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

[1]\fnmPing-Hsun \surChuang

\equalcont

The author contributed to the appendix.

[1]\orgdivDepartment of Mathematics, \orgnameNational Taiwan University, \orgaddress\cityTaipei, \postcode10617, \countryTaiwan

On the Periods of Twisted Moments of the Kloosterman Connection

[email protected]    \fnmJeng-Daw \surYu [email protected] *
Abstract

This paper aims to study the Betti homology and de Rham cohomology of twisted symmetric powers of the Kloosterman connection of rank two on the torus. We compute the period pairing and, with respect to certain bases, interpret these associated period numbers in terms of the Bessel moments. Via the rational structures on Betti homology and de Rham cohomology, we prove the \mathbb{Q}-linear and quadratic relations among these Bessel moments.

keywords:
Periods, Bessel moments, Kloosterman connection
pacs:
[

MSC Classification]32G20, 33C10, 34M35

Statements and Declarations

This research was supported by the grant 108-2115-M-002-003-MY2 of the Ministry of Science and Technology. The authors have no relevant financial or non-financial interests to disclose.

1 Introduction

Let 𝔾m,z=Spec([z,z1])\mathbb{G}_{\mathrm{m},z}=\operatorname{Spec}\,(\mathbb{Q}[z,z^{-1}]) be the algebraic torus over \mathbb{Q} with variable zz, and similarly for the torus 𝔾m,t\mathbb{G}_{\mathrm{m},t} with variable tt. Let Kl2{\mathrm{Kl}}_{2} be the Kloosterman connection (of rank two) on 𝔾m,z\mathbb{G}_{\mathrm{m},z} corresponding to the differential operator (zz)2z(z\partial_{z})^{2}-z. (For details, see section 2.) In [1], in order to study the Hodge aspects of the symmetric powers SymkKl2\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}, Fresán, Sabbah, and Yu consider the following settings. Let [2]:𝔾m,t𝔾m,z\left[2\right]:\mathbb{G}_{\mathrm{m},t}\rightarrow\mathbb{G}_{\mathrm{m},z} be the double cover induced by the ring homomorphism [z,z1][t,t1]\mathbb{Q}[z,z^{-1}]\rightarrow\mathbb{Q}[t,t^{-1}], given by zt2z\mapsto t^{2}. One obtains the pullback connection

Kl~2=[2]+Kl2.\widetilde{{\mathrm{Kl}}}_{2}=\left[2\right]^{+}{\mathrm{Kl}}_{2}.

The structure of Kl~2\widetilde{{\mathrm{Kl}}}_{2} is much simpler since it is the restriction to 𝔾m\mathbb{G}_{\mathrm{m}} of the Fourier transform of a regular holonomic module on the affine line. In addition, the symmetric power SymkKl2\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} appears in the pushforward [2]+SymkKl~2\left[2\right]_{+}\operatorname{Sym}^{k}{\widetilde{\mathrm{Kl}}}_{2} naturally in the decomposition [1, p. 1662]

[2]+SymkKl~2SymkKl2zSymkKl2,\left[2\right]_{+}\operatorname{Sym}^{k}{\widetilde{\mathrm{Kl}}}_{2}\cong\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\oplus\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2},

where zSymkKl2=(𝒪𝔾m,d+dz2z)SymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}=\left(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},{\mathrm{d}}+\frac{{\mathrm{d}}z}{2z}\right)\otimes\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}. In [2], Fresán, Sabbah, and Yu compute the de Rham cohomology and Betti homology for SymkKl2\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}. In this paper, we study the analogues for zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}.

