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

On single-variable Witten zeta functions of rank 2 root systems

Kam Cheong Au University of Cologne
Department of Mathematics and Computer Science
Weyertal 86-90, 50931 Cologne, Germany
[email protected]
Abstract.

We use a Mellin-transform-based method to simultaneously analyse single-variable Witten zeta functions of all rank 22 irreducible root systems. We obtain satisfactory details regarding their pole locations, values of residues and special values. We also mention a connection between Witten zeta function to Eisenstein series.

Key words and phrases:
Witten zeta functions, analytic continuation, residues, asymptotic formula
2020 Mathematics Subject Classification:
Primary: 11M32, 11M35. Secondary: 11P82

Introduction

Let 𝔀\mathfrak{g} be a simple Lie algebra over β„‚\mathbb{C},

΢𝔀​(s):=βˆ‘Ο1(dim ​ρ)s,ρ∈{finite-dimensional irreducible representations of 𝔀},\zeta_{\mathfrak{g}}(s):=\sum_{\rho}\frac{1}{(\text{dim }\rho)^{s}},\quad\rho\in\{\text{finite-dimensional irreducible representations of $\mathfrak{g}$}\},

be its Witten zeta function. If Ξ¦\Phi is the root system associated to 𝔀\mathfrak{g}, we abuse notation by writing ΢𝔀​(s)\zeta_{\mathfrak{g}}(s) as ΢Φ​(s)\zeta_{\Phi}(s).

There are vast literatures on Witten zeta function ([12], [14], [17], [10], [11], [9], [16]), especially on functional relations of their multi-variable generalizations. Finer analytic properties of single-variable ΢Φ​(s)\zeta_{\Phi}(s), however, received less attention and seem not easily derivable from their multi-variable versions. In this article, we look at following properties of ΢Φ​(s),Ξ¦=A2,B2,G2\zeta_{\Phi}(s),\Phi=A_{2},B_{2},G_{2}:

  • β€’

    values of residue at abscissa of convergence;

  • β€’

    values at 0 and negative integers;

  • β€’

    derivative at zero;

  • β€’

    location of all poles.

Answers to them are nice yet non-trivial, this leads one to suspect ΢Φ​(s)\zeta_{\Phi}(s) underlies deep analytic and arithmetic properties that were up-to-now overlooked. We denote

Ο‰A​(s):=2βˆ’s​΢A2​(s)\displaystyle\omega_{A}(s):=2^{-s}\zeta_{A_{2}}(s) =βˆ‘n,mβ‰₯11ms​ns​(m+n)s;\displaystyle=\sum_{n,m\geq 1}\frac{1}{m^{s}n^{s}(m+n)^{s}};
Ο‰B​(s):=6βˆ’s​΢B2​(s)\displaystyle\omega_{B}(s):=6^{-s}\zeta_{B_{2}}(s) =βˆ‘n,mβ‰₯11ms​ns​(m+n)s​(m+2​n)s;\displaystyle=\sum_{n,m\geq 1}\frac{1}{m^{s}n^{s}(m+n)^{s}(m+2n)^{s}};
Ο‰G​(s):=120βˆ’s​΢G2​(s)\displaystyle\omega_{G}(s):=120^{-s}\zeta_{G_{2}}(s) =βˆ‘n,mβ‰₯11ms​ns​(m+n)s​(m+2​n)s​(m+3​n)s​(2​m+3​n)s.\displaystyle=\sum_{n,m\geq 1}\frac{1}{m^{s}n^{s}(m+n)^{s}(m+2n)^{s}(m+3n)^{s}(2m+3n)^{s}}.

Major properties of these functions are tabulated below.

Ο‰βˆ™β€‹(s)\omega_{\bullet}(s) AA BB GG
Converges when β„œβ‘(s)>23\Re(s)>\frac{2}{3} β„œβ‘(s)>12\Re(s)>\frac{1}{2} β„œβ‘(s)>13\Re(s)>\frac{1}{3}
Residue at abscissa
of convergence
Γ​(13)32​3​π\dfrac{\Gamma(\frac{1}{3})^{3}}{2\sqrt{3}\pi} Γ​(14)28​2​π\dfrac{\Gamma(\frac{1}{4})^{2}}{8\sqrt{2\pi}} Γ​(13)328/3​33/2​π\dfrac{\Gamma(\frac{1}{3})^{3}}{2^{8/3}3^{3/2}\pi}
ω​(0)\omega(0) 13\frac{1}{3} 38\frac{3}{8} 512\frac{5}{12}
ω​(βˆ’n)\omega(-n) Vanishes for nβˆˆβ„•n\in\mathbb{N} Vanishes for n∈2​ℕn\in 2\mathbb{N}
ω′​(0)\omega^{\prime}(0) log⁑2+log⁑π\log 2+\log\pi 54​log⁑2+32​log⁑π\frac{5}{4}\log 2+\frac{3}{2}\log\pi 2​log⁑2βˆ’12​log⁑3+52​log⁑π2\log 2-\frac{1}{2}\log 3+\frac{5}{2}\log\pi
Location of
other poles
s=k2s=\frac{k}{2}
k≀1,k≑1(mod2)k\leq 1,k\equiv 1\pmod{2}
s=k3s=\frac{k}{3}
k≀1,k≑1,5(mod6)k\leq 1,k\equiv 1,5\pmod{6}
s=k5s=\frac{k}{5}
k≀1,k≑1,3,7,9(mod10)k\leq 1,k\equiv 1,3,7,9\pmod{10}
Table 1. Major properties of Ο‰A​(s),Ο‰B​(s)\omega_{A}(s),\omega_{B}(s) and Ο‰G​(s)\omega_{G}(s).

All facts concerning Ο‰A​(s)\omega_{A}(s) proved in this article are known, thanks to a plethora of literatures on Tornheim’s double zeta function ([17], [3], [19], [15]), which specializes to Ο‰A​(s)\omega_{A}(s). Main novelty are thus corresponding facts on Ο‰B​(s)\omega_{B}(s) and Ο‰G​(s)\omega_{G}(s), which received much less investigation. Bringmann et. al in [4] proved first three rows for Ο‰B​(s)\omega_{B}(s); Rutard [20] establishes the rows ω​(0),ω′​(0)\omega(0),\omega^{\prime}(0).

This article proves all (and re-proves some) entries of table above. We analyse Ο‰A​(s),Ο‰B​(s),Ο‰G​(s)\omega_{A}(s),\omega_{B}(s),\omega_{G}(s) in a uniform manner, this is done by introducing a more general series

ΞΆf​(s):=βˆ‘n,mβ‰₯1(n​md+1​f​(nm))βˆ’s,\zeta_{f}(s):=\sum_{n,m\geq 1}\left(nm^{d+1}f(\frac{n}{m})\right)^{-s},

where f​(x)f(x) is a degree dd polynomial with real coefficient such that f​(ℝβ‰₯0)βŠ‚β„>0f(\mathbb{R}^{\geq 0})\subset\mathbb{R}^{>0}. We recover Ο‰A​(s),Ο‰B​(s)\omega_{A}(s),\omega_{B}(s) and Ο‰G​(s)\omega_{G}(s) by letting ff to be

(0.1) fA​(x):=1+x,fB​(x):=(1+x)​(1+2​x),fG​(x):=(1+x)​(1+2​x)​(1+3​x)​(2+3​x).f_{A}(x):=1+x,\quad f_{B}(x):=(1+x)(1+2x),\quad f_{G}(x):=(1+x)(1+2x)(1+3x)(2+3x).
Theorem 0.1.

Let ff be a degree dd polynomial with real coefficients such that f​(ℝβ‰₯0)βŠ‚β„>0f(\mathbb{R}^{\geq 0})\subset\mathbb{R}^{>0}, then ΞΆf​(s)\zeta_{f}(s) has meromorphic continuation to β„‚\mathbb{C}, also

  1. (1)

    ΞΆf​(s)\zeta_{f}(s) has abscissa of convergence s=2d+2s=\frac{2}{d+2}, it has a simple pole at this point with residue

    1d+2β€‹βˆ«0∞(x​f​(x))βˆ’2/(d+2)​𝑑x;\frac{1}{d+2}\int_{0}^{\infty}(xf(x))^{-2/(d+2)}dx;
  2. (2)

    ΞΆf​(s)\zeta_{f}(s) is of moderate increase along imaginary direction, that is, for any compact subset KβŠ‚β„K\subset\mathbb{R}, there exists M>0M>0 such that

    |ΞΆf​(Οƒ+i​t)|=O​(|t|M),ΟƒβˆˆK,|t|≫1;|\zeta_{f}(\sigma+it)|=O(|t|^{M}),\qquad\sigma\in K,|t|\gg 1;
  3. (3)

    ΞΆf​(s)\zeta_{f}(s) is analytic at s=0,βˆ’1,βˆ’2,β‹―s=0,-1,-2,\cdots, with an explicit formula to calculate them (Theorem 3.1);

  4. (4)

    Apart from the pole at s=2d+2s=\frac{2}{d+2}, all other poles are at s∈11+d​℀≀1s\in\frac{1}{1+d}\mathbb{Z}^{\leq 1}, they are at most simple (Theorem 4.1);

  5. (5)

    If roots of ff are rational numbers, then ΞΆf′​(0)\zeta^{\prime}_{f}(0) is a linear combination of Ξ³\gamma (Euler’s constant), ΢′​(βˆ’1)\zeta^{\prime}(-1) and log⁑Γ​(r)\log\Gamma(r) for various rβˆˆβ„šr\in\mathbb{Q}.

We will see Ο‰βˆ™β€‹(s)\omega_{\bullet}(s) manifest unusual properties not found in generic ΞΆf​(s)\zeta_{f}(s). Here we give two examples: (1) ΞΆf​(βˆ’2​n)β‰ 0\zeta_{f}(-2n)\neq 0 for generic ff, whereas Ο‰βˆ™β€‹(βˆ’2​n)=0\omega_{\bullet}(-2n)=0; (2) for generic ff of degree 44, ΞΆf​(s)\zeta_{f}(s) has poles at s=k/5s=k/5 for all k≀1k\leq 1 such that k/5βˆ‰β„€k/5\notin\mathbb{Z}, whereas for Ο‰G​(s)\omega_{G}(s), only those kk which are odd give poles. These properties indicate Witten zeta functions are nice objects to study and provide motivation to investigate the corresponding situation at higher rank.

We structure the article as follows: In Β§1, we develop prerequisite materials on our Mellin transform setup, which is needed in Β§2 for meromorphic continuation of ΞΆf​(s)\zeta_{f}(s). Β§3 looks at values of ΞΆf\zeta_{f} at non-positive integers. Β§4 classifies all poles of ΞΆf\zeta_{f}. Β§5 calculates ΞΆf′​(0)\zeta_{f}^{\prime}(0). Β§6 applies our previously gained knowledge on ΞΆG2​(s)\zeta_{G_{2}}(s) to derive an asymptotic formula for number of representation of Lie algebra 𝔀2\mathfrak{g}_{2}.

Acknowledgment

The author thanks Prof. Kathrin Bringmann valuable discussions and feedbacks. The author has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 101001179).

1. An integration kernel

In this section, we assemble materials that we will need for our investigation of ΞΆf​(s)\zeta_{f}(s). Let f​(x)f(x) be a degree dd polynomial with real coefficient such that f​(ℝβ‰₯0)βŠ‚β„>0f(\mathbb{R}^{\geq 0})\subset\mathbb{R}^{>0}, define

Ff​(a,s)=∫0∞(f​(x))βˆ’a​xs​d​xx,dβ€‹β„œβ‘(a)>β„œβ‘(s)>0.F_{f}(a,s)=\int_{0}^{\infty}(f(x))^{-a}x^{s}\frac{dx}{x},\qquad d\Re(a)>\Re(s)>0.

It is absolutely convergent on the region dβ€‹β„œβ‘(a)>β„œβ‘(s)>0d\Re(a)>\Re(s)>0, so defines an analytic function there.

Theorem 1.1.

The function Ff​(a,s)F_{f}(a,s) has meromorphic continuation to β„‚2\mathbb{C}^{2}, and

Kf​(a,s)=Γ​(a)Γ​(s)​Γ​(d​aβˆ’s)​Ff​(a,s)K_{f}(a,s)=\frac{\Gamma(a)}{\Gamma(s)\Gamma(da-s)}F_{f}(a,s)

is an entire function.

Proof.

Let CC be a contour on complex plane that starts at +∞+\infty with small negative imaginary part, goes around 0, wraps around positive real axis and ends at +∞+\infty with small positive imaginary part.

[Uncaptioned image]

Then, by agreeing 0≀arg⁑(x)<2​π0\leq\arg(x)<2\pi, we have

(1.1) Ff​(a,s)=1e2​π​i​sβˆ’1β€‹βˆ«Cf​(x)βˆ’a​xs​d​xxF_{f}(a,s)=\frac{1}{e^{2\pi is}-1}\int_{C}f(x)^{-a}x^{s}\frac{dx}{x}

now since the path of integral no longer passes through x=0x=0, only rapid decrease at x=∞x=\infty (i.e. dβ€‹β„œβ‘(a)>β„œβ‘(s)d\Re(a)>\Re(s)) is needed to ensure ∫Cf​(x)βˆ’a​xs​d​xx\int_{C}f(x)^{-a}x^{s}\frac{dx}{x} being analytic, thus we obtain a continuation on a larger region with the condition β„œβ‘(s)>0\Re(s)>0 removed. From the poles of 1/(e2​π​i​sβˆ’1)1/(e^{2\pi is}-1), we see that Ff​(a,s)F_{f}(a,s) has simple poles at s=0,βˆ’1,βˆ’2,β‹―s=0,-1,-2,\cdots, they are cancelled by zeroes of 1/Γ​(s)1/\Gamma(s), so

Ff​(a,s)Γ​(s)​ is analytic on ​dβ€‹β„œβ‘(a)>β„œβ‘(s).\frac{F_{f}(a,s)}{\Gamma(s)}\text{ is analytic on }d\Re(a)>\Re(s).

Let g​(x)=xd​f​(1/x)βˆˆβ„β€‹[x]g(x)=x^{d}f(1/x)\in\mathbb{R}[x], then substitution x↦1/xx\mapsto 1/x in the original integral definition of Ff​(a,s)F_{f}(a,s) shows Ff​(a,s)=Fg​(a,a​dβˆ’s)F_{f}(a,s)=F_{g}(a,ad-s). Apply this conclusion to gg:

Fg​(a,a​dβˆ’s)Γ​(a​dβˆ’s)​ is analytic on ​dβ€‹β„œβ‘(a)>β„œβ‘(a​dβˆ’s),\frac{F_{g}(a,ad-s)}{\Gamma(ad-s)}\text{ is analytic on }d\Re(a)>\Re(ad-s),

that is Ff​(a,s)/Γ​(a​dβˆ’s)F_{f}(a,s)/\Gamma(ad-s) is analytic on β„œβ‘(s)>0\Re(s)>0. From integral representation (1.1) again: only poles on β„œβ‘(s)≀0\Re(s)\leq 0 are simple poles at s=0,βˆ’1,βˆ’2,β‹―s=0,-1,-2,\cdots, so Ff​(a,s)Γ​(s)​Γ​(a​dβˆ’s)\frac{F_{f}(a,s)}{\Gamma(s)\Gamma(ad-s)} is an entire function in aa and ss.

To prove Kf​(a,s)K_{f}(a,s) is still entire, we need to show Ff​(βˆ’n,s)=0F_{f}(-n,s)=0 when n=0,1,2,β‹―n=0,1,2,\cdots. By analytic continuation, we can assume ss is sufficiently negative. By representation (1.1), we need to prove

∫Cf​(x)n​xs​d​xx=0\int_{C}f(x)^{n}x^{s}\frac{dx}{x}=0

Let C​(R)C(R) be truncation of CC at large real number RR, let C′​(R)C^{\prime}(R) be a circle of large radius RR as in figure.

[Uncaptioned image]

Since nβ‰₯0n\geq 0, f​(x)n​xs​1xf(x)^{n}x^{s}\frac{1}{x} is analytic inside the contour, so

∫C​(R)f​(x)n​xs​d​xx+∫C′​(R)f​(x)n​xs​d​xx=0\int_{C(R)}f(x)^{n}x^{s}\frac{dx}{x}+\int_{C^{\prime}(R)}f(x)^{n}x^{s}\frac{dx}{x}=0

and when ss is sufficiently negative, the integral along C′​(R)C^{\prime}(R) tends to 0 as Rβ†’βˆžR\to\infty, so integral along CC is indeed zero. ∎

Remark 1.2.

