,j,s) \right]\neq 0. \end{align*}(2) There exists a constant $C(t_0, \epsilon,n) \gt 0$ such that for any $k \gt C(t_0, \epsilon,n)$ and any $s=\sigma + it_0$, with
\begin{equation*}\frac{2k-3}{4} \lt \sigma \lt \frac {2k-1}{4} - \epsilon \ \ and \ \ \frac {2k-1}{4} + \epsilon \lt \sigma \lt \frac{2k+1}4,\end{equation*}there exist a cusp form $f\in S^+_{k-\frac12}$ and $j\in\{1,\ldots, 2\}$ such that
\begin{align*} \frac{d^n}{ds^n} \left[ 4^s L^*(f,j,s) \right] \neq 0. \end{align*}
Funding Statement
The first author was supported by the National Research Foundation of Korea grant funded by the Ministry of Science and ICT (MSIT)(No. RS-2024-00346031).