1.1 Historical results and our results

Let I0(t)I_{0}(t) and K0(t)K_{0}(t) be modified Bessel functions. Define the Bessel moments

IKMk(a,b)=0I0(t)aK0(t)katbdt\displaystyle{\mathrm{IKM}}_{k}(a,b)=\int_{0}^{\infty}I_{0}(t)^{a}K_{0}(t)^{k-a}t^{b}{\mathrm{d}}t (1)

provided that 0ak0\leq a\leq k are non-negative integers, bb\in\mathbb{Z}, and the convergence of this integral. The particular Bessel moments of the form IKMa+b(a,2c1){\mathrm{IKM}}_{a+b}(a,2c-1) appear in two-dimensional quantum field theory as Feynman integrals [3, 4, 5]. From a mathematical point of view, these moments are realized as period integrals of SymkKl2\operatorname{Sym}^{k}\mathrm{Kl}_{2} and zSymkKl2\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2}. For the details, we refer to [2]. In that paper, Fresán, Sabbah, and Yu developed the Hodge theory on symmetric powers of the generalized Kloosterman connection Kln+1\mathrm{Kl}_{n+1} of rank (n+1)(n+1).

Sum rule identities

In [4, (220)], the authors provide the following conjecture on the (π)\mathbb{Q}(\pi) linear relation of Bessel moments which is called the “sum rule” in their paper.

Conjecture 1.

For each pair of integers (n,k)\left(n,k\right) with n2k2n\geq 2k\geq 2, the following combination of Bessel moments vanish

m=0n/2(1)m(n2m)πn2mIKM2n(n2m,n2k)=0.\sum_{m=0}^{\lfloor n/2\rfloor}(-1)^{m}\binom{n}{2m}\pi^{n-2m}{\mathrm{IKM}}_{2n}(n-2m,n-2k)=0.

Later in [6, (1.5)], Zhou uses the Hilbert transformation to prove this conjecture. Moreover, he also proves a “sum rule”:

Formula 2.

For each pair of integers (n,k)\left(n,k\right) with n12k2n-1\geq 2k\geq 2, the following conbination of Bessel moments vanish

m=1(n+1)/2(1)m(n2m1)πn2m+1IKM2n(n2m+1,n2k1)=0.\sum_{m=1}^{\lfloor(n+1)/2\rfloor}(-1)^{m}\binom{n}{2m-1}\pi^{n-2m+1}{\mathrm{IKM}}_{2n}(n-2m+1,n-2k-1)=0.

When the involved exponents of tt in (1) are odd, these two identities are both reproved by Fresán, Sabbah, and Yu [2]. The proof is to study the connection SymkKl2\operatorname{Sym}^{k}\mathrm{Kl}_{2} on 𝔾m\mathbb{G}_{\mathrm{m}} whose period integrals are those Bessel moments IKMk(a,b)\mathrm{IKM}_{k}(a,b) with odd bb. In this paper, by studying the connection zSymkKl2\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2} on 𝔾m\mathbb{G}_{\mathrm{m}}, we provide proofs of these two identities involving even powers of tt using a similar approach in section 5. The key point to consider the twisted connection zSymkKl2\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2} is that the period integrals of zSymkKl2\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2} are those Bessel moments IKMk(a,b)\mathrm{IKM}_{k}(a,b) with even bb. For example, we have the following result:

Formula 3 (Corollary 28).

For k=4r+4k=4r+4, a multiple of 44, one has

j=0r(k/22j)(1)jπ2jIKMk(2j,2i)={(1)rπ2r+2IKMk(2r+2,2i)if 0ir,(1)rπ2r+2IKMkreg(2r+2,2i)if r+1ik12.\sum_{j=0}^{r}\binom{k/2}{2j}(-1)^{j}\pi^{2j}{\mathrm{IKM}}_{k}(2j,2i)=\begin{cases}(-1)^{r}\pi^{2r+2}{\mathrm{IKM}}_{k}(2r+2,2i)&\text{if $0\leq i\leq r$},\\ (-1)^{r}\pi^{2r+2}{\mathrm{IKM}}_{k}^{\mathrm{reg}}(2r+2,2i)&\text{if $r+1\leq i\leq\lfloor\frac{k-1}{2}\rfloor$}.\end{cases}

The notation IKMkreg(2r+2,2i)\mathrm{IKM}^{\mathrm{reg}}_{k}(2r+2,2i) above denotes the regularized Bessel moments (see Lemma 21). Roughly speaking, the regularized Bessel moments are obtained from those Bessel moments IKMk(a,b)\mathrm{IKM}_{k}(a,b) with parameters k,a,bk,a,b that makes IKMk(a,b)\mathrm{IKM}_{k}(a,b) diverge but minus their divergent asymptotic. Therefore, our sum rule generalizes the sum rules in [6, 4].