For d=1,2d=1,2, Ff​(a,s)F_{f}(a,s) and hence Kf​(a,s)K_{f}(a,s) can be expressed in terms of well-known special functions:

(1.2) Kf​(a,s)\displaystyle K_{f}(a,s) =Ξ±1βˆ’s\displaystyle=\alpha_{1}^{-s} f​(x)=1+Ξ±1​x\displaystyle f(x)=1+\alpha_{1}x
(1.3) Kf​(a,s)\displaystyle K_{f}(a,s) =Γ​(a)Γ​(2​a)​α1βˆ’s​F12​[.a,s2​a.;1βˆ’Ξ±2Ξ±1]\displaystyle=\frac{\Gamma(a)}{\Gamma(2a)}\alpha_{1}^{-s}{}_{2}F_{1}{\left[\genfrac{.}{.}{0.0pt}{}{a{\mathchar 44\relax}\mskip 8.0mus}{2a};1-\frac{\alpha_{2}}{\alpha_{1}}\right]}\qquad\qquad f​(x)=(1+Ξ±1​x)​(1+Ξ±2​x)\displaystyle f(x)=(1+\alpha_{1}x)(1+\alpha_{2}x)

with F12{}_{2}F_{1} being Gauss hypergeometric function, they follow from the general formula:

∫0∞xsβˆ’1(1+α​x)a​𝑑x=Ξ±βˆ’s​Γ​(s)​Γ​(aβˆ’s)Γ​(a)\int_{0}^{\infty}\frac{x^{s-1}}{(1+\alpha x)^{a}}dx=\alpha^{-s}\frac{\Gamma(s)\Gamma(a-s)}{\Gamma(a)}
(1.4) ∫0∞xsβˆ’1​(1+x)βˆ’a​(1+λ​x)βˆ’b​𝑑x=Γ​(s)​Γ​(a+bβˆ’s)Γ​(a+b)Γ—F12​[.b,sa+b.;1βˆ’Ξ»]\int_{0}^{\infty}x^{s-1}(1+x)^{-a}(1+\lambda x)^{-b}dx=\frac{\Gamma(s)\Gamma(a+b-s)}{\Gamma(a+b)}\times{}_{2}F_{1}{\left[\genfrac{.}{.}{0.0pt}{}{b{\mathchar 44\relax}\mskip 8.0mus}{a+b};1-\lambda\right]}

For higher dd, an explicit formula is in general not possible. Nonetheless, most of the properties of Kf​(a,s)K_{f}(a,s) we need can be derived without relying on an explicit representation using special functions.

Proposition 1.3.

When nn is a non-negative integer,

Kf​(a,βˆ’n)=(βˆ’1)n​n!​Γ​(a)Γ​(a​d+n)​[f​(x)βˆ’a]​[xn]K_{f}(a,-n)=(-1)^{n}n!\frac{\Gamma(a)}{\Gamma(ad+n)}[f(x)^{-a}][x^{n}]

where [f​(x)βˆ’a]​[xn][f(x)^{-a}][x^{n}] is the coefficient of xnx^{n} in the series expansion of f​(x)βˆ’af(x)^{-a} around x=0x=0.

Proof.

By analytic continuation, we only need to prove this when aa is sufficiently large. In this case

Kf​(a,s)=1e2​π​i​sβˆ’1​Γ​(a)Γ​(s)​Γ​(d​aβˆ’s)β€‹βˆ«Cf​(x)βˆ’a​xs​d​xxK_{f}(a,s)=\frac{1}{e^{2\pi is}-1}\frac{\Gamma(a)}{\Gamma(s)\Gamma(da-s)}\int_{C}f(x)^{-a}x^{s}\frac{dx}{x}

here CC is the contour used in proof of previous theorem. When ss tends to βˆ’n-n, we have

Kf​(a,βˆ’n)=(βˆ’1)n​n!​Γ​(a)Γ​(d​a+n)​(12​π​i)β€‹βˆ«Cf​(x)βˆ’a​xβˆ’n​d​xxK_{f}(a,-n)=\frac{(-1)^{n}n!\Gamma(a)}{\Gamma(da+n)}(\frac{1}{2\pi i})\int_{C}f(x)^{-a}x^{-n}\frac{dx}{x}

Now the integrand remains single-valued across the positive real axis, the only pole is at x=0x=0, its residue there is [f​(x)βˆ’a]​[xn][f(x)^{-a}][x^{n}]. ∎

Proposition 1.4.

When n,nn,n are non-negative integers,

Kf​(βˆ’m,βˆ’n)={(βˆ’1)(d+1)​m​d​n!​(d​mβˆ’n)!m!​[f​(x)m]​[xn]n≀d​m(βˆ’1)n+m​n!m!​(nβˆ’d​mβˆ’1)!​[βˆ’f​(x)m​log⁑f​(x)]​[xn]n>d​m.K_{f}(-m,-n)=\begin{cases}(-1)^{(d+1)m}d\frac{n!(dm-n)!}{m!}[f(x)^{m}][x^{n}]&\quad n\leq dm\\ (-1)^{n+m}\frac{n!}{m!(n-dm-1)!}[-f(x)^{m}\log f(x)][x^{n}]&\quad n>dm\end{cases}.

In particular, if f​(x)βˆˆβ„šβ€‹[x]f(x)\in\mathbb{Q}[x], then

Kf​(βˆ’n,m)βˆˆβ„šnβˆˆβ„€β‰₯0,mβˆˆβ„€.K_{f}(-n,m)\in\mathbb{Q}\qquad n\in\mathbb{Z}^{\geq 0},m\in\mathbb{Z}.
Proof.

When n≀d​mn\leq dm, we simply let aβ†’βˆ’ma\to-m, here we note that both Γ​(a)\Gamma(a) and Γ​(a​d+n)\Gamma(ad+n) has simple pole at a=βˆ’ma=-m, this gives the formula in the case n≀d​mn\leq dm. When n>d​mn>dm, Γ​(a)\Gamma(a) has a pole at a=βˆ’ma=-m, Γ​(a​d+n)\Gamma(ad+n) is regular with value (d​mβˆ’nβˆ’1)!(dm-n-1)!, and [f​(x)βˆ’a]​[xn][f(x)^{-a}][x^{n}] has a zero at a=βˆ’ma=-m,111because f​(x)mf(x)^{m} is a polynomial of degree d​mdm and we are extracting coefficient higher than the degree.. By differentiating with respect to aa, we obtain its first order term:

[f​(x)βˆ’a]​[xn]=[βˆ’f​(x)m​log⁑f​(x)]​(a+m)+O​((a+m)2)aβ†’βˆ’m[f(x)^{-a}][x^{n}]=[-f(x)^{m}\log f(x)](a+m)+O((a+m)^{2})\qquad a\to-m

therefore the formula in the case n>d​mn>dm.

Regarding the last claim of rationality, this is evident from the formula when m≀0m\leq 0. For m>0m>0, let g​(x)=xd​f​(1/x)g(x)=x^{d}f(1/x), as in the proof of previous theorem, Kf​(a,s)=Kg​(a,a​dβˆ’s)K_{f}(a,s)=K_{g}(a,ad-s), thus Kf​(βˆ’n,m)=Kg​(βˆ’n,βˆ’n​dβˆ’m)K_{f}(-n,m)=K_{g}(-n,-nd-m) with βˆ’n​dβˆ’m<0-nd-m<0, so we can apply the formula again with ff replaced by gg and produces the same conclusion. ∎

Proposition 1.5.

Let f​(x)=c0​(1+Ξ±1​x)​⋯​(1+Ξ±d​x)f(x)=c_{0}(1+\alpha_{1}x)\cdots(1+\alpha_{d}x),222note that our restriction on ff implies c0>0c_{0}>0 and each Ξ±iβˆˆβ„‚βˆ’(βˆ’βˆž,0]\alpha_{i}\in\mathbb{C}-(-\infty,0], then

Kf​(0,s)=βˆ‘i=1dΞ±iβˆ’sK_{f}(0,s)=\sum_{i=1}^{d}\alpha_{i}^{-s}
Proof.

By analytic continuation, it suffices to prove this when ss is sufficiently negative, so the following representation is valid:

Kf​(a,s)=1e2​π​i​sβˆ’1​Γ​(a)Γ​(s)​Γ​(a​dβˆ’s)β€‹βˆ«Cf​(x)βˆ’a​xs​d​xxK_{f}(a,s)=\frac{1}{e^{2\pi is}-1}\frac{\Gamma(a)}{\Gamma(s)\Gamma(ad-s)}\int_{C}f(x)^{-a}x^{s}\frac{dx}{x}

with CC same contour as in proof of previous theorem. Letting a→0a\to 0 gives

Kf​(0,s)=βˆ’1e2​π​i​sβˆ’1​1Γ​(s)​Γ​(βˆ’s)β€‹βˆ«Cxs​log⁑f​(x)​d​xx:=h​(s)β€‹βˆ«Cxs​log⁑f​(x)​d​xxK_{f}(0,s)=-\frac{1}{e^{2\pi is}-1}\frac{1}{\Gamma(s)\Gamma(-s)}\int_{C}x^{s}\log f(x)\frac{dx}{x}:=h(s)\int_{C}x^{s}\log f(x)\frac{dx}{x}

We know from remark above that Kf​(0,s)=Ξ±βˆ’sK_{f}(0,s)=\alpha^{-s} when f=1+α​xf=1+\alpha x, so

∫Cxs​log⁑(1+α​x)​d​xx=Ξ±βˆ’s​h​(s)βˆ’1\int_{C}x^{s}\log(1+\alpha x)\frac{dx}{x}=\alpha^{-s}h(s)^{-1}

So when f​(x)=c0​(1+Ξ±1​x)​⋯​(1+Ξ±d​x)f(x)=c_{0}(1+\alpha_{1}x)\cdots(1+\alpha_{d}x),

Kf​(0,s)=h​(s)β€‹βˆ«Cxs​log⁑[c​(1+Ξ±1​x)​⋯​(1+Ξ±d​x)]​d​xx=h​(s)​(log⁑c0)β€‹βˆ«Cxs​d​xx⏟=0+βˆ‘i=1dh​(s)β€‹βˆ«Cxs​log⁑(1+Ξ±i​x)​d​xx=h​(s)β€‹βˆ‘i=1dh​(s)βˆ’1​αiβˆ’s.\begin{aligned} K_{f}(0,s)&=h(s)\int_{C}x^{s}\log[c(1+\alpha_{1}x)\cdots(1+\alpha_{d}x)]\frac{dx}{x}\\ &=h(s)(\log c_{0})\underbrace{\int_{C}x^{s}\frac{dx}{x}}_{=0}+\sum_{i=1}^{d}h(s)\int_{C}x^{s}\log(1+\alpha_{i}x)\frac{dx}{x}\\ &=h(s)\sum_{i=1}^{d}h(s)^{-1}\alpha_{i}^{-s}\end{aligned}.

as desired. ∎

Proposition 1.6.

Write

f​(x)=c0​(1+Ξ±1​x)​⋯​(1+Ξ±d​x)=c0+c1​x+β‹―+cd​xd,f(x)=c_{0}(1+\alpha_{1}x)\cdots(1+\alpha_{d}x)=c_{0}+c_{1}x+\cdots+c_{d}x^{d},

we have

Kf​(s,1βˆ’s)=cdβˆ’1cd+s​(cdβˆ’1cd​((dβˆ’1)β€‹Ξ³βˆ’log⁑cd)+(d+1)β€‹βˆ‘ilog⁑αiΞ±i)+O​(s2),sβ†’0K_{f}(s,1-s)=\frac{c_{d-1}}{c_{d}}+s\left(\frac{c_{d-1}}{c_{d}}((d-1)\gamma-\log c_{d})+(d+1)\sum_{i}\frac{\log\alpha_{i}}{\alpha_{i}}\right)+O(s^{2}),\qquad s\to 0

where Ξ³\gamma is Euler’s constant.

We will need this result during evaluation of ΞΆf′​(0)\zeta^{\prime}_{f}(0).

Proof.

The constant term is Kf​(0,1)=βˆ‘iΞ±βˆ’1=cdβˆ’1/cdK_{f}(0,1)=\sum_{i}\alpha^{-1}=c_{d-1}/c_{d}, coefficient of ss is

dd​s|s=0​Kf​(s,1βˆ’s)=dd​s|s=0​Kf​(s,1)βˆ’dd​s|s=1​Kf​(0,s)\left.\frac{d}{ds}\right|_{s=0}K_{f}(s,1-s)=\left.\frac{d}{ds}\right|_{s=0}K_{f}(s,1)-\left.\frac{d}{ds}\right|_{s=1}K_{f}(0,s)

where chain rule has been used. The second term is easy to evaluate by the previous proposition: dd​s|s=1​Kf​(0,s)=βˆ’βˆ‘i(log⁑αi)/Ξ±i\left.\frac{d}{ds}\right|_{s=1}K_{f}(0,s)=-\sum_{i}(\log\alpha_{i})/\alpha_{i}. For first term, let g​(x)=xd​f​(x)g(x)=x^{d}f(x),

dd​s|s=0​Kf​(s,1)=dd​s|s=0​Kg​(s,s​dβˆ’1)=dd​s|s=0​Kg​(s,βˆ’1)+d​dd​s|s=βˆ’1​Kg​(0,s)\left.\frac{d}{ds}\right|_{s=0}K_{f}(s,1)=\left.\frac{d}{ds}\right|_{s=0}K_{g}(s,sd-1)=\left.\frac{d}{ds}\right|_{s=0}K_{g}(s,-1)+d\left.\frac{d}{ds}\right|_{s=-1}K_{g}(0,s)

Since Kg​(0,s)=βˆ‘iΞ±sK_{g}(0,s)=\sum_{i}\alpha^{s}, the second term derivative on RHS equals βˆ‘i(log⁑αi)/Ξ±i\sum_{i}(\log\alpha_{i})/\alpha_{i}. For first term, by Proposition 1.3,

Kg​(s,βˆ’1)=βˆ’Ξ“β€‹(s)Γ​(s​d+1)​[f​(x)βˆ’s]​[x1]=Γ​(s+1)Γ​(s​d+1)​cdβˆ’sβˆ’1​cdβˆ’1K_{g}(s,-1)=-\frac{\Gamma(s)}{\Gamma(sd+1)}[f(x)^{-s}][x^{1}]=\frac{\Gamma(s+1)}{\Gamma(sd+1)}c_{d}^{-s-1}c_{d-1}

Evaluating derivatives of the gamma functions gives the claim. ∎

Following two technical lemmas concern growth of Ff​(a,s)F_{f}(a,s) when arguments a,sa,s have large imaginary part.

Lemma 1.7.

(a) There exists a positive integer NN and polynomials p0,β‹―,pNβˆˆβ„β€‹[a,s]p_{0},\cdots,p_{N}\in\mathbb{R}[a,s] such that Ff​(a,s)F_{f}(a,s) satisfies a recurrence of form

βˆ‘i=0Npi​(a,s)​Ff​(a,s+i)=0\sum_{i=0}^{N}p_{i}(a,s)F_{f}(a,s+i)=0

(b) Denote s=Οƒ+i​ts=\sigma+it to be its real and imaginary part. There exists a positive number Ξ΄>0\delta>0, depending only on ff, such that for any Ξ΅>0\varepsilon>0,

Ff​(a,Οƒ+i​t)=OK,a​(eβˆ’(Ξ΄βˆ’Ξ΅)​|t|)|t|>1,ΟƒβˆˆK​ a compact subset of ​ℝF_{f}(a,\sigma+it)=O_{K,a}(e^{-(\delta-\varepsilon)|t|})\qquad|t|>1,\sigma\in K\textit{ a compact subset of }\mathbb{R}

We can choose Ξ΄=Ο€\delta=\pi if all roots of ff are real.

Proof.

(a) This is a direct consequence of creative telescoping, which guarantees existence of pip_{i} and a rational function q​(a,s,x)βˆˆβ„β€‹(a,s,x)q(a,s,x)\in\mathbb{R}(a,s,x) such that

βˆ‘i=0Npi​(a,s)​(f​(x))βˆ’a​xs+i​1x=dd​x​(q​(a,s,x)​(f​(x))βˆ’a​xs+i​1x)\sum_{i=0}^{N}p_{i}(a,s)(f(x))^{-a}x^{s+i}\frac{1}{x}=\frac{d}{dx}\left(q(a,s,x)(f(x))^{-a}x^{s+i}\frac{1}{x}\right)

Then integrate above expression with respect to xx along contour CC mentioned in proof of Theorem 1.1, integral of RHS is zero by fundamental theorem of calculus.

(b) For fixed aa, we only need to prove the statement when supK<dβ€‹β„œβ‘(a)\sup K<d\Re(a): the recurrence in (a) allows us to extend this estimate to higher β„œβ‘(s)=Οƒ\Re(s)=\sigma. Moreover, since ff is assumed to be in ℝ​[x]\mathbb{R}[x], Ff​(a,s)Β―=Ff​(aΒ―,sΒ―)\overline{F_{f}(a,s)}=F_{f}(\overline{a},\overline{s}), we can assume t>0t>0. Let Ξ²1,β‹―,Ξ²d\beta_{1},\cdots,\beta_{d} be roots of f​(x)f(x), I claim we can let Ξ΄\delta to be

0<Ξ΄=min1≀i≀d⁑{arg⁑βi}0<arg⁑βi<2​π0<\delta=\min_{1\leq i\leq d}\{\arg\beta_{i}\}\qquad 0<\arg\beta_{i}<2\pi

From (1.1), we have

Ff​(a,s)=1e2​π​i​sβˆ’1β€‹βˆ«Cf​(x)βˆ’a​xs​d​xx,dβ€‹β„œβ‘(a)>β„œβ‘(s)F_{f}(a,s)=\frac{1}{e^{2\pi is}-1}\int_{C}f(x)^{-a}x^{s}\frac{dx}{x},\quad d\Re(a)>\Re(s)

with CC the same contour as in the proof there. Fix r>0r>0 such that all roots of ff have absolute value >r>r. For given small Ξ΅>0\varepsilon>0, write Ξ΄β€²=Ξ΄βˆ’Ξ΅\delta^{\prime}=\delta-\varepsilon. We deform the contour CC into another contour that consists of three parts:

  1. (1)

    a ray from eδ′​iβ€‹βˆže^{\delta^{\prime}i}\infty to eδ′​i​re^{\delta^{\prime}i}r

  2. (2)

    a circular arc from r​eδ′​ire^{\delta^{\prime}i} to r​eβˆ’Ξ΄β€²β€‹ire^{-\delta^{\prime}i}

  3. (3)

    a ray from eβˆ’Ξ΄β€²β€‹i​re^{-\delta^{\prime}i}r to eβˆ’Ξ΄β€²β€‹iβ€‹βˆže^{-\delta^{\prime}i}\infty.

This is allowed by our choice of Ξ΄\delta: f​(x)βˆ’af(x)^{-a} is analytic on the sector |arg⁑x|<Ξ΄β€²|\arg x|<\delta^{\prime}. The integral along (1) is

βˆ’eδ′​i​(Οƒ+i​t)β€‹βˆ«r∞f​(eδ′​i​x)βˆ’a​xΟƒ+i​t​d​xx-e^{\delta^{\prime}i(\sigma+it)}\int_{r}^{\infty}f(e^{\delta^{\prime}i}x)^{-a}x^{\sigma+it}\frac{dx}{x}

whose absolute value is upper bounded by

eβˆ’Ξ΄β€²β€‹t​|∫r∞f​(eδ′​i​x)βˆ’a​xσ​d​xx|=OK,a​(eβˆ’Ξ΄β€²β€‹t)e^{-\delta^{\prime}t}\left|\int_{r}^{\infty}f(e^{\delta^{\prime}i}x)^{-a}x^{\sigma}\frac{dx}{x}\right|=O_{K,a}(e^{-\delta^{\prime}t})

Similarly, integral along (3) is OK,a​(eβˆ’(2β€‹Ο€βˆ’Ξ΄β€²)​t)O_{K,a}(e^{-(2\pi-\delta^{\prime})t}). For integral along (2): its absolute value is

|βˆ«Ξ΄β€²2β€‹Ο€βˆ’Ξ΄β€²f​(r​ei​θ)βˆ’a​(r​ei​θ)Οƒ+i​t​𝑑θ|=OK,a​(βˆ«Ξ΄β€²2β€‹Ο€βˆ’Ξ΄β€²eβˆ’t​θ​𝑑θ)=OK,a​(eβˆ’Ξ΄β€²β€‹t)+OK,a​(eβˆ’(2β€‹Ο€βˆ’Ξ΄β€²)​t)\left|\int_{\delta^{\prime}}^{2\pi-\delta^{\prime}}f(re^{i\theta})^{-a}(re^{i\theta})^{\sigma+it}d\theta\right|=O_{K,a}\left(\int_{\delta^{\prime}}^{2\pi-\delta^{\prime}}e^{-t\theta}d\theta\right)=O_{K,a}(e^{-\delta^{\prime}t})+O_{K,a}(e^{-(2\pi-\delta^{\prime})t})

therefore

Ff​(a,Οƒ+i​t)=OK,a​(eβˆ’Ξ΄β€²β€‹t)+OK,a​(eβˆ’(2β€‹Ο€βˆ’Ξ΄β€²)​t)e2​π​i​(Οƒ+i​t)βˆ’1F_{f}(a,\sigma+it)=\frac{O_{K,a}(e^{-\delta^{\prime}t})+O_{K,a}(e^{-(2\pi-\delta^{\prime})t})}{e^{2\pi i(\sigma+it)}-1}

since Ξ΄β€²<Ο€\delta^{\prime}<\pi and t>0t>0, above is OK,a​(eβˆ’Ξ΄β€²β€‹t)O_{K,a}(e^{-\delta^{\prime}t}), as desired. ∎

We say a meromorphic function f​(s)f(s) has moderate increase along imaginary direction if for every compact subset KβŠ‚β„K\subset\mathbb{R}, there exists M>0M>0 such that

|f​(Οƒ+i​t)|=O​(|t|M)ΟƒβˆˆK,|t|≫1|f(\sigma+it)|=O(|t|^{M})\qquad\sigma\in K,\quad|t|\gg 1
Lemma 1.8.

(a) Ff​(s,1βˆ’s)F_{f}(s,1-s) has moderate increase along imaginary direction.

(b) For fixed zz, Ff​(s,z)F_{f}(s,z) has moderate increase along imaginary direction in variable ss.

Proof.

(a) By creative telescoping, f​(s)=Ff​(s,1βˆ’s)f(s)=F_{f}(s,1-s) satisfies a recurrence of some order with coefficient rational functions of ss. So it suffices to prove moderate growth when β„œβ‘(s)\Re(s) is large, formula (1.1) is valid and

f​(s)=1eβˆ’2​π​i​sβˆ’1β€‹βˆ«C(x​f​(x))βˆ’s​𝑑x.f(s)=\frac{1}{e^{-2\pi is}-1}\int_{C}(xf(x))^{-s}dx.

For small r>0r>0 that we will fix later, deform contour CC into three pieces:

  • β€’

    straight line path from ∞\infty to rr above real axis;

  • β€’

    full counter-clockwise circle from rr to rr;

  • β€’

    straight line path from rr to ∞\infty below real axis.

Giving

(1.5) f​(s)=∫r∞(x​f​(x))βˆ’s​𝑑x+1eβˆ’2​π​i​sβˆ’1β€‹βˆ«02​π[r​ei​θ​f​(r​ei​θ)]βˆ’s​(i​r​ei​θ)​𝑑θf(s)=\int_{r}^{\infty}(xf(x))^{-s}dx+\frac{1}{e^{-2\pi is}-1}\int_{0}^{2\pi}[re^{i\theta}f(re^{i\theta})]^{-s}(ire^{i\theta})d\theta

Writing s=Οƒ+i​ts=\sigma+it, first term is of moderate increase:

|∫r∞(x​f​(x))βˆ’s​𝑑x|β‰€βˆ«r∞(x​f​(x))βˆ’Οƒβ€‹π‘‘x=O​(rβˆ’Οƒ+1)\left|\int_{r}^{\infty}(xf(x))^{-s}dx\right|\leq\int_{r}^{\infty}(xf(x))^{-\sigma}dx=O(r^{-\sigma+1})

Since c:=f​(0)β‰ 0c:=f(0)\neq 0 and c>0c>0, we have

f​(r​ei​θ)βˆ’s=cβˆ’s​(1+O​(r​ei​θ))βˆ’s=cβˆ’s​exp⁑[βˆ’s​log⁑(1+O​(r​ei​θ))]=cβˆ’s​exp⁑[O​(r​s​ei​θ)]​(1+O​(βˆ’r2​s)),f(re^{i\theta})^{-s}=c^{-s}(1+O(re^{i\theta}))^{-s}=c^{-s}\exp[-s\log(1+O(re^{i\theta}))]=c^{-s}\exp[O(rse^{i\theta})](1+O(-r^{2}s)),

this term is O​(1)O(1) if we choose r=|t|βˆ’1r=|t|^{-1}, thus second term in (1.5) is bounded above by

r1βˆ’Οƒβ€‹cβˆ’Οƒ|eβˆ’2​π​i​sβˆ’1|​O​(∫02​πe(1βˆ’s)​i​θ​𝑑θ)=r1βˆ’Οƒβ€‹cβˆ’Οƒ|eβˆ’2​π​i​sβˆ’1|​O​(∫02​πet​θ​𝑑θ)=O​(1)\frac{r^{1-\sigma}c^{-\sigma}}{|e^{-2\pi is}-1|}O\left(\int_{0}^{2\pi}e^{(1-s)i\theta}d\theta\right)=\frac{r^{1-\sigma}c^{-\sigma}}{|e^{-2\pi is}-1|}O\left(\int_{0}^{2\pi}e^{t\theta}d\theta\right)=O(1)

Combining both terms f​(s)=O​(rβˆ’Οƒ+1)+O​(1)=O​(|t|Οƒβˆ’1)f(s)=O(r^{-\sigma+1})+O(1)=O(|t|^{\sigma-1}), so is of moderate growth.

(b) For fixed zz, we only need to do this for β„œβ‘(s)\Re(s) sufficiently large (again by creative telescoping),

Ff​(s,z)=∫r∞f​(x)βˆ’s​xz​d​xx+i​re2​π​i​zβˆ’1β€‹βˆ«02​πf​(r​ei​θ)βˆ’s​(r​ei​θ)z​𝑑θF_{f}(s,z)=\int_{r}^{\infty}f(x)^{-s}x^{z}\frac{dx}{x}+\frac{ir}{e^{2\pi iz}-1}\int_{0}^{2\pi}f(re^{i\theta})^{-s}(re^{i\theta})^{z}d\theta

choose r=|t|βˆ’1r=|t|^{-1} and same argument as above shows the claim. ∎

2. Meromorphic continuation and residue at abscissa of convergence

We retain all notations from previous section: f​(x)βˆˆβ„β€‹[x]f(x)\in\mathbb{R}[x] is a polynomial which is positive on [0,∞)[0,\infty), d=deg⁑fd=\deg f,

Ff​(a,s)=∫0∞(f​(x))a​xsβˆ’1​𝑑x,Kf​(a,s)=Γ​(a)Γ​(s)​Γ​(a​dβˆ’s)​Ff​(a,s)F_{f}(a,s)=\int_{0}^{\infty}(f(x))^{a}x^{s-1}dx,\qquad K_{f}(a,s)=\frac{\Gamma(a)}{\Gamma(s)\Gamma(ad-s)}F_{f}(a,s)
Proposition 2.1.

The series

ΞΆf​(s)=βˆ‘n,mβ‰₯1(n​md+1​f​(nm))βˆ’s\zeta_{f}(s)=\sum_{n,m\geq 1}\left(nm^{d+1}f(\frac{n}{m})\right)^{-s}

converges absolutely when β„œβ‘(s)>2d+2\Re(s)>\frac{2}{d+2} and defines an analytic function in this region.

Proof.

Let g​(x)=xd​f​(1/x)g(x)=x^{d}f(1/x), then

ΞΆf​(s)=βˆ‘nβ‰₯m(nd+1​m​g​(mn))βˆ’s+βˆ‘m>n(n​md+1​f​(nm))βˆ’s\zeta_{f}(s)=\sum_{n\geq m}(n^{d+1}mg(\frac{m}{n}))^{-s}+\sum_{m>n}(nm^{d+1}f(\frac{n}{m}))^{-s}

Since c1<f​(x)<c2c_{1}<f(x)<c_{2} for positive constants cic_{i} on x∈[0,1]x\in[0,1]. Above two sums converge absolutely if and only if

βˆ‘nβ‰₯m(nd+1​m)βˆ’s,βˆ‘m>n(n​md+1)βˆ’s\sum_{n\geq m}(n^{d+1}m)^{-s},\qquad\sum_{m>n}(nm^{d+1})^{-s}

both converge. From βˆ‘1≀n<mnβˆ’s=O​(m1βˆ’s)\sum_{1\leq n<m}n^{-s}=O(m^{1-s}), we see this happens if and only if β„œβ‘(s)>2/(d+2)\Re(s)>2/(d+2). ∎

We remark that abscissa of convergence for series much more general than ΞΆf​(s)\zeta_{f}(s) is known, see for example [8].

From our definition of FfF_{f} and inverse Mellin transform, we have

f​(x)βˆ’a=12​π​iβ€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžFf​(a,z)​xβˆ’z​𝑑zdβ€‹β„œβ‘(a)>c>0,x>0.f(x)^{-a}=\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}F_{f}(a,z)x^{-z}dz\quad d\Re(a)>c>0,x>0.

Therefore

ΞΆf​(s)\displaystyle\zeta_{f}(s) =12​π​iβ€‹βˆ‘n,mβ‰₯1∫cβˆ’iβ€‹βˆžc+iβ€‹βˆžnβˆ’s​mβˆ’(d+1)​s​Ff​(s,z)​(nm)βˆ’z​𝑑z\displaystyle=\frac{1}{2\pi i}\sum_{n,m\geq 1}\int_{c-i\infty}^{c+i\infty}n^{-s}m^{-(d+1)s}F_{f}(s,z)(\frac{n}{m})^{-z}dz
=12​π​iβ€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžFf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z\displaystyle=\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}F_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)dz
=12​π​i​1Γ​(s)β€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžΞ“β€‹(z)​Γ​(s​dβˆ’z)​Kf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z,\displaystyle=\frac{1}{2\pi i}\frac{1}{\Gamma(s)}\int_{c-i\infty}^{c+i\infty}\Gamma(z)\Gamma(sd-z)K_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)dz,

the exchange of summation and integral holds when the series is absolute convergent, that is

β„œβ‘(s)+c>1β„œβ‘((d+1)​s)βˆ’c>1dβ€‹β„œβ‘(s)>c>0,\Re(s)+c>1\quad\Re((d+1)s)-c>1\quad d\Re(s)>c>0,

in particular when β„œβ‘(s)\Re(s) is sufficiently large. We shall use the above integral representation of ΞΆf​(s)\zeta_{f}(s) to extend it meromorphically to all β„‚\mathbb{C}.

Due to the poles of Ξ“\Gamma and ΞΆ\zeta, it is best to visualize this situation using a "singularity diagram". By Theorem 1.1, the integrand

Γ​(z)​Γ​(s​dβˆ’z)​Kf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)\Gamma(z)\Gamma(sd-z)K_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)

has poles333here we highlight importance of Theorem 1.1, which says Kf​(s,z)K_{f}(s,z) is an entire function. at

z=βˆ’n,z=d​s+n,z=1βˆ’s,z=(d+1)​sβˆ’1​ with ​nβˆˆβ„€β‰₯0z=-n,\quad z=ds+n,\quad z=1-s,\quad z=(d+1)s-1\text{ with }n\in\mathbb{Z}^{\geq 0}

The diagram is constructed by plotting these lines with respect to coordinates (β„œβ‘(s),β„œβ‘(z))=(β„œβ‘(s),c)(\Re(s),\Re(z))=(\Re(s),c). Note that his diagram depends only on d=deg⁑fd=\deg f, not on particular appearance of ff.

Refer to caption
Figure 1. Singularity diagram when d=2d=2. Readers can convince themselves this diagram remain essentially same for other dd.