\mathbb{Q}-dimension of Bessel moments

In [7], Zhou considers the \mathbb{Q}-vector subspace spanned by the Bessel moments in \mathbb{C}. This vector subspace is finite-dimensional due to the sum rule. Similarly, we have the following upper bound of the dimension.

Theorem 4 (Corollary 30).

For any kk and any 0a(k1)/20\leq a\leq\lfloor(k-1)/2\rfloor, the dimension of the \mathbb{Q}-vector space generated by the Bessel moments has an upper bound:

dimspan{IKMk(a,2j)j{0}}(k+1)/2.\dim{\mathrm{span}}_{\mathbb{Q}}\left\{{\mathrm{IKM}}_{k}(a,2j)\mid j\in\left\{0\right\}\cup\mathbb{N}\right\}\leq\lfloor(k+1)/2\rfloor.

For kk even, the dimension of the \mathbb{Q}-vector space generated by the regularized Bessel moments has an upper bound:

dimspan{IKMkreg(k/2,2j)j{0}}(k+1)/2.\dim{\mathrm{span}}_{\mathbb{Q}}\left\{{\mathrm{IKM}}^{\mathrm{reg}}_{k}\left(k/2,2j\right)\mid j\in\left\{0\right\}\cup\mathbb{N}\right\}\leq\lfloor(k+1)/2\rfloor.

Note that our statement involves the regularized Bessel moments. This conclusion is a more general result than the one given by Zhou.

Quadratic relations of Bessel moments

In [8], the authors prove a general result of quadratic relations between periods given by a self-dual connection. We apply this result and obtain the quadratic relation between the Bessel moments. Under certain bases of cohomologies, let BB be the topological pairing matrix on Betti homology, DD be the Poincaré pairing matrix on de Rham cohomology, and P,PcP,P_{\mathrm{c}} be the period pairing matrices between these two homology and cohomology. Then, the quadratic relations on these periods (Bessel moments) are given by

PD1Pc=(1)k(2π1)k+1B.\displaystyle PD^{-1}P_{\mathrm{c}}=(-1)^{k}(2\pi\sqrt{-1})^{k+1}B. (2)

The entries of the matrices PP and PcP_{\mathrm{c}} consist of \mathbb{Q}-linear combinations of Bessel moments and regularized Bessel moments, which are obtained in Section 5. Moreover, due to the rational structure of Betti homology and de Rham cohomology, the corresponding pairing matrices D,BD,B consist of rational numbers.

Note that the quadratic relations among Bessel moments does not depend on the choices of the bases. The effect of changing the bases of these homologies and cohomologies is just the conjugation of the matrices P,D,Pc,BP,D,P_{\mathrm{c}},B and thus on (2).

Determinants of Bessel moment matrix

Another interesting result is to compute the determinants of certain matrices consisting of Bessel moments. In [3, Conjectures 4 and 7], Broadhurst conjectures closed formulae of the determinants of the following two r×rr\times r matrices 𝐌r\mathbf{M}_{r} and 𝐍r\mathbf{N}_{r} involving the Bessel moments:

𝐌r=(IKM2r+1(a,2b1))1a,br,𝐍r=(IKM2r+2(a,2b1))1a,br.\mathbf{M}_{r}=\big{(}{\mathrm{IKM}}_{2r+1}(a,2b-1)\big{)}_{1\leq a,b\leq r},\quad\mathbf{N}_{r}=\big{(}{\mathrm{IKM}}_{2r+2}(a,2b-1)\big{)}_{1\leq a,b\leq r}.

Later, in [9], Zhou uses an analytic method to prove these two determinant formulae. Using a similar method as Zhou, we give explicit determinant formulae:

Formula 5 (Corollary 39).

For r1r\geq 1, we have

det(IKM2r1(i1,2j2))1i,jr\displaystyle\det\big{(}{\mathrm{IKM}}_{2r-1}(i-1,2j-2)\big{)}_{1\leq i,j\leq r} =πr(r+1)2r(r3)a=1r1ara2a+12a+1,\displaystyle=\sqrt{\pi}^{r(r+1)}\sqrt{2}^{r(r-3)}\prod_{a=1}^{r-1}\dfrac{a^{r-a}}{\sqrt{2a+1}^{2a+1}},
det(IKM2r(i1,2j2))1i,jr\displaystyle\det\big{(}{\mathrm{IKM}}_{2r}(i-1,2j-2)\big{)}_{1\leq i,j\leq r} =π(r+1)2Γ(r+12)12r(r+3)a=1r1(2a+1)ra(a+1)a+1.\displaystyle=\dfrac{\sqrt{\pi}^{(r+1)^{2}}}{\Gamma\left(\frac{r+1}{2}\right)}\dfrac{1}{\sqrt{2}^{r(r+3)}}\prod_{a=1}^{r-1}\dfrac{(2a+1)^{r-a}}{(a+1)^{a+1}}.

1.2 Approach

In [10], Bloch and Esnault study irregular connections on curves and provide the associated homology theory. Due to their results, we study the de Rham cohomology and Betti homology of zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} on 𝔾m\mathbb{G}_{\mathrm{m}} and provide explicit bases in order to find the periods.

In section 2, we introduce the twisted kk-th symmetric power of the Kloosterman connection zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}, which is the main object in this paper. We discuss the rational structures on the de Rham cohomology and Betti homology of the connection. Moreover, since the connection is self-dual, we introduce its algebraic and topological self-pairings. These pairings will play an important role in our computations.

In section 3, we study the de Rham cohomology and the de Rham cohomology with compact support of zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} and write down certain elements in these two cohomologies. Next, we introduce the Poincaré pairing between them and compute the pairing with respect to the elements we have constructed. Using the dimension result of de Rham cohomology, along with the non-vanishing determinant of the Poincaré pairing, in Corollary 14, we conclude that the explicit elements in de Rham cohomology form bases.

We study parallelly the Betti homology of zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} in section 4. Since our ambient space is a non-compact space ×\mathbb{C}^{\times}, we need to modify our Betti homology theory by allowing the chain to go to 0 or \infty. By controlling the growth behaviors of the horizontal sections, we study the moderate decay Betti homology and rapid decay Betti homology on zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}. Similarly, We first write down some elements in the moderate decay homology and rapid decay homology and compute their topological pairing explicitly. Moreover, by the duality of de Rham cohomology and Betti homology, the dimension of Betti homology is the same as the de Rham cohomology. Together with the topological pairing, we conclude that they are bases in Corollary 20.

Finally, in section 5, we compute the period pairing between the de Rham cohomologies and the Betti homologies and interpret them in terms of the Bessel moments. Note that our variety 𝔾m=Spec[z,z1]\mathbb{G}_{\mathrm{m}}=\operatorname{Spec}\mathbb{Q}[z,z^{-1}] and the connection zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} are defined over \mathbb{Q} and therefore, the de Rham cohomology and Betti homology are naturally endowed with a \mathbb{Q}-vector space structure. From the dimension constraint of homologies, after computing the period pairing, we obtain the \mathbb{Q}-linear relation of Bessel moments (Formula 3) and an upper bound of \mathbb{Q}-dimension of space spanned by the Bessel moments (Theorem 4). In addition, the self duality of zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} gives quadratic relations between these Bessel moments (2).

In appendix A.1, we provide an accurate analysis of the symmetric powers of the modified Bessel differential operator. The first usage belongs to section 3, which enables us to determine the dimension of the de Rham cohomology HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}. The second usage belongs to appendix A.2, which allows us to analyze the leading term of the Vanhove operator. This helps us to obtain the determinant formula (Formula 5).