These lines cut the plane into different connected components. The component of point AA corresponds to the choice of (β„œβ‘(s),c)(\Re(s),c) for which our original integral representation

ΞΆf​(s)=12​π​i​1Γ​(s)β€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžΞ“β€‹(z)​Γ​(s​dβˆ’z)​Kf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z:=12​π​iβ€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžHf​(s,z)​𝑑z\zeta_{f}(s)=\frac{1}{2\pi i}\frac{1}{\Gamma(s)}\int_{c-i\infty}^{c+i\infty}\Gamma(z)\Gamma(sd-z)K_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)dz:=\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}H_{f}(s,z)dz

is valid. Since

Hf​(s,z)=Ff​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z),H_{f}(s,z)=F_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z),

Lemma 1.7 tells us the integrand is exponentially decreasing along path of integration, we can shift the contour of integration with β„œβ‘(z)=c\Re(z)=c to an arbitrary cβ€²c^{\prime}, picking up the residue in between:

∫cβˆ’iβ€‹βˆžc+iβ€‹βˆžHf​(s,z)​𝑑z=∫cβ€²βˆ’iβ€‹βˆžcβ€²+iβ€‹βˆžHf​(s,z)​𝑑zΒ±(Sum of residues at poles between ​c​ and ​cβ€²)\int_{c-i\infty}^{c+i\infty}H_{f}(s,z)dz=\int_{c^{\prime}-i\infty}^{c^{\prime}+i\infty}H_{f}(s,z)dz\pm\left(\text{Sum of residues at poles between }c\text{ and }c^{\prime}\right)

here ++ is taken if cβ€²<cc^{\prime}<c and βˆ’- if cβ€²>cc^{\prime}>c.

For example, we can cross the line β„œβ‘(z)=1βˆ’β„œβ‘(s)\Re(z)=1-\Re(s) to reach point BB, then

ΞΆf​(s)\displaystyle\zeta_{f}(s) =Resz=1βˆ’s​[Hf​(s,z)]+12​π​iβ€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžHf​(s,z)​𝑑z\displaystyle=\text{Res}_{z=1-s}[H_{f}(s,z)]+\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}H_{f}(s,z)dz
=Γ​(1βˆ’s)​Γ​(d​s+sβˆ’1)Γ​(s)​΢​((d+2)​sβˆ’1)​Kf​(s,1βˆ’s)+12​π​iβ€‹βˆ«cβˆ’iβ€‹βˆžc+iβ€‹βˆžHf​(s,z)​𝑑z\displaystyle=\frac{\Gamma(1-s)\Gamma(ds+s-1)}{\Gamma(s)}\zeta((d+2)s-1)K_{f}(s,1-s)+\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}H_{f}(s,z)dz

where now (β„œβ‘(s),c)(\Re(s),c) in last integral should be in the connected component of point BB, it is analytic as a function of ss for whatever β„œβ‘(s)\Re(s) in this region. In particular, it is analytic near s=2/(d+2)s=2/(d+2), the term in front has a simple pole at this point, so we established

Theorem 2.2.

ΞΆf​(s)\zeta_{f}(s) has a simple pole at s=2/(d+2)s=2/(d+2), with residue

1d+2​Ff​(2d+2,dd+2)=1d+2β€‹βˆ«0∞(x​f​(x))βˆ’2/(d+2)​𝑑x\frac{1}{d+2}F_{f}(\frac{2}{d+2},\frac{d}{d+2})=\frac{1}{d+2}\int_{0}^{\infty}(xf(x))^{-2/(d+2)}dx

For general ff, one does not expect this integral to simplify further. However, for ff arising from rank 22 root systems, they all can be evaluated nicely:

Theorem 2.3.

Residue of Ο‰A​(s)\omega_{A}(s) at s=2/3s=2/3 is

Γ​(13)32​3​π.\frac{\Gamma(\frac{1}{3})^{3}}{2\sqrt{3}\pi}.

Residue of Ο‰B​(s)\omega_{B}(s) at s=1/2s=1/2 is

Γ​(14)28​2​π.\frac{\Gamma\left(\frac{1}{4}\right)^{2}}{8\sqrt{2\pi}}.

Residue of Ο‰G​(s)\omega_{G}(s) at s=1/3s=1/3 is

Γ​(13)328/3​33/2​π.\frac{\Gamma\left(\frac{1}{3}\right)^{3}}{2^{8/3}3^{3/2}\pi}.
Proof.

We skip proofs for Ο‰A,Ο‰B\omega_{A},\omega_{B}, partly because evaluation of integrals ∫0∞(x​f​(x))βˆ’2/(d+2)\int_{0}^{\infty}(xf(x))^{-2/(d+2)} are trivial in these cases, partly because they are known result ([19] for Ο‰A\omega_{A} and [4] for Ο‰B\omega_{B}). We concentrate on Ο‰G\omega_{G}, the claim is equivalent to non-trivial evaluation

∫0∞d​xx​(1+x)​(1+2​x)​(1+3​x)​(2+3​x)3=125/3​31/2​π​Γ​(13)3.\int_{0}^{\infty}\frac{dx}{\sqrt[3]{x(1+x)(1+2x)(1+3x)(2+3x)}}=\frac{1}{2^{5/3}3^{1/2}\pi}\Gamma\left(\frac{1}{3}\right)^{3}.

In order not to interrupt our discussion, we put its proof in Appendix. ∎

Back to meromorphic continuation of ΞΆf​(s)\zeta_{f}(s). In the diagram, we can continue from point BB to CC, to DD and so on, picking up residues at z=0,βˆ’1,βˆ’2,β‹―z=0,-1,-2,\cdots in the process, so we obtain

(2.1) ΞΆf​(s)=Γ​(1βˆ’s)​Γ​(d​s+sβˆ’1)Γ​(s)​΢​((d+2)​sβˆ’1)​Kf​(s,1βˆ’s)+βˆ‘i=0M(βˆ’1)ii!​Γ​(s​d+i)Γ​(s)​Kf​(s,βˆ’i)​΢​(sβˆ’i)​΢​((d+1)​s+i)+1Γ​(s)​12​π​iβ€‹βˆ«βˆ’Mβˆ’1/2βˆ’iβ€‹βˆžβˆ’Mβˆ’1/2+iβ€‹βˆžΞ“β€‹(z)​Γ​(s​dβˆ’z)​Kf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z\zeta_{f}(s)=\frac{\Gamma(1-s)\Gamma(ds+s-1)}{\Gamma(s)}\zeta((d+2)s-1)K_{f}(s,1-s)+\sum_{i=0}^{M}\frac{(-1)^{i}}{i!}\frac{\Gamma(sd+i)}{\Gamma(s)}K_{f}(s,-i)\zeta(s-i)\zeta((d+1)s+i)\\ +\frac{1}{\Gamma(s)}\frac{1}{2\pi i}\int_{-M-1/2-i\infty}^{-M-1/2+i\infty}\Gamma(z)\Gamma(sd-z)K_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)dz

for any non-negative integer MM. From the diagram, we see that as MM gets sufficiently large, the latter integral becomes analytic at any point sβˆˆβ„‚s\in\mathbb{C}, so we proved the meromorphic continuation of ΞΆf​(s)\zeta_{f}(s).

Proposition 2.4.

ΞΆf​(s)\zeta_{f}(s) is moderately increasing along imaginary direction, that is, for any compact subset KβŠ‚β„K\subset\mathbb{R}, there exists M>0M>0 such that

|ΞΆf​(Οƒ+i​t)|=O​(|t|M),ΟƒβˆˆK,|t|≫1|\zeta_{f}(\sigma+it)|=O(|t|^{M}),\qquad\sigma\in K,|t|\gg 1
Proof.

Rewrite equation 2.1 as

ΞΆf​(s)=΢​((d+2)​sβˆ’1)​Ff​(s,1βˆ’s)+βˆ‘i=0M΢​(sβˆ’i)​΢​((d+1)​s+i)​[f​(x)βˆ’s]​[xi]+12​π​iβ€‹βˆ«βˆ’Mβˆ’1/2βˆ’iβ€‹βˆžβˆ’Mβˆ’1/2+iβ€‹βˆžΞΆβ€‹(s+z)​΢​((d+1)​sβˆ’z)​Ff​(s,z)​𝑑z\zeta_{f}(s)=\zeta((d+2)s-1)F_{f}(s,1-s)+\sum_{i=0}^{M}\zeta(s-i)\zeta((d+1)s+i)[f(x)^{-s}][x^{i}]\\ +\frac{1}{2\pi i}\int_{-M-1/2-i\infty}^{-M-1/2+i\infty}\zeta(s+z)\zeta((d+1)s-z)F_{f}(s,z)dz

Recall ΢​(s)\zeta(s) is moderately increasing, Ff​(s,1βˆ’s)F_{f}(s,1-s) is too (Lemma 1.8), so first two terms are moderately increasing. As for the integral, same lemma implies Ff​(s,z)F_{f}(s,z) is moderately increasing in ss when β„œβ‘(z)\Re(z) is fixed. ∎

Actually, a much more general moderate growth result on zeta functions of hyperplane arrangement is known, see [7].

3. Values at non-positive integers

In the representation (2.1), when let ss tends to βˆ’n-n, where nn is a non-negative integer, the last term involving the integral is 0 because of 1/Γ​(s)1/\Gamma(s) at the front. As for the sum,

βˆ‘i=0M(βˆ’1)ii!​Γ​(s​d+i)Γ​(s)​Kf​(s,βˆ’i)​΢​(sβˆ’i)​΢​((d+1)​s+i)\sum_{i=0}^{M}\frac{(-1)^{i}}{i!}\frac{\Gamma(sd+i)}{\Gamma(s)}K_{f}(s,-i)\zeta(s-i)\zeta((d+1)s+i)

when sβ†’βˆ’ns\to-n, only terms that are non-zero come from ii such that 0≀i≀n​d0\leq i\leq nd and i=(d+1)​n+1i=(d+1)n+1. Now it is easy to write down the formula of ΞΆf​(βˆ’n)\zeta_{f}(-n):

ΞΆf​(βˆ’n)=(n!)2d+1​(βˆ’1)d​n+1((d+1)​n+1)!​΢​(βˆ’(d+2)​nβˆ’1)​[Kf​(βˆ’n,1+n)+Kf​(βˆ’n,βˆ’1βˆ’(d+1)​n)]+βˆ‘i=0n​d(βˆ’1)n​(d+1)​n!dΓ—i!​(n​dβˆ’i)!​Kf​(βˆ’n,βˆ’i)​΢​(βˆ’nβˆ’i)​΢​(βˆ’(d+1)​n+i)\zeta_{f}(-n)=\frac{(n!)^{2}}{d+1}\frac{(-1)^{dn+1}}{((d+1)n+1)!}\zeta(-(d+2)n-1)\left[K_{f}(-n,1+n)+K_{f}(-n,-1-(d+1)n)\right]\\ +\sum_{i=0}^{nd}\frac{(-1)^{n(d+1)}n!}{d\times i!(nd-i)!}K_{f}(-n,-i)\zeta(-n-i)\zeta(-(d+1)n+i)

Note that Kf​(βˆ’n,1+n)=Kg​(βˆ’n,βˆ’1βˆ’(d+1)​n)K_{f}(-n,1+n)=K_{g}(-n,-1-(d+1)n) with g​(x)=xd​f​(1/x)g(x)=x^{d}f(1/x). Some simplifications using Proposition 1.4 yields the following form:

Theorem 3.1.

ΞΆf​(s)\zeta_{f}(s) is analytic at s=0,βˆ’1,βˆ’2,β‹―s=0,-1,-2,\cdots and

ΞΆf​(βˆ’n)=΢​(βˆ’(d+2)​nβˆ’1)d+1​([βˆ’f​(x)n​log⁑f​(x)]​[x1+(1+d)​n]+[βˆ’g​(x)n​log⁑g​(x)]​[x1+(1+d)​n])+βˆ‘i=0n​d[f​(x)n]​[xi]​΢​(βˆ’nβˆ’i)​΢​(βˆ’(d+1)​n+i),\zeta_{f}(-n)=\frac{\zeta(-(d+2)n-1)}{d+1}\left([-f(x)^{n}\log f(x)][x^{1+(1+d)n}]+[-g(x)^{n}\log g(x)][x^{1+(1+d)n}]\right)\\ +\sum_{i=0}^{nd}[f(x)^{n}][x^{i}]\zeta(-n-i)\zeta(-(d+1)n+i),

where g​(x)=xd​f​(1/x)g(x)=x^{d}f(1/x).

Corollary 3.2.

When d=deg⁑fd=\deg f is odd and nn is a non-negative integer, we have

ΞΆf​(βˆ’2​nβˆ’1)=0\zeta_{f}(-2n-1)=0
Proof.

Conditions on parity of dd and nn ensure every term in above expression is zero because Riemann’s zeta at negative even integers are 0. ∎

Example 3.3.

Let f​(x)=(1+x)​(1+3​x)f(x)=(1+x)(1+3x),

ΞΆf​(s)=βˆ‘n,mβ‰₯11ns​ms​(n+m)s​(n+3​m)s.\zeta_{f}(s)=\sum_{n,m\geq 1}\frac{1}{n^{s}m^{s}(n+m)^{s}(n+3m)^{s}}.

ΞΆf​(0),ΞΆf​(βˆ’1),β‹―,ΞΆf​(βˆ’4)\zeta_{f}(0),\zeta_{f}(-1),\cdots,\zeta_{f}(-4) equal respectively

43108,βˆ’2387480,80972171,1082823151963120,2383543667157837977.\frac{43}{108},\quad-\frac{23}{87480},\quad\frac{809}{72171},\quad\frac{10828231}{51963120},\quad\frac{2383543667}{157837977}.

The values of ΞΆf​(βˆ’n)\zeta_{f}(-n) with ff coming from root system, i.e., Ο‰A​(s),Ο‰B​(s)\omega_{A}(s),\omega_{B}(s) and Ο‰G​(s)\omega_{G}(s) are much more interesting.

Example 3.4.

For Ο‰A​(s)\omega_{A}(s), one calculates, using above formula, that Ο‰A​(0)=1/3\omega_{A}(0)=1/3 and one quickly notices Ο‰A​(βˆ’n)=0\omega_{A}(-n)=0 for first few nn.

This is indeed true in general. For odd nn, this is explained by above corollary; the vanishing of Ο‰A\omega_{A} at even nn is a much deeper fact (see below).

Example 3.5.

For Ο‰B​(s)\omega_{B}(s), its values at s=0,βˆ’1,β‹―,βˆ’6s=0,-1,\cdots,-6 are

38,βˆ’114480,0,βˆ’45811576960,0,βˆ’287820799443256502005760,0.\frac{3}{8},\quad-\frac{11}{4480},\quad 0,\quad-\frac{4581}{1576960},\quad 0,\quad-\frac{287820799443}{256502005760},\quad 0.

For Ο‰G​(s)\omega_{G}(s), its values at s=0,βˆ’1,β‹―,βˆ’6s=0,-1,\cdots,-6 are

512,332052612736,0,3130718204745814251580471680499712,0,340671926141776900457938568106968752320107746327814532497408,0.\frac{5}{12},\quad\frac{33205}{2612736},\quad 0,\quad\frac{313071820474581425}{1580471680499712},\quad 0,\quad\frac{34067192614177690045793856810696875}{2320107746327814532497408},\quad 0.

These values have already been reported in [20]. The obvious pattern Ο‰B​(βˆ’2​n)=Ο‰G​(βˆ’2​n)=0\omega_{B}(-2n)=\omega_{G}(-2n)=0 for nβ‰₯1n\geq 1 is indeed true. It is special case of the following deep result.

Theorem 3.6 ([1]).

Let Ξ¦\Phi be a root system, mm be a negative even integer, the order of vanishing of ΢Φ​(s)\zeta_{\Phi}(s) at s=ms=m is at least the rank of Ξ¦\Phi, that is 444Here ords=s0​f​(s)\text{ord}_{s=s_{0}}f(s) for a meromorphic function is order of zero or pole at point s=s0s=s_{0}.

ords=m​(΢Φ​(s))β‰₯Β Rank ofΒ Ξ¦.\text{ord}_{s=m}(\zeta_{\Phi}(s))\geq\text{ Rank of $\Phi$}.

Its proof relies on root system’s special symmetry and employs technique of different nature. It follows Ο‰A​(s),Ο‰B​(s),Ο‰G​(s)\omega_{A}(s),\omega_{B}(s),\omega_{G}(s) not only vanish at s=βˆ’2​ns=-2n, but also these are at least double zeroes. For Ο‰A​(s)\omega_{A}(s), this is already shown in [17].

3.1. A connection to Eisenstein series

We make some interesting observations on vanishing of Ο‰B​(βˆ’2​n),Ο‰G​(βˆ’2​n)\omega_{B}(-2n),\omega_{G}(-2n).

First we write out, using Theorem 3.1, an explicit expression that is equivalent to vanishing of Ο‰βˆ™β€‹(βˆ’2​n)\omega_{\bullet}(-2n) with nβ‰₯1n\geq 1.