2 The Kloosterman connection and its twisted symmetric powers

In this section, we recall the definition and basic properties of Kloosterman connection and its symmetric powers from [1, 2]. Besides, we recall the twisted connection on 𝔾m\mathbb{G}_{\mathrm{m}} obtained from the decomposition of the pushforward of trivial connection under the cyclic cover of 𝔾m\mathbb{G}_{\mathrm{m}}. Combining these connections, we obtain the twisted symmetric powers of the Kloosterman connection. Moreover, since these connections are all self-dual, the duality induces the algebraic pairings on them and the topological pairings on the sheaves of horizontal sections.

2.1 Self-duality and pairing on Kl2\mathrm{Kl}_{2}

The connection Kl2=(𝒪𝔾m,zv0𝒪𝔾m,zv1,){\mathrm{Kl}}_{2}=(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}}v_{0}\oplus\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}}v_{1},\nabla) consists of a rank 22 free sheaf on 𝔾m,z=Spec[z,z1]\mathbb{G}_{\mathrm{m},z}=\operatorname{Spec}\mathbb{Q}[z,z^{-1}] with basis of sections v0v_{0} and v1v_{1} and the connection \nabla on it given by

z(v0,v1)=(v0,v1)(0z10)dz.\displaystyle z\nabla\left(v_{0},v_{1}\right)=\left(v_{0},v_{1}\right)\begin{pmatrix}0&z\\ 1&0\end{pmatrix}{\mathrm{d}}z.

That is, zv0=v1dzz\nabla v_{0}=v_{1}\mathrm{d}z and v1=v0dz\nabla v_{1}=v_{0}\mathrm{d}z. The connection Kl2\mathrm{Kl}_{2} is self-dual in the sense that there exists an algebraic horizontal pairing ,alg\langle\ ,\,\rangle_{\text{alg}} on it:

(vi,vjalg)0i,j1=(0110)\left(\left\langle v_{i},v_{j}\right\rangle_{\mathrm{alg}}\right)_{0\leq i,j\leq 1}=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}

such that λ:Kl2Kl2\lambda:\mathrm{Kl}_{2}\rightarrow\mathrm{Kl}_{2}^{\vee} by (v0,v1)(v1,v0)\left(v_{0},v_{1}\right)\mapsto\left(v_{1}^{\vee},-v_{0}^{\vee}\right) makes the following diagram commute.

Kl2×Kl2{{\mathrm{Kl}}_{2}\times{\mathrm{Kl}}_{2}}(𝒪𝔾m,d){(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},\mathrm{d})}Kl2×Kl2{{\mathrm{Kl}}_{2}^{\vee}\times{\mathrm{Kl}}_{2}},alg\scriptstyle{\langle\ ,\,\rangle_{\text{alg}}}λ×1\scriptstyle{\lambda\times 1}natural

Here Kl2{\mathrm{Kl}}_{2}^{\vee} denotes the dual connection with the dual basis {v0,v1}\{v_{0}^{\vee},v_{1}^{\vee}\}.

Recall that the modified Bessel functions I0(t)I_{0}(t) and K0(t)K_{0}(t) satisfy the differential equation ((tt)2t2)y=0((t\partial_{t})^{2}-t^{2})y=0 and the Wronskian relation

I0(t)K0(t)I0(t)K0(t)=1t.I_{0}(t)K_{0}^{\prime}(t)-I_{0}^{\prime}(t)K_{0}(t)=\frac{-1}{t}. (3)

Under the change of variable z=t24z=\frac{t^{2}}{4}, the differential equation ((tt)2t2)y=0((t\partial_{t})^{2}-t^{2})y=0 becomes 4((zz)2z)y=04((z\partial_{z})^{2}-z)y=0. Define A0,B0A_{0},B_{0} be the fundamental solutions to the differential equation ((zz)2z)y=0((z\partial_{z})^{2}-z)y=0 by rescaling the modified Bessel functions. In addition, define A1,B1A_{1},B_{1} by zzz\partial_{z} differential of A0,B0A_{0},B_{0}:

A0(z)=2I0(2z),\displaystyle A_{0}(z)=-2I_{0}(2\sqrt{z}), A1(z)=zzA0(z);\displaystyle\qquad A_{1}(z)=z\partial_{z}A_{0}(z);
B0(z)=2K0(2z),\displaystyle B_{0}(z)=2K_{0}(2\sqrt{z}), B1(z)=zzB0(z).\displaystyle\qquad B_{1}(z)=z\partial_{z}B_{0}(z).