For Ο‰A\omega_{A}, it is better to directly use the fact that Kf​(a,s)≑1K_{f}(a,s)\equiv 1 for f​(x)=1+xf(x)=1+x, arriving at:

(3.1) Ο‰A​(βˆ’2​n)=0⇔(2​n)!(4​n+1)!​΢​(βˆ’6​nβˆ’1)=βˆ‘i=02​n1i!​(2​nβˆ’i)!​΢​(βˆ’iβˆ’2​n)​΢​(iβˆ’4​n)\omega_{A}(-2n)=0\iff\frac{(2n)!}{(4n+1)!}\zeta(-6n-1)=\sum_{i=0}^{2n}\frac{1}{i!(2n-i)!}\zeta(-i-2n)\zeta(i-4n)

For Ο‰B\omega_{B} and Ο‰G\omega_{G}, denote fB​(x)=(1+x)​(1+2​x)f_{B}(x)=(1+x)(1+2x) and fG​(x)=(1+x)​(1+2​x)​(1+3​x)​(2+3​x)f_{G}(x)=(1+x)(1+2x)(1+3x)(2+3x),

Ο‰B​(βˆ’2​n)=0⇔΢​(βˆ’8​nβˆ’1)3​(1+2βˆ’1βˆ’4​n)​[fB​(x)2​n​log⁑fB​(x)]​[x1+6​n]=βˆ‘i=04​n[fB​(x)2​n]​[xi]​΢​(βˆ’iβˆ’2​n)​΢​(iβˆ’6​n)\omega_{B}(-2n)=0\iff\frac{\zeta(-8n-1)}{3}(1+2^{-1-4n})[f_{B}(x)^{2n}\log f_{B}(x)][x^{1+6n}]=\sum_{i=0}^{4n}[f_{B}(x)^{2n}][x^{i}]\zeta(-i-2n)\zeta(i-6n)
Ο‰G​(βˆ’2​n)=0⇔΢​(βˆ’12​nβˆ’1)5​(1+3βˆ’1βˆ’6​n)​[fG​(x)2​n​log⁑fG​(x)]​[x1+10​n]=βˆ‘i=08​n[fG​(x)2​n]​[xi]​΢​(βˆ’iβˆ’2​n)​΢​(iβˆ’10​n)\omega_{G}(-2n)=0\iff\frac{\zeta(-12n-1)}{5}(1+3^{-1-6n})[f_{G}(x)^{2n}\log f_{G}(x)][x^{1+10n}]=\sum_{i=0}^{8n}[f_{G}(x)^{2n}][x^{i}]\zeta(-i-2n)\zeta(i-10n)

Recall following Eisenstein series, for Ο„\tau in upper-half plane,

E2​k​(Ο„)=12​΢​(2​k)β€‹βˆ‘(m,n)βˆˆβ„€2(m,n)β‰ (0,0)1(m+n​τ)2​kkβ‰₯2E_{2k}(\tau)=\frac{1}{2\zeta(2k)}\sum_{\begin{subarray}{c}(m,n)\in\mathbb{Z}^{2}\\ (m,n)\neq(0,0)\end{subarray}}\frac{1}{(m+n\tau)^{2k}}\qquad k\geq 2

We also agree that Ek​(Ο„)=0E_{k}(\tau)=0 for kβ‰₯3k\geq 3 odd. Note that limℑ⁑(Ο„)β†’βˆžE2​k​(Ο„)=1\lim_{\Im(\tau)\to\infty}E_{2k}(\tau)=1.

Romik ([19]) observed that, in equation (3.1), if we replace every occurrence of ΢​(1βˆ’2​k)\zeta(1-2k) by ΢​(1βˆ’2​k)​E2​k​(Ο„)\zeta(1-2k)E_{2k}(\tau), we still obtain a valid equality. That is

(2​n)!(4​n+1)!​΢​(βˆ’6​nβˆ’1)​E6​n+2​(Ο„)=βˆ‘i=02​n1i!​(2​nβˆ’i)!​΢​(βˆ’iβˆ’2​n)​΢​(iβˆ’4​n)​E2​n+i+1​(Ο„)​E4​nβˆ’i+1​(Ο„)\frac{(2n)!}{(4n+1)!}\zeta(-6n-1)E_{6n+2}(\tau)=\sum_{i=0}^{2n}\frac{1}{i!(2n-i)!}\zeta(-i-2n)\zeta(i-4n)E_{2n+i+1}(\tau)E_{4n-i+1}(\tau)

the original equality (3.1) is recovered by letting ℑ⁑(Ο„)β†’βˆž\Im(\tau)\to\infty. Romik gave a proof that applies to both (3.1) and the Eisenstein series version.

Surprisingly, when one performs the same replacement to the equalities coming from Ο‰B,Ο‰G\omega_{B},\omega_{G}, they seem still hold555They have been checked for n≀10n\leq 10. We formulate them as conjectures.

Conjecture 3.7.

For positive integers nn, the following two equalities hold:

΢​(βˆ’8​nβˆ’1)​E8​n+2​(Ο„)3​(1+2βˆ’1βˆ’4​n)Γ—[fB​(x)2​n​log⁑fB​(x)]​[x1+6​n]=βˆ‘i=04​n[fB​(x)2​n]​[xi]​΢​(βˆ’iβˆ’2​n)​΢​(iβˆ’6​n)​E2​n+i+1​(Ο„)​E6​nβˆ’i+1​(Ο„)\frac{\zeta(-8n-1)E_{8n+2}(\tau)}{3}(1+2^{-1-4n})\times[f_{B}(x)^{2n}\log f_{B}(x)][x^{1+6n}]\\ =\sum_{i=0}^{4n}[f_{B}(x)^{2n}][x^{i}]\zeta(-i-2n)\zeta(i-6n)E_{2n+i+1}(\tau)E_{6n-i+1}(\tau)
΢​(βˆ’12​nβˆ’1)​E12​n+2​(Ο„)5​(1+3βˆ’1βˆ’6​n)Γ—[fG​(x)2​n​log⁑fG​(x)]​[x1+10​n]=βˆ‘i=08​n[fG​(x)2​n]​[xi]​΢​(βˆ’iβˆ’2​n)​΢​(iβˆ’10​n)​E2​n+i+1​(Ο„)​E10​nβˆ’i+1​(Ο„)\frac{\zeta(-12n-1)E_{12n+2}(\tau)}{5}(1+3^{-1-6n})\times[f_{G}(x)^{2n}\log f_{G}(x)][x^{1+10n}]\\ =\sum_{i=0}^{8n}[f_{G}(x)^{2n}][x^{i}]\zeta(-i-2n)\zeta(i-10n)E_{2n+i+1}(\tau)E_{10n-i+1}(\tau)

4. Poles of ΞΆf​(s)\zeta_{f}(s)

Apart from the pole of ΞΆf​(s)\zeta_{f}(s) at abscissa of convergence s=2/(d+2)s=2/(d+2), it also has a family of simple poles at other ss, they are easily described.

Theorem 4.1.

Let s0β‰ 2/(d+2)s_{0}\neq 2/(d+2) such that ΞΆf​(s)\zeta_{f}(s) have a pole at s=s0s=s_{0}. Then this pole is (at most) simple, s0βˆ‰β„€s_{0}\notin\mathbb{Z}, s0​(1+d)βˆ’1=βˆ’nβˆˆβ„€β‰€0s_{0}(1+d)-1=-n\in\mathbb{Z}^{\leq 0}, the residue at this point is666recall [h​(x)]​[xm][h(x)][x^{m}] means coefficient of xmx^{m} in series expansion of h​(x)h(x) around x=0x=0.

΢​((d+2)​s0βˆ’1)d+1​([f​(x)βˆ’s0]​[xn]+[g​(x)βˆ’s0]​[xn])\frac{\zeta((d+2)s_{0}-1)}{d+1}\left([f(x)^{-s_{0}}][x^{n}]+[g(x)^{-s_{0}}][x^{n}]\right)

where g​(x)=xd​f​(1/x)g(x)=x^{d}f(1/x).

Proof.

We have seen that ΞΆf​(s)\zeta_{f}(s) is analytic at integer ss, so s0s_{0} is not an integer. From (2.1),

ΞΆf​(s)=Γ​(1βˆ’s)​Γ​((d+1)​sβˆ’1)Γ​(s)​΢​((d+2)​sβˆ’1)​Kf​(s,1βˆ’s)+βˆ‘i=0M(βˆ’1)ii!​Γ​(s​d+i)Γ​(s)​Kf​(s,βˆ’i)​΢​(sβˆ’i)​΢​((d+1)​s+i)+1Γ​(s)​12​π​iβ€‹βˆ«βˆ’Mβˆ’1/2βˆ’iβ€‹βˆžβˆ’Mβˆ’1/2+iβ€‹βˆžΞ“β€‹(z)​Γ​(s​dβˆ’z)​Kf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z\zeta_{f}(s)=\frac{\Gamma(1-s)\Gamma((d+1)s-1)}{\Gamma(s)}\zeta((d+2)s-1)K_{f}(s,1-s)+\sum_{i=0}^{M}\frac{(-1)^{i}}{i!}\frac{\Gamma(sd+i)}{\Gamma(s)}K_{f}(s,-i)\zeta(s-i)\zeta((d+1)s+i)\\ +\frac{1}{\Gamma(s)}\frac{1}{2\pi i}\int_{-M-1/2-i\infty}^{-M-1/2+i\infty}\Gamma(z)\Gamma(sd-z)K_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)dz

For sβˆ‰β„€s\notin\mathbb{Z}, the only possible pole that could arise are those coming from

΢​((d+2)​sβˆ’1)​ or ​Γ​((d+1)​sβˆ’1)​ or ​Γ​(s​d+i).\zeta((d+2)s-1)\text{ or }\Gamma((d+1)s-1)\text{ or }\Gamma(sd+i).

The first choice gives s=2/(d+2)s=2/(d+2), which we already excluded. The third case cannot give a pole either: let sβ†’βˆ’n0/ds\to-n_{0}/d where n0β‰₯β„€β‰₯0n_{0}\geq\mathbb{Z}^{\geq 0}, then residue of the sum βˆ‘i=0M\sum_{i=0}^{M} term is

βˆ‘i=0n0(βˆ’1)ii!​Γ​(βˆ’n0/d)​(βˆ’1)n0βˆ’id​(n0βˆ’i)!​Kf​(βˆ’n0d,βˆ’i)​΢​(βˆ’n0dβˆ’i)​΢​(βˆ’n0d+iβˆ’n0)\sum_{i=0}^{n_{0}}\frac{(-1)^{i}}{i!\Gamma(-n_{0}/d)}\frac{(-1)^{n_{0}-i}}{d(n_{0}-i)!}K_{f}(-\frac{n_{0}}{d},-i)\zeta(-\frac{n_{0}}{d}-i)\zeta(-\frac{n_{0}}{d}+i-n_{0})

for this range of ii, Proposition 1.3, together with the assumption βˆ’n0/dβˆ‰β„€-n_{0}/d\notin\mathbb{Z}, implies Kf​(βˆ’n0d,βˆ’i)=0K_{f}(-\frac{n_{0}}{d},-i)=0, so above sum is zero.

So only the second case remains: sβ†’s0s\to s_{0} such that s0​(1+d)βˆ’1=βˆ’n0s_{0}(1+d)-1=-n_{0} with nβˆˆβ„€β‰₯0n\in\mathbb{Z}^{\geq 0}. At this point (d+2)​sβˆ’1,s​d+iβˆ‰β„€(d+2)s-1,sd+i\notin\mathbb{Z}, so only contribution of residue comes from the term Γ​(1βˆ’s)​Γ​((d+1)​sβˆ’1)Γ​(s)​΢​((d+2)​sβˆ’1)​Kf​(s,1βˆ’s)\frac{\Gamma(1-s)\Gamma((d+1)s-1)}{\Gamma(s)}\zeta((d+2)s-1)K_{f}(s,1-s) and inside summation with i=βˆ’(d+1)​s0+1=ni=-(d+1)s_{0}+1=n. A straightforward calculation says residue is

(βˆ’1)nn!​Γ​(1βˆ’s0)Γ​(s0)​΢​((d+2)​s0βˆ’1)​11+d​[Kf​(s0,1βˆ’s0)+Kf​(s0,βˆ’n)]\frac{(-1)^{n}}{n!}\frac{\Gamma(1-s_{0})}{\Gamma(s_{0})}\zeta((d+2)s_{0}-1)\frac{1}{1+d}\left[K_{f}(s_{0},1-s_{0})+K_{f}(s_{0},-n)\right]

Note that Kf​(s0,1βˆ’s0)=Kg​(s0,βˆ’n)K_{f}(s_{0},1-s_{0})=K_{g}(s_{0},-n) and Proposition 1.3 completes the proof. ∎

Corollary 4.2.

Let f​(x)=c0+c1​x+β‹―+cd​xdf(x)=c_{0}+c_{1}x+\cdots+c_{d}x^{d}. Apart from s=2/(2+d)s=2/(2+d), the only other pole of ΞΆf​(s)\zeta_{f}(s) with β„œβ‘(s)β‰₯0\Re(s)\geq 0 is s=1/(1+d)s=1/(1+d), it is simple and residue there is

c0βˆ’1/(d+1)+cdβˆ’1/(d+1)d+1​΢​(1d+1)\frac{c_{0}^{-1/(d+1)}+c_{d}^{-1/(d+1)}}{d+1}\zeta(\frac{1}{d+1})
Proof.

This is special case s0=1/(d+1),n=0s_{0}=1/(d+1),n=0 in above theorem. ∎

Example 4.3.

For Ο‰A​(s)\omega_{A}(s), we have d=1d=1, and nn has to be even. Letting k=2​nk=2n, residue at s=1/2βˆ’ks=1/2-k with kβ‰₯0k\geq 0 is

΢​(12βˆ’3​k)​(kβˆ’1/22​k)\zeta(\frac{1}{2}-3k)\binom{k-1/2}{2k}

a result already obtained by Romik [19].

Proposition 4.4.

(a) Let nn be a non-negative integer, residue of Ο‰B​(s)\omega_{B}(s) at s=(1βˆ’n)/3s=(1-n)/3 is

13​΢​(1βˆ’4​n3)​(1+2(βˆ’1βˆ’2​n)/3)​bn\frac{1}{3}\zeta(\frac{1-4n}{3})(1+2^{(-1-2n)/3})b_{n}

with bnβˆˆβ„šb_{n}\in\mathbb{Q} equals the coefficient of xnx^{n} in power series expansion of ((1+x)​(1+2​x))(nβˆ’1)/3\left((1+x)(1+2x)\right)^{(n-1)/3}. Moreover, bnb_{n} satisfies the recurrence

27​(n+4)​(n+6)​bn+6βˆ’4​(n+2)​(n+5)​bn=027(n+4)(n+6)b_{n+6}-4(n+2)(n+5)b_{n}=0

(b) Let nn a non-negative integer, residue of Ο‰G​(s)\omega_{G}(s) at s=(1βˆ’n)/5s=(1-n)/5 is

15​΢​(1βˆ’6​n5)​(1+3(βˆ’3​nβˆ’2)/5)​2(nβˆ’1)/5​gn\frac{1}{5}\zeta(\frac{1-6n}{5})(1+3^{(-3n-2)/5})2^{(n-1)/5}g_{n}

with gnβˆˆβ„šg_{n}\in\mathbb{Q} equals the coefficient of xnx^{n} in power series expansion of ((1+x)​(1+2​x)​(1+3​x)​(1+32​x))(nβˆ’1)/5\left((1+x)(1+2x)(1+3x)(1+\frac{3}{2}x)\right)^{(n-1)/5}. Moreover, gng_{n} satisfies the recurrence

729​(n+4)​(n+9)​(n+14)​(n+19)​(6​n+103)​gnβˆ’1875​(n+14)​(n+19)​(216​n3+7596​n2+83694​n+290603)​gn+10+50000​(n+12)​(n+16)​(n+18)​(n+20)​(6​n+43)​gn+20=0729(n+4)(n+9)(n+14)(n+19)(6n+103)g_{n}-1875(n+14)(n+19)\left(216n^{3}+7596n^{2}+83694n+290603\right)g_{n+10}\\ +50000(n+12)(n+16)(n+18)(n+20)(6n+43)g_{n+20}=0
Proof.