Here, the function z\sqrt{z} is taken to be the principal branch on the range |argz|<π|\arg z\,|<\pi. For other zz, these functions are defined via the analytic continuation. Throughout this paper, the multivalued functions such as zk/2z^{k/2} or z1/4z^{-1/4} are all treated in this way without a mention. The functions A0(z)A_{0}(z) and B0(z)B_{0}(z) are annihilated by the operator (zz)2z\left(z\partial_{z}\right)^{2}-z and real-valued on the ray >0\mathbb{R}_{>0}. This gives

zA1(z)=A0(z),zB1(z)=B0(z).\partial_{z}A_{1}(z)=A_{0}(z),\qquad\partial_{z}B_{1}(z)=B_{0}(z).

Together with the Wronskian relation A0B1A1B0=2A_{0}B_{1}-A_{1}B_{0}=2 from (3), we obtain a basis of horizontal sections of \nabla on Kl2{\mathrm{Kl}}_{2}

e0=12(A0v1A1v0),e1=12π1(B0v1B1v0).\displaystyle e_{0}=\frac{1}{2}(A_{0}v_{1}-A_{1}v_{0}),\qquad e_{1}=\frac{1}{2\pi\sqrt{-1}}(B_{0}v_{1}-B_{1}v_{0}). (4)

Denote Kl2\mathrm{Kl}_{2}^{\nabla} the local system of \mathbb{Q}-vector space generated by e0,e1e_{0},e_{1}. There exists a topological pairing ,top=2π1,alg\langle\ ,\,\rangle_{\mathrm{top}}=2\pi\sqrt{-1}\langle\ ,\,\rangle_{\mathrm{alg}} on Kl2\mathrm{Kl}_{2}^{\nabla}:

(ei,ejtop)0i,j1=(0110).\left(\left\langle e_{i},e_{j}\right\rangle_{\mathrm{top}}\right)_{0\leq i,j\leq 1}=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}.

2.2 Rational structures and pairings on (𝒪𝔾m,d+12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}).

Consider the double cover [2]:𝔾m,t𝔾m,z\left[2\right]:\mathbb{G}_{\mathrm{m},t}\rightarrow\mathbb{G}_{\mathrm{m},z} induced by the ring homomorphism [z,z1][t,t1]\mathbb{Q}[z,z^{-1}]\rightarrow\mathbb{Q}[t,t^{-1}], zt2z\mapsto t^{2}. Let T=(𝒪𝔾m,t,d)T=(\mathcal{O}_{\mathbb{G}_{\mathrm{m},t}},{\mathrm{d}}) be the trivial connection on 𝔾m,t\mathbb{G}_{\mathrm{m},t}. Via the ring homomorphism [z,z1][t,t1]\mathbb{Q}[z,z^{-1}]\rightarrow\mathbb{Q}[t,t^{-1}], we view [t,t1]\mathbb{Q}[t,t^{-1}] as the [z,z1]\mathbb{Q}[z,z^{-1}]-module. Then, from the decomposition of [z,z1]\mathbb{Q}[z,z^{-1}]-modules

[t,t1]=[z,z1]t[z,z1],\mathbb{Q}[t,t^{-1}]=\mathbb{Q}[z,z^{-1}]\oplus t\cdot\mathbb{Q}[z,z^{-1}],

the pushforward connection [2]+T\left[2\right]_{+}T decomposes into the direct sum

(𝒪𝔾m,z,d)(t𝒪𝔾m,z,d).(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},{\mathrm{d}})\oplus\left(t\cdot\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},{\mathrm{d}}\right).