Firstly, note that bn=0b_{n}=0 when n≑1(mod3)n\equiv 1\pmod{3} and gn=0g_{n}=0 when n≑1(mod5)n\equiv 1\pmod{5}, because ((1+x)​(1+2​x))(nβˆ’1)/3\left((1+x)(1+2x)\right)^{(n-1)/3} and ((1+x)​(1+2​x)​(1+3​x)​(1+32​x))(nβˆ’1)/5\left((1+x)(1+2x)(1+3x)(1+\frac{3}{2}x)\right)^{(n-1)/5} become polynomials, this is consistent with the fact that Ο‰B​(s),Ο‰G​(s)\omega_{B}(s),\omega_{G}(s) being analytic at integral ss.

The assertion on expression of residue follows immediately from Theorem 4.1. Here one notes that, for

fB​(x)=(1+x)​(1+2​x),gB​(x)=x2​fB​(1/x),f_{B}(x)=(1+x)(1+2x),\quad g_{B}(x)=x^{2}f_{B}(1/x),

gB​(2​x)=2​fB​(x)⟹[gB​(x)βˆ’s]​[xn]=2βˆ’sβˆ’n​[fB​(x)s]​[xn]g_{B}(2x)=2f_{B}(x)\implies[g_{B}(x)^{-s}][x^{n}]=2^{-s-n}[f_{B}(x)^{s}][x^{n}]; and similarly

fG​(x)=(1+x)​(1+2​x)​(1+3​x)​(2+3​x),gG​(x)=x4​fG​(1/x),f_{G}(x)=(1+x)(1+2x)(1+3x)(2+3x),\quad g_{G}(x)=x^{4}f_{G}(1/x),

gG​(3​x)=9​fG​(x)⟹[gG​(x)βˆ’s]​[xn]=3βˆ’2​sβˆ’n​[fG​(x)s]​[xn]g_{G}(3x)=9f_{G}(x)\implies[g_{G}(x)^{-s}][x^{n}]=3^{-2s-n}[f_{G}(x)^{s}][x^{n}]. It remains to prove the recurrences satisfied by bnb_{n} and gng_{n}, this is follows from creative telescoping: we have

bn=12​π​iβ€‹βˆ«C​(Ξ΅)(fB​(x))(nβˆ’1)/3​xβˆ’nβˆ’1​𝑑x,gn=12​π​iβ€‹βˆ«C​(Ξ΅)(fG​(x))(nβˆ’1)/5​xβˆ’nβˆ’1​𝑑x,b_{n}=\frac{1}{2\pi i}\int_{C(\varepsilon)}(f_{B}(x))^{(n-1)/3}x^{-n-1}dx,\qquad g_{n}=\frac{1}{2\pi i}\int_{C(\varepsilon)}(f_{G}(x))^{(n-1)/5}x^{-n-1}dx,

where C​(Ξ΅)C(\varepsilon) is a small counter-clockwise circle around x=0x=0. Let F​(x,n)=(fB​(x))(nβˆ’1)/3​xβˆ’nβˆ’1F(x,n)=(f_{B}(x))^{(n-1)/3}x^{-n-1}, then one checks,

27​(4+n)​(6+n)​F​(x,n+6)βˆ’4​(2+n)​(5+n)​F​(x,n)=dd​x​(3​(x+1)​(2​x+1)​(2​n​x4βˆ’6​n​x3βˆ’39​n​x2βˆ’36​n​xβˆ’9​n+10​x4βˆ’12​x3βˆ’150​x2βˆ’144​xβˆ’36)x5​F​(x,n)).27(4+n)(6+n)F(x,n+6)-4(2+n)(5+n)F(x,n)\\ =\frac{d}{dx}\left(\frac{3(x+1)(2x+1)\left(2nx^{4}-6nx^{3}-39nx^{2}-36nx-9n+10x^{4}-12x^{3}-150x^{2}-144x-36\right)}{x^{5}}F(x,n)\right).

Applying ∫C​(Ξ΅)𝑑x\int_{C(\varepsilon)}dx on both sides shows the recurrence for bnb_{n}. Analogue method works for gng_{n} (see remark below). ∎

Remark 4.5.

In the above proof, certificate of creative telescoping777it refers to the complicated rational function inside dd​x​(⋯​⋯​F​(x,n))\frac{d}{dx}\left(\cdots\cdots F(x,n)\right). is obtained with help of Mathematica package HolonomicFunctions developed by C. Koutschan, [13]. One simply executes the following command:

CreativeTelescoping[((1+x)(1+2x))^((n-1)/3)*x^(-n-1),Der[x],{S[n]}] // Factor

For gng_{n}, one make corresponding modification, this certificate is too long to be displayed here.

For a generic f​(x)f(x) of degree dd, residue of ΞΆf​(s)\zeta_{f}(s) at s=s0s=s_{0} allowed by Theorem 4.1 is in general non-zero. Witten zeta functions display here another unexpected property: residues at many more points turn out to be zero, ΞΆf​(s)\zeta_{f}(s) is thus analytic there.

Theorem 4.6.

(a) Ο‰B​(s)\omega_{B}(s) has a pole at s=1/2s=1/2 and also at s=k3s=\frac{k}{3} with k≑1,5(mod6)k\equiv 1,5\pmod{6} and k≀1k\leq 1.

(b) Ο‰G​(s)\omega_{G}(s) has a pole at s=1/3s=1/3 and also (possibly) at s=k5s=\frac{k}{5} with k≑1,3,7,9(mod10)k\equiv 1,3,7,9\pmod{10} and k≀1k\leq 1.

In both cases, all poles are simple, these are only poles of Ο‰B​(s)\omega_{B}(s) and Ο‰G​(s)\omega_{G}(s).

Proof.

Poles of Ο‰B​(s)\omega_{B}(s) at s=1/2s=1/2 and of Ο‰G​(s)\omega_{G}(s) at s=1/3s=1/3 comes from the abscissa of convergence.

For remaining poles, we concentrate on Ο‰G\omega_{G} since the following argument for Ο‰B\omega_{B} is similar. For k≑0,5(mod10)k\equiv 0,5\pmod{10}, Ο‰G​(s)\omega_{G}(s) is analytic at s=k/5s=k/5 because k/5k/5 is an integer. To show Ο‰G​(s)\omega_{G}(s) is also analytic at s=k/5s=k/5 for k≑2,4,6,8(mod10)k\equiv 2,4,6,8\pmod{10}, it suffices to show gn=0g_{n}=0 for n≑3,5,7,9(mod10)n\equiv 3,5,7,9\pmod{10}. This can be easily checked from the recurrence: for example, one computes g3=g13=0g_{3}=g_{13}=0, the recurrence implies g3+10​n=0g_{3+10n}=0. ∎

Ο‰B​(s)\omega_{B}(s) has a pole of exact order 11 at s=k3s=\frac{k}{3} for k≀1,k≑1,5(mod6)k\leq 1,k\equiv 1,5\pmod{6}: this follows from the non-vanishing of bnb_{n} at n≑0,2(mod6)n\equiv 0,2\pmod{6}. For Ο‰G\omega_{G}, non-vanishing of gng_{n} at n≑0,2,4,8(mod10)n\equiv 0,2,4,8\pmod{10} implies we can remove the word "possibly" in theorem above. This should not be difficult to prove by deriving an asymptotic of corresponding second order recurrence (PoincarΓ©-Perron theorem [18]).

Conjecture 4.7.

Consider above defined sequence {gn}nβ‰₯0\{g_{n}\}_{n\geq 0}, then gnβ‰ 0g_{n}\neq 0 when n≑0,2,4,8(mod10)n\equiv 0,2,4,8\pmod{10}.

Results of this section concerning Ο‰B​(s)\omega_{B}(s) are already obtained by Bringmann et al. in [4]. Ο‰G​(s)\omega_{G}(s) having possible poles at s=k/5s=k/5 was already known in [9], however, the fact that half of such kk actually does not give poles seems unknown until now.

5. Derivative at s=0s=0

In this section, we show how to evaluate ΞΆf′​(0)\zeta_{f}^{\prime}(0) when all roots of ff are rational. We will be brief because ΞΆf′​(0)\zeta_{f}^{\prime}(0) was already evaluated in [20], using a different approach.

As a preliminary, we need to evaluate the following integral

ℐ​(Ξ±)=12​π​iβ€‹βˆ«βˆ’3/2βˆ’iβ€‹βˆžβˆ’3/2+iβ€‹βˆžΞ“β€‹(z)​΢​(z)β€‹Ξ±βˆ’z​Γ​(βˆ’z)​΢​(βˆ’z)​𝑑zΞ±>0\mathcal{I}(\alpha)=\frac{1}{2\pi i}\int_{-3/2-i\infty}^{-3/2+i\infty}\Gamma(z)\zeta(z)\alpha^{-z}\Gamma(-z)\zeta(-z)dz\qquad\alpha>0
Proposition 5.1.

Let Ξ±=p/q\alpha=p/q, with p,qβˆˆβ„•p,q\in\mathbb{N} coprime. Then ℐ​(Ξ±)\mathcal{I}(\alpha) is a β„š\mathbb{Q}-linear combination of following constants:

1,Ξ³,΢′​(0),΢′​(βˆ’1),log⁑p,log⁑q,log⁑Γ​(ip)​ with ​1≀i≀pβˆ’1,log⁑Γ​(iq)​ with ​1≀i≀qβˆ’11,\quad\gamma,\quad\zeta^{\prime}(0),\quad\zeta^{\prime}(-1),\quad\log p,\quad\log q,\quad\log\Gamma(\frac{i}{p})\text{ with }1\leq i\leq p-1,\quad\log\Gamma(\frac{i}{q})\text{ with }1\leq i\leq q-1

Here we point out that ΢′​(0)=βˆ’log⁑(2​π)/2\zeta^{\prime}(0)=-\log(2\pi)/2 which a well-known conclusion. We now prove the Proposition. Denote

(5.1) ℐ​(s,Ξ±)=1Γ​(s)​12​π​iβ€‹βˆ«3/2βˆ’iβ€‹βˆž3/2+iβ€‹βˆžΞ“β€‹(z)​΢​(s)β€‹Ξ±βˆ’z​Γ​(sβˆ’z)​΢​(sβˆ’z)​𝑑zβ„œβ‘(s)>5/2\mathcal{I}(s,\alpha)=\frac{1}{\Gamma(s)}\frac{1}{2\pi i}\int_{3/2-i\infty}^{3/2+i\infty}\Gamma(z)\zeta(s)\alpha^{-z}\Gamma(s-z)\zeta(s-z)dz\qquad\Re(s)>5/2
Lemma 5.2.

When β„œβ‘(s)>5/2\Re(s)>5/2, we have

ℐ​(s,Ξ±)=1Γ​(s)β€‹βˆ«0∞xsβˆ’1(eα​xβˆ’1)​(exβˆ’1)​𝑑x\mathcal{I}(s,\alpha)=\frac{1}{\Gamma(s)}\int_{0}^{\infty}\frac{x^{s-1}}{(e^{\alpha x}-1)(e^{x}-1)}dx
Proof.

This is a simple consequence of Parseval identity: if

Fi​(s)=∫0∞xsβˆ’1​fi​(x)​𝑑xΞ±i<β„œβ‘(s)<Ξ²iF_{i}(s)=\int_{0}^{\infty}x^{s-1}f_{i}(x)dx\qquad\alpha_{i}<\Re(s)<\beta_{i}

then

∫0∞xsβˆ’1​f1​(x)​f2​(x)​𝑑x=12​π​iβ€‹βˆ«cβˆ’βˆžc+∞F1​(z)​F2​(sβˆ’z)​𝑑zΞ±2+c<β„œβ‘(s)<Ξ²2+c,Ξ±1<c<Ξ²1\int_{0}^{\infty}x^{s-1}f_{1}(x)f_{2}(x)dx=\frac{1}{2\pi i}\int_{c-\infty}^{c+\infty}F_{1}(z)F_{2}(s-z)dz\qquad\alpha_{2}+c<\Re(s)<\beta_{2}+c,\alpha_{1}<c<\beta_{1}

Specialize this to f1​(x)=1exβˆ’1,f2​(x)=1eα​xβˆ’1f_{1}(x)=\frac{1}{e^{x}-1},f_{2}(x)=\frac{1}{e^{\alpha x}-1}, with corresponding F1​(s)=Γ​(s)​΢​(s)F_{1}(s)=\Gamma(s)\zeta(s) and F2​(s)=Ξ±βˆ’s​Γ​(s)​΢​(s)F_{2}(s)=\alpha^{-s}\Gamma(s)\zeta(s). ∎

Recall polylogarithm and Hurwitz zeta function:

Lis⁑(z)=βˆ‘nβ‰₯1znns,΢​(s,a)=βˆ‘nβ‰₯01(n+a)s\operatorname{Li}_{s}(z)=\sum_{n\geq 1}\frac{z^{n}}{n^{s}},\qquad\zeta(s,a)=\sum_{n\geq 0}\frac{1}{(n+a)^{s}}
Lemma 5.3.

When ΞΌβ‰ 1\mu\neq 1 is an NN-th root of unity,

dd​s|s=0​Lis⁑(ΞΌ)=ΞΌΞΌβˆ’1​log⁑N+βˆ‘j=1Nβˆ’1ΞΌj​log⁑Γ​(jN)\left.\frac{d}{ds}\right|_{s=0}\operatorname{Li}_{s}(\mu)=\frac{\mu}{\mu-1}\log N+\sum_{j=1}^{N-1}\mu^{j}\log\Gamma(\frac{j}{N})
Proof.

For β„œβ‘(s)>1\Re(s)>1, we have

Lis⁑(ΞΌ)=βˆ‘nβ‰₯1ΞΌnns=βˆ‘i=0βˆžβˆ‘j=1NΞΌj(j+i​N)s=Nβˆ’sβ€‹βˆ‘j=1NΞΌj​΢​(s,jN),\operatorname{Li}_{s}(\mu)=\sum_{n\geq 1}\frac{\mu^{n}}{n^{s}}=\sum_{i=0}^{\infty}\sum_{j=1}^{N}\frac{\mu^{j}}{(j+iN)^{s}}=N^{-s}\sum_{j=1}^{N}\mu^{j}\zeta(s,\frac{j}{N}),

Recall following two classical identities: ([21])

΢​(0,a)=12βˆ’a,dd​s|s=0​΢​(s,a)=log⁑Γ​(a)βˆ’log⁑(2​π)2.\zeta(0,a)=\frac{1}{2}-a,\qquad\left.\frac{d}{ds}\right|_{s=0}\zeta(s,a)=\log\Gamma(a)-\frac{\log(2\pi)}{2}.

Differentiate initial displayed equation at s=0s=0 and plug in these two identities on ΢​(s,a)\zeta(s,a) completes the proof. ∎

Let Ξ±=p/q\alpha=p/q be a positive rational number indicated in Proposition, and let

qβˆ’s​ℐ​(s,Ξ±)=c0​(p,q)+c1​(p,q)​s+O​(s2)sβ†’0q^{-s}\mathcal{I}(s,\alpha)=c_{0}(p,q)+c_{1}(p,q)s+O(s^{2})\qquad s\to 0
Lemma 5.4.

We have

c0​(p,q)=14+112​(pq+qp).c_{0}(p,q)=\frac{1}{4}+\frac{1}{12}(\frac{p}{q}+\frac{q}{p}).

and

c1​(p,q)=΢′​(βˆ’1)p​qβˆ’p+q2​p​q​΢′​(0)+log⁑qq​rq​(p)+log⁑pp​rp​(q)+1qβ€‹βˆ‘j=1qsq,p​(j)​log⁑Γ​(jq)+1pβ€‹βˆ‘j=1psp,q​(j)​log⁑Γ​(jp)c_{1}(p,q)=\frac{\zeta^{\prime}(-1)}{pq}-\frac{p+q}{2pq}\zeta^{\prime}(0)+\frac{\log q}{q}r_{q}(p)+\frac{\log p}{p}r_{p}(q)+\frac{1}{q}\sum_{j=1}^{q}s_{q,p}(j)\log\Gamma\left(\frac{j}{q}\right)+\frac{1}{p}\sum_{j=1}^{p}s_{p,q}(j)\log\Gamma\left(\frac{j}{p}\right)

where

sq,p​(j)=βˆ’qβˆ’12+(the unique ​0≀i≀qβˆ’1​ such that ​i​p+j≑0(modq))s_{q,p}(j)=-\frac{q-1}{2}+\left(\text{the unique }0\leq i\leq q-1\text{ such that }ip+j\equiv 0\pmod{q}\right)

and

rq​(p)=1qβ€‹βˆ‘i=1qβˆ’1i​sq,p​(i)r_{q}(p)=\frac{1}{q}\sum_{i=1}^{q-1}is_{q,p}(i)
Proof.
qβˆ’s​ℐ​(s,Ξ±)=1Γ​(s)β€‹βˆ«0∞xsβˆ’1(ep​xβˆ’1)​(eq​xβˆ’1)​𝑑xq^{-s}\mathcal{I}(s,\alpha)=\frac{1}{\Gamma(s)}\int_{0}^{\infty}\frac{x^{s-1}}{(e^{px}-1)(e^{qx}-1)}dx

c0​(p,q)c_{0}(p,q) is simply residue of 1x​(ep​xβˆ’1)​(eq​xβˆ’1)\frac{1}{x(e^{px}-1)(e^{qx}-1)} at x=0x=0. c1​(p,q)c_{1}(p,q) requires a more elaborate computation. Partial fraction decomposition

1(xpβˆ’1)​(xqβˆ’1)=1p​q​1(xβˆ’1)2+2βˆ’pβˆ’q2​p​q​1(xβˆ’1)+βˆ‘Ξ½β‰ 1,Ξ½q=1Ξ½q​(Ξ½pβˆ’1)​1xβˆ’Ξ½+βˆ‘ΞΌ:ΞΌβ‰ 1,ΞΌp=1ΞΌp​(ΞΌqβˆ’1)​1xβˆ’ΞΌ\frac{1}{(x^{p}-1)(x^{q}-1)}=\frac{1}{pq}\frac{1}{(x-1)^{2}}+\frac{2-p-q}{2pq}\frac{1}{(x-1)}+\sum_{\nu\neq 1,\nu^{q}=1}\frac{\nu}{q(\nu^{p}-1)}\frac{1}{x-\nu}+\sum_{\mu:\mu\neq 1,\mu^{p}=1}\frac{\mu}{p(\mu^{q}-1)}\frac{1}{x-\mu}

together with Mellin transforms

∫0∞xsβˆ’1​d​x(exβˆ’1)2=(΢​(sβˆ’1)βˆ’ΞΆβ€‹(s))​Γ​(s),∫0∞xsβˆ’1​d​xexβˆ’a=Γ​(s)​Lis⁑(a)a\int_{0}^{\infty}\frac{x^{s-1}dx}{(e^{x}-1)^{2}}=(\zeta(s-1)-\zeta(s))\Gamma(s),\quad\int_{0}^{\infty}\frac{x^{s-1}dx}{e^{x}-a}=\frac{\Gamma(s)\operatorname{Li}_{s}(a)}{a}

produce

qβˆ’s​ℐ​(s,Ξ±)=1p​q​(΢​(sβˆ’1)βˆ’ΞΆβ€‹(s))+2βˆ’pβˆ’q2​p​q​΢​(s)+βˆ‘Ξ½β‰ 1,Ξ½q=1Lis⁑(Ξ½)q​(Ξ½pβˆ’1)+βˆ‘ΞΌ:ΞΌβ‰ 1,ΞΌp=1Lis⁑(ΞΌ)p​(ΞΌqβˆ’1)q^{-s}\mathcal{I}(s,\alpha)=\frac{1}{pq}(\zeta(s-1)-\zeta(s))+\frac{2-p-q}{2pq}\zeta(s)+\sum_{\nu\neq 1,\nu^{q}=1}\frac{\operatorname{Li}_{s}(\nu)}{q(\nu^{p}-1)}+\sum_{\mu:\mu\neq 1,\mu^{p}=1}\frac{\operatorname{Li}_{s}(\mu)}{p(\mu^{q}-1)}

From Lemma above, we obtain the expression of c1​(p,q)c_{1}(p,q) as in statement with

rq​(p)=βˆ‘ΞΌβ‰ 1,ΞΌq=1ΞΌ(ΞΌβˆ’1)​(ΞΌpβˆ’1),sq,p​(j)=βˆ‘ΞΌβ‰ 1,ΞΌq=1ΞΌjΞΌpβˆ’1r_{q}(p)=\sum_{\mu\neq 1,\mu^{q}=1}\frac{\mu}{(\mu-1)(\mu^{p}-1)},\qquad s_{q,p}(j)=\sum_{\mu\neq 1,\mu^{q}=1}\frac{\mu^{j}}{\mu^{p}-1}

here the sum is over all qq-th roots of unity that is not 11. Simplification of these two trigonometric sums gives the expression in statement. ∎

Therefore

(5.2) ℐ​(s,Ξ±)=c0​(p,q)+[c0​(p,q)​log⁑q+c1​(p,q)]​s+O​(s2).\mathcal{I}(s,\alpha)=c_{0}(p,q)+[c_{0}(p,q)\log q+c_{1}(p,q)]s+O(s^{2}).

The original representation ℐ​(s,Ξ±)\mathcal{I}(s,\alpha) in (5.1) is valid only for β„œβ‘(s)>5/2\Re(s)>5/2, in order to get a representation that is valid at s=0s=0, we shift the contour from β„œβ‘(z)=3/2\Re(z)=3/2 to β„œβ‘(z)=βˆ’3/2\Re(z)=-3/2, picking up residues at z=1,0,βˆ’1z=1,0,-1 to yield:

ℐ​(s,Ξ±)=΢​(sβˆ’1)α​(sβˆ’1)+112​(α​s​΢​(s+1)βˆ’6​΢​(s))+1Γ​(s)​12​π​iβ€‹βˆ«βˆ’3/2βˆ’iβ€‹βˆžβˆ’3/2+iβ€‹βˆžΞ“β€‹(z)​΢​(s)β€‹Ξ±βˆ’z​Γ​(sβˆ’z)​΢​(sβˆ’z)​𝑑z.\mathcal{I}(s,\alpha)=\frac{\zeta(s-1)}{\alpha(s-1)}+\frac{1}{12}(\alpha s\zeta(s+1)-6\zeta(s))+\frac{1}{\Gamma(s)}\frac{1}{2\pi i}\int_{-3/2-i\infty}^{-3/2+i\infty}\Gamma(z)\zeta(s)\alpha^{-z}\Gamma(s-z)\zeta(s-z)dz.

This is now valid at s=0s=0, its ss coefficient is

γ​α2+3​α​log⁑(2​π)+1βˆ’12​΢′​(βˆ’1)12​α+ℐ​(Ξ±)\frac{\gamma\alpha^{2}+3\alpha\log(2\pi)+1-12\zeta^{\prime}(-1)}{12\alpha}+\mathcal{I}(\alpha)

Comparing with (5.2) finally gives an explicit expression of ℐ​(Ξ±)\mathcal{I}(\alpha):

ℐ​(Ξ±)=c0​(p,q)​log⁑q+c1​(p,q)βˆ’Ξ³β€‹Ξ±2+3​α​log⁑(2​π)+1βˆ’12​΢′​(βˆ’1)12​α\mathcal{I}(\alpha)=c_{0}(p,q)\log q+c_{1}(p,q)-\frac{\gamma\alpha^{2}+3\alpha\log(2\pi)+1-12\zeta^{\prime}(-1)}{12\alpha}

here ci​(p,q)c_{i}(p,q) are computed according to formula in Lemma above.

For example:

ℐ​(1)\displaystyle\mathcal{I}(1) =2​΢′​(βˆ’1)βˆ’Ξ³12βˆ’112+14​log⁑(2​π);\displaystyle=2\zeta^{\prime}(-1)-\frac{\gamma}{12}-\frac{1}{12}+\frac{1}{4}\log(2\pi);
ℐ​(12)\displaystyle\mathcal{I}(\frac{1}{2}) =5​΢′​(βˆ’1)2βˆ’16βˆ’Ξ³24+11​log⁑224+log⁑π4;\displaystyle=\frac{5\zeta^{\prime}(-1)}{2}-\frac{1}{6}-\frac{\gamma}{24}+\frac{11\log 2}{24}+\frac{\log\pi}{4};
ℐ​(13)\displaystyle\mathcal{I}(\frac{1}{3}) =103​΢′​(βˆ’1)βˆ’14βˆ’Ξ³36+11​log⁑336+112​log⁑(2​π)+13​log⁑Γ​(13);\displaystyle=\frac{10}{3}\zeta^{\prime}(-1)-\frac{1}{4}-\frac{\gamma}{36}+\frac{11\log 3}{36}+\frac{1}{12}\log(2\pi)+\frac{1}{3}\log\Gamma\left(\frac{1}{3}\right);
ℐ​(32)\displaystyle\mathcal{I}(\frac{3}{2}) =56​΢′​(βˆ’1)βˆ’Ξ³8βˆ’118+19​log⁑272+log⁑π12βˆ’log⁑39+13​log⁑Γ​(23).\displaystyle=\frac{5}{6}\zeta^{\prime}(-1)-\frac{\gamma}{8}-\frac{1}{18}+\frac{19\log 2}{72}+\frac{\log\pi}{12}-\frac{\log 3}{9}+\frac{1}{3}\log\Gamma\left(\frac{2}{3}\right).

Now we go back to the problem of evaluating ΞΆf′​(0)\zeta_{f}^{\prime}(0). Let

f​(x)=c0​(1+Ξ±1​x)​⋯​(1+Ξ±d​x)=c0+c1​x+β‹―+cd​xd,f(x)=c_{0}(1+\alpha_{1}x)\cdots(1+\alpha_{d}x)=c_{0}+c_{1}x+\cdots+c_{d}x^{d},

recall that Kf​(0,z)=βˆ‘iΞ±iβˆ’zK_{f}(0,z)=\sum_{i}\alpha_{i}^{-z} (Proposition 1.5). From (2.1), the following representation is valid at a neighbourhood of s=0s=0:

ΞΆf​(s)=Γ​(1βˆ’s)​Γ​(d​s+sβˆ’1)Γ​(s)​΢​((d+2)​sβˆ’1)​Kf​(s,1βˆ’s)+βˆ‘i=01΢​(sβˆ’i)​΢​((d+1)​s+i)​[f​(x)βˆ’s]​[xi]+1Γ​(s)​12​π​iβ€‹βˆ«βˆ’3/2βˆ’iβ€‹βˆžβˆ’3/2+iβ€‹βˆžΞ“β€‹(z)​Γ​(s​dβˆ’z)​Kf​(s,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z\zeta_{f}(s)=\frac{\Gamma(1-s)\Gamma(ds+s-1)}{\Gamma(s)}\zeta((d+2)s-1)K_{f}(s,1-s)+\sum_{i=0}^{1}\zeta(s-i)\zeta((d+1)s+i)[f(x)^{-s}][x^{i}]\\ +\frac{1}{\Gamma(s)}\frac{1}{2\pi i}\int_{-3/2-i\infty}^{-3/2+i\infty}\Gamma(z)\Gamma(sd-z)K_{f}(s,z)\zeta(s+z)\zeta((d+1)s-z)dz

Let AfA_{f} be sum of coefficients of ss in Taylor expansion of first three terms in the front, then

ΞΆf′​(0)\displaystyle\zeta_{f}^{\prime}(0) =Af+12​π​iβ€‹βˆ«βˆ’3/2βˆ’iβ€‹βˆžβˆ’3/2+iβ€‹βˆžΞ“β€‹(z)​Γ​(s​dβˆ’z)​Kf​(0,z)​΢​(s+z)​΢​((d+1)​sβˆ’z)​𝑑z\displaystyle=A_{f}+\frac{1}{2\pi i}\int_{-3/2-i\infty}^{-3/2+i\infty}\Gamma(z)\Gamma(sd-z)K_{f}(0,z)\zeta(s+z)\zeta((d+1)s-z)dz
=Af+βˆ‘i=1dℐ​(Ξ±i),\displaystyle=A_{f}+\sum_{i=1}^{d}\mathcal{I}(\alpha_{i}),

with same ℐ​(Ξ±)\mathcal{I}(\alpha) as before. Observe

βˆ‘i=01΢​(sβˆ’i)​΢​((d+1)​s+i)​[f​(x)βˆ’s]​[xi]=΢​(s)​΢​((d+1)​s)​c0βˆ’s+΢​(sβˆ’1)​΢​((d+1)​s+1)​(βˆ’s​c0βˆ’sβˆ’1​c1);\sum_{i=0}^{1}\zeta(s-i)\zeta((d+1)s+i)[f(x)^{-s}][x^{i}]=\zeta(s)\zeta((d+1)s)c_{0}^{-s}+\zeta(s-1)\zeta((d+1)s+1)\left(-sc_{0}^{-s-1}c_{1}\right);

Proposition 1.6 enables us to expand Kf​(s,1βˆ’s)K_{f}(s,1-s) upto O​(s2)O(s^{2}), so AfA_{f} can be found. When all Ξ±iβˆˆβ„š\alpha_{i}\in\mathbb{Q}, we showed previously how to evaluate ℐ​(Ξ±i)\mathcal{I}(\alpha_{i}). This completes our description on how to evaluate ΞΆf′​(0)\zeta^{\prime}_{f}(0).

Example 5.5.

Let f​(x)=(1+x)​(1+3​x)f(x)=(1+x)(1+3x),

ΞΆf​(s)=βˆ‘n,mβ‰₯11ns​ms​(n+m)s​(n+3​m)s\zeta_{f}(s)=\sum_{n,m\geq 1}\frac{1}{n^{s}m^{s}(n+m)^{s}(n+3m)^{s}}

an explicit calculation gives

ΞΆf′​(0)=βˆ’49​΢′​(βˆ’1)βˆ’25108​log⁑3+43​log⁑(2​π)+13​log⁑Γ​(13)\zeta_{f}^{\prime}(0)=-\frac{4}{9}\zeta^{\prime}(-1)-\frac{25}{108}\log 3+\frac{4}{3}\log(2\pi)+\frac{1}{3}\log\Gamma(\frac{1}{3})

In view of occurrence of ΢′​(βˆ’1)\zeta^{\prime}(-1) and log⁑Γ​(1/3)\log\Gamma(1/3) in this generic example, simplicity of following derivatives might be surprising:

Theorem 5.6.
Ο‰A′​(0)\displaystyle\omega_{A}^{\prime}(0) =log⁑2+log⁑π\displaystyle=\log 2+\log\pi
Ο‰B′​(0)\displaystyle\omega_{B}^{\prime}(0) =54​log⁑2+32​log⁑π\displaystyle=\frac{5}{4}\log 2+\frac{3}{2}\log\pi
Ο‰G′​(0)\displaystyle\omega_{G}^{\prime}(0) =2​log⁑2βˆ’12​log⁑3+52​log⁑π\displaystyle=2\log 2-\frac{1}{2}\log 3+\frac{5}{2}\log\pi

The author does not know a conceptual explanation for their simplicities, the terms ΢′​(βˆ’1)\zeta^{\prime}(-1) or log⁑Γ​(r)\log\Gamma(r) end up cancelling each other during the computation888For Ο‰G\omega_{G}, one also needs relation Γ​(1/3)​Γ​(2/3)=2​π/3\Gamma(1/3)\Gamma(2/3)=2\pi/\sqrt{3}., analogous miraculous cancellation also takes place when one uses the approach in [20]. The case of Ο‰A\omega_{A} were also proved in [2], [3] and [17].

6. Number of representations for Lie algebra G2G_{2}

Let 𝔀\mathfrak{g} be a finite dimensional simple Lie algebra over β„‚\mathbb{C}, denote r𝔀​(n)r_{\mathfrak{g}}(n) to be number of non-isomorphic representation of 𝔀\mathfrak{g} with dimension nn. We abuse notation to write r𝔀​(n)r_{\mathfrak{g}}(n) as rΦ​(n)r_{\Phi}(n) if Ξ¦\Phi is corresponding root system. For Ξ¦=A2,B2,G2\Phi=A_{2},B_{2},G_{2},

βˆ‘nβ‰₯0rA2​(n)​qn\displaystyle\sum_{n\geq 0}r_{A_{2}}(n)q^{n} =∏n,mβ‰₯1(1βˆ’qm​n​(m+n)/2)βˆ’1;\displaystyle=\prod_{n,m\geq 1}(1-q^{mn(m+n)/2})^{-1};
βˆ‘nβ‰₯0rB2​(n)​qn\displaystyle\sum_{n\geq 0}r_{B_{2}}(n)q^{n} =∏n,mβ‰₯1(1βˆ’qm​n​(m+n)​(m+2​n)/6)βˆ’1;\displaystyle=\prod_{n,m\geq 1}(1-q^{mn(m+n)(m+2n)/6})^{-1};
βˆ‘nβ‰₯0rG2​(n)​qn\displaystyle\sum_{n\geq 0}r_{G_{2}}(n)q^{n} =∏n,mβ‰₯1(1βˆ’qm​n​(m+n)​(m+2​n)​(m+3​n)​(2​m+3​n)/120)βˆ’1.\displaystyle=\prod_{n,m\geq 1}(1-q^{mn(m+n)(m+2n)(m+3n)(2m+3n)/120})^{-1}.