The second component (t𝒪𝔾m,z,d)\left(t\cdot\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},{\mathrm{d}}\right) is isomorphic to (𝒪𝔾m,z,d+12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}) via the following diagram

𝒪𝔾m,z{{\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}}}𝒪𝔾m,zΩ𝔾m,z1{{\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}\otimes\Omega_{\mathbb{G}_{\mathrm{m}},z}^{1}}}t𝒪𝔾m,z{{t\cdot\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}}}t𝒪𝔾m,zΩ𝔾m,z1{{t\cdot\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}\otimes\Omega_{\mathbb{G}_{\mathrm{m}},z}^{1}}}d+dtt=d+12dzz\scriptstyle{\mathrm{d}+\frac{\mathrm{d}t}{t}=\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}}t\scriptstyle{t}\scriptstyle{\wr}t\scriptstyle{t}\scriptstyle{\wr}d\scriptstyle{\mathrm{d}} (5)

The dual connection of (𝒪𝔾m,z,d+12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}) is given by (𝒪𝔾m,z,d12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}-\frac{1}{2}\frac{\mathrm{d}z}{z}), and the two are isomorphic via multiplication by zz, that is, the following diagram commutes.

𝒪𝔾m,z{{\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}}}𝒪𝔾m,zΩ𝔾m,z1{{\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}\otimes\Omega_{\mathbb{G}_{\mathrm{m}},z}^{1}}}𝒪𝔾m,z{{\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}}}𝒪𝔾m,zΩ𝔾m,z1{{\mathcal{O}_{\mathbb{G}_{\mathrm{m}},z}\otimes\Omega_{\mathbb{G}_{\mathrm{m}},z}^{1}}}d+12dzz\scriptstyle{\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}}z\scriptstyle{z}\scriptstyle{\wr}z\scriptstyle{z}\scriptstyle{\wr}d12dzz\scriptstyle{\mathrm{d}-\frac{1}{2}\frac{\mathrm{d}z}{z}}

This induces a perfect algebraic horizontal pairing ,alg\langle\ ,\,\rangle_{\mathrm{alg}} on (𝒪𝔾m,z,d+12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}) given by

1,1alg=z.\langle 1,1\rangle_{\mathrm{alg}}=z.

On the other hand, the rational structure of the local system of horizontal sections of (𝒪𝔾m,t,d)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},t}},\mathrm{d}) is generated by 11. Under the isomorphism (5), the rational structure of local system of horizontal sections of (𝒪𝔾m,z,d+12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}) is generated by 1t=1z\frac{1}{t}=\frac{1}{\sqrt{z}}. Its dual connection (𝒪𝔾m,z,d12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}-\frac{1}{2}\frac{\mathrm{d}z}{z}) has local system of horizontal sections generated by z\sqrt{z}. This induces a rational topological pairing ,top\langle\ ,\,\rangle_{\mathrm{top}} on (𝒪𝔾m,z,d+12dzz)(\mathcal{O}_{\mathbb{G}_{\mathrm{m},z}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z})^{\nabla}

1z,1ztop=1.\left\langle\frac{1}{\sqrt{z}},\frac{1}{\sqrt{z}}\right\rangle_{\mathrm{top}}=1.

2.3 Algebraic and topological pairings on zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}

The kk-th symmetric product of Kl2\mathrm{Kl}_{2}, SymkKl2\operatorname{Sym}^{k}\mathrm{Kl}_{2}, is a rank k+1k+1 free sheaf over 𝒪𝔾m\mathcal{O}_{\mathbb{G}_{\mathrm{m}}} with basis of sections

v0av1ka=1|𝔖k|σ𝔖kσ(v0av1ka)a=0,1,,k,v_{0}^{a}v_{1}^{k-a}=\frac{1}{|\mathfrak{S}_{k}|}\sum_{\sigma\in\mathfrak{S}_{k}}\sigma\left(v_{0}^{\otimes a}\otimes v_{1}^{\otimes k-a}\right)\qquad a=0,1,\ldots,k,

where 𝔖k