Circle method, adapted in [6] and [4], can be used to find asymptotic of r𝔀​(n)r_{\mathfrak{g}}(n) provided we have enough information on poles location and certain special values of ΢𝔀​(s)\zeta_{\mathfrak{g}}(s).

Asymptotic of rA2​(n),rB2​(n)r_{A_{2}}(n),r_{B_{2}}(n) has been found previously ([19], [5], [4]):

rA2​(n)\displaystyle r_{A_{2}}(n) ∼C1n3/5​exp⁑(A1​n2/5+A2​n3/10+A3​n1/5+A4​n1/10)\displaystyle\sim\frac{C_{1}}{n^{3/5}}\exp\left(A_{1}n^{2/5}+A_{2}n^{3/10}+A_{3}n^{1/5}+A_{4}n^{1/10}\right)
rB2​(n)\displaystyle r_{B_{2}}(n) ∼C2n7/12​exp⁑(B1​n1/3+B2​n2/9+B3​n1/9+B4)\displaystyle\sim\frac{C_{2}}{n^{7/12}}\exp\left(B_{1}n^{1/3}+B_{2}n^{2/9}+B_{3}n^{1/9}+B_{4}\right)

for explicit constants C1,C2,Ai,BiC_{1},C_{2},A_{i},B_{i}. Our gathered information on ΞΆG2​(s)\zeta_{G_{2}}(s) enables us to do the same for G2G_{2} by applying a formula given in Theorem 4.4 of [4]. Some technical hypotheses have to be satisfied (e.g. poles being simple and moderate growth along imaginary direction), they are met for ΞΆA2​(s),ΞΆB2​(s),ΞΆG2​(s)\zeta_{A_{2}}(s),\zeta_{B_{2}}(s),\zeta_{G_{2}}(s) (Theorems 1.8, 4.1).

For ΞΆG2​(s)\zeta_{G_{2}}(s), these hypotheses have been checked in Rutard [20], and we have

rG2​(n)∼C3n9/16​exp⁑(G1​n1/4+G2​n3/20+G3​n1/20),nβ†’βˆž,r_{G_{2}}(n)\sim\frac{C_{3}}{n^{9/16}}\exp\left(G_{1}n^{1/4}+G_{2}n^{3/20}+G_{3}n^{1/20}\right),\qquad n\to\infty,

for some constants C3,G1,G2,G3C_{3},G_{1},G_{2},G_{3}. However, in [20], residues at s=1/5,1/3s=1/5,1/3 remains unevaluated, plugging in these residues, which we figured out in this article, we obtain values of these constants:

Theorem 6.1.
rG2​(n)∼C3n9/16​exp⁑(G1​n1/4+G2​n3/20+G3​n1/20),nβ†’βˆžr_{G_{2}}(n)\sim\frac{C_{3}}{n^{9/16}}\exp\left(G_{1}n^{1/4}+G_{2}n^{3/20}+G_{3}n^{1/20}\right),\qquad n\to\infty

with

C3=279/48​39/32​57/16​π31/16​΢​(43)1/16​Γ​(13)1/4,G1=51/4​Γ​(13)3313/8​(2​΢​(4/3)Ο€)3/4,C_{3}=2^{79/48}3^{9/32}5^{7/16}\pi^{31/16}\zeta\left(\frac{4}{3}\right)^{1/16}\Gamma\left(\frac{1}{3}\right)^{1/4},\qquad G_{1}=\frac{5^{1/4}\Gamma\left(\frac{1}{3}\right)^{3}}{3^{13/8}}\left(\frac{2\zeta(4/3)}{\pi}\right)^{3/4},
G2=213/20​31/8​(1+32/5)​π3/20​΢​(15)​΢​(65)​Γ​(15)517/20​΢​(43)3/20​Γ​(13)3/5,G3=323/8​(1+32/5)2​΢​(15)2​΢​(65)2​Γ​(15)229/20​579/20​Γ​(13)21/5​(π΢​(43))21/20G_{2}=\frac{2^{13/20}3^{1/8}(1+3^{2/5})\pi^{3/20}\zeta(\frac{1}{5})\zeta(\frac{6}{5})\Gamma\left(\frac{1}{5}\right)}{5^{17/20}\zeta(\frac{4}{3})^{3/20}\Gamma\left(\frac{1}{3}\right)^{3/5}},\qquad G_{3}=\frac{3^{23/8}(1+3^{2/5})^{2}\zeta\left(\frac{1}{5}\right)^{2}\zeta\left(\frac{6}{5}\right)^{2}\Gamma\left(\frac{1}{5}\right)^{2}}{2^{9/20}5^{79/20}\Gamma\left(\frac{1}{3}\right)^{21/5}}\left(\frac{\pi}{\zeta\left(\frac{4}{3}\right)}\right)^{21/20}

7. Appendix: evaluation of a definite integral

Here we prove the evaluation

∫0∞d​xx​(1+x)​(1+2​x)​(1+3​x)​(2+3​x)3=125/3​31/2​π​Γ​(13)3,\int_{0}^{\infty}\frac{dx}{\sqrt[3]{x(1+x)(1+2x)(1+3x)(2+3x)}}=\frac{1}{2^{5/3}3^{1/2}\pi}\Gamma\left(\frac{1}{3}\right)^{3},

which is related to residue of Ο‰G​(s)\omega_{G}(s) at the point s=1/3s=1/3. We remark this evaluation seems highly non-trivial: algebraic curve y3=x​(1+x)​(1+2​x)​(1+3​x)​(2+3​x)y^{3}=x(1+x)(1+2x)(1+3x)(2+3x) being genus 44. The fact it can be expressed in terms of gamma function at all indicates deep properties of Ο‰G​(s)\omega_{G}(s).

We shall give an indirect proof, it is following Lemma specialized at u=βˆ’1/6u=-1/6.

Lemma 7.1.

For βˆ’1/2<β„œβ‘(u)<1/6-1/2<\Re(u)<1/6, we have

∫0∞(x​(1+x)​(1+2​x))2​u​((1+3​x)​(2+3​x))βˆ’2​uβˆ’2/3​𝑑x=3βˆ’3​uβˆ’3/2​Γ​(16βˆ’u)​Γ​(u+12)22/3​Γ​(23).\int_{0}^{\infty}(x(1+x)(1+2x))^{2u}((1+3x)(2+3x))^{-2u-2/3}dx=\frac{3^{-3u-3/2}\Gamma\left(\frac{1}{6}-u\right)\Gamma\left(u+\frac{1}{2}\right)}{2^{2/3}\Gamma\left(\frac{2}{3}\right)}.
Proof.

The key idea is to do a creative telescoping. Let F​(x,u)F(x,u) be the integrand,

I1​(u):=∫0∞F​(x,u)​𝑑x,I2​(u):=∫CF​(x,u)​𝑑x=(e4​π​i​uβˆ’1)​I1​(u),I_{1}(u):=\int_{0}^{\infty}F(x,u)dx,\qquad I_{2}(u):=\int_{C}F(x,u)dx=(e^{4\pi iu}-1)I_{1}(u),

here CC is the contour wrapping around positive real axis mentioned in the proof of Theorem 1.1. Note that I1​(u)I_{1}(u) is analytic on βˆ’1/2<β„œβ‘(u)<1/6-1/2<\Re(u)<1/6 and I2​(u)I_{2}(u) on β„œβ‘(u)<1/6\Re(u)<1/6. Creative telescoping produces the following relation

(2​u+1)​F​(x,u)+9​(6​u+5)​F​(x,u+1)=βˆ‚βˆ‚x​(2​x​(x+1)​(2​x+1)​(3​x2+3​x+1)(3​x+1)​(3​x+2)​F​(x,u)).(2u+1)F(x,u)+9(6u+5)F(x,u+1)=\frac{\partial}{\partial x}\left(\frac{2x(x+1)(2x+1)\left(3x^{2}+3x+1\right)}{(3x+1)(3x+2)}F(x,u)\right).

Integrate with respect to xx along CC, fundamental theorem of calculus says integral on the RHS equals 0, so we have

(2​u+1)​I2​(u)+9​(6​u+5)​I2​(u+1)=0.(2u+1)I_{2}(u)+9(6u+5)I_{2}(u+1)=0.

This recurrence amounts to say that

(7.1) J​(u)=33​uΓ​(16βˆ’u)​Γ​(12+u)​(1βˆ’e4​π​i​u)​I1​(u)J(u)=\frac{3^{3u}}{\Gamma\left(\frac{1}{6}-u\right)\Gamma\left(\frac{1}{2}+u\right)}(1-e^{4\pi iu})I_{1}(u)

is a function of period 11: J​(u+1)=J​(u)J(u+1)=J(u).

Since (1βˆ’e4​π​i​u)​I1​(u)=I2​(u)(1-e^{4\pi iu})I_{1}(u)=I_{2}(u) is analytic on β„œβ‘(u)<1/6\Re(u)<1/6, J​(u)J(u) is entire. Note that for any Ξ΅>0\varepsilon>0, yy a large positive real number, |Γ​(aΒ±i​y)|β‰₯O​(e(βˆ’Ο€/2βˆ’Ξ΅)​|y|)|\Gamma(a\pm iy)|\geq O(e^{(-\pi/2-\varepsilon)|y|}) and |I1​(Β±i​y)|=O​(eΡ​|y|)|I_{1}(\pm iy)|=O(e^{\varepsilon|y|}). We have

J​(i​y)=O​(e(Ο€+Ξ΅)​y),J​(βˆ’i​y)=O​(e(5​π+Ξ΅)​y),yβ†’+∞.J(iy)=O(e^{(\pi+\varepsilon)y}),\qquad J(-iy)=O(e^{(5\pi+\varepsilon)y}),\qquad y\to+\infty.

These two growth conditions, together with J​(u+1)=J​(u)J(u+1)=J(u) imply J​(u)=a+b​e2​π​i​u+c​e4​π​i​uJ(u)=a+be^{2\pi iu}+ce^{4\pi iu} for some a,b,cβˆˆβ„‚a,b,c\in\mathbb{C}.

In (7.1), set u=0u=0 gives J​(0)=0⇔a+b+c=0J(0)=0\iff a+b+c=0. Also, since I1​(u)I_{1}(u) has a simple pole at u=βˆ’1/2u=-1/2, and 1βˆ’e4​π​i​uΓ​(12+u)\frac{1-e^{4\pi iu}}{\Gamma(\frac{1}{2}+u)} has double zero at u=βˆ’1/2u=-1/2, J​(βˆ’1/2)=0⇔aβˆ’b+c=0J(-1/2)=0\iff a-b+c=0. Therefore J​(u)=a​(1βˆ’e4​π​i​u)J(u)=a(1-e^{4\pi iu}). So

I1​(u)=3βˆ’3​u​Γ​(16βˆ’u)​Γ​(12+u)​aI_{1}(u)=3^{-3u}\Gamma\left(\frac{1}{6}-u\right)\Gamma\left(\frac{1}{2}+u\right)a

for some constant aa independent of uu. Thus it suffices to evaluate I1​(u)I_{1}(u) at any point, we will find I1​(βˆ’1/3)I_{1}(-1/3), from equation (1.4),

I1​(βˆ’1/3)=∫0∞(x​(1+x)​(1+2​x))βˆ’2/3​𝑑x=3Γ—F12​(1/3,2/3;4/3;βˆ’1),I_{1}(-1/3)=\int_{0}^{\infty}(x(1+x)(1+2x))^{-2/3}dx=3\times{{}_{2}}F_{1}(1/3,2/3;4/3;-1),

and its evaluation follows from Kummer’s formula

F12​(a,b;1+aβˆ’b;βˆ’1)=Γ​(1+aβˆ’b)​Γ​(1+a/2)Γ​(1+a)​Γ​(1+a/2βˆ’b).{{}_{2}}F_{1}(a,b;1+a-b;-1)=\frac{\Gamma(1+a-b)\Gamma(1+a/2)}{\Gamma(1+a)\Gamma(1+a/2-b)}.

∎

References

  • [1] KamΒ Cheong Au. Vanishing of Witten zeta function at negative integers. arXiv preprint arXiv:2412.11879, 2024.
  • [2] DavidΒ H Bailey and JonathanΒ M Borwein. Computation and experimental evaluation of Mordell-Tornheim-Witten sum derivatives. Experimental Mathematics, 27(3):370–376, 2018.
  • [3] JonathanΒ M Borwein and Karl Dilcher. Derivatives and fast evaluation of the Tornheim zeta function. The Ramanujan Journal, 45:413–432, 2018.
  • [4] Walter Bridges, Benjamin Brindle, Kathrin Bringmann, and Johann Franke. Asymptotic expansions for partitions generated by infinite products. Mathematische Annalen, pages 1–40, 2024.
  • [5] Walter Bridges, Kathrin Bringmann, and Johann Franke. On the number of irreducible representations of 𝔰​𝔲​(3)\mathfrak{su}(3). arXiv preprint arXiv:2308.02319, 2023.
  • [6] Gregory Debruyne and GΓ©rald Tenenbaum. The saddle-point method for general partition functions. Indagationes Mathematicae, 31(4):728–738, 2020.
  • [7] Driss Essouabri. SingularitΓ© de sΓ©ries de Dirichlet associΓ©es Γ  des polynΓ΄mes de plusieurs variables et applications en thΓ©orie analytique des nombres. In Annales de l’institut Fourier, volumeΒ 47, pages 429–483, 1997.
  • [8] Jokke HΓ€sΓ€ and Alexander Stasinski. Representation growth of compact linear groups. Transactions of the American Mathematical Society, 372(2):925–980, 2019.
  • [9] Yasushi Komori, Kohji Matsumoto, and Hirofumi Tsumura. On Witten multiple zeta-functions associated with semisimple Lie algebras II. Journal of the Mathematical Society of Japan, 62(2):355–394, 2010.
  • [10] Yasushi Komori, Kohji Matsumoto, and Hirofumi Tsumura. On Witten multiple zeta-functions associated with semisimple lie algebras III. Multiple Dirichlet series, L-functions and automorphic forms, pages 223–286, 2012.
  • [11] Yasushi Komori, Kohji Matsumoto, and Hirofumi Tsumura. On Witten multiple zeta-functions associated with semi-simple Lie algebras V. Glasgow Mathematical Journal, 57(1):107–130, 2015.
  • [12] Yasushi Komori, Kohji Matsumoto, and Hirofumi Tsumura. The Theory of Zeta-functions of Root Systems. Springer, 2023.
  • [13] Christoph Koutschan. Advanced Applications of the Holonomic Systems Approach. PhD thesis, RISC, Johannes Kepler University, Linz, Austria, 2009.
  • [14] Kohji Matsumoto and Hirofumi Tsumura. On Witten multiple zeta-functions associated with semisimple Lie algebras I. In Annales de l’institut Fourier, volumeΒ 56, pages 1457–1504, 2006.
  • [15] Takashi Nakamura. A functional relation for the Tornheim double zeta function. Acta Arithmetica, 125(3):257–263, 2006.
  • [16] Takuya Okamoto. Multiple zeta values related with the zeta-function of the root system type A2A_{2}, B2B_{2}, and G2G_{2}. Commentarii mathematici Universitatis Sancti Pauli, 61(1):9–27, 2012.
  • [17] Kazuhiro Onodera. A functional relation for Tornheim’s double zeta functions. Acta Arithmetica, 162(4):337–354, 2014.
  • [18] Oskar Perron. Über Summengleichungen und poincarΓ©sche Differenzengleichungen. Mathematische Annalen, 84(1):1–15, 1921.
  • [19] Dan Romik. On the number of nn-dimensional representations of S​U​(3)SU(3), the Bernoulli numbers, and the Witten zeta function. Acta Arithmetica, 180(2):111–159, 2017.
  • [20] Simon Rutard. Values and derivative values at nonpositive integers of generalized multiple Hurwitz zeta functions, applications to Witten zeta functions. arXiv preprint arXiv:2312.04725, 2023.
  • [21] EdmundΒ Taylor Whittaker and GeorgeΒ Neville Watson. A course of modern analysis: an introduction to the general theory of infinite processes and of analytic functions; with an account of the principal transcendental functions. University press, 1920.
  • [22] Don Zagier. Values of zeta functions and their applications. In First European Congress of Mathematics, volumeΒ 2, 1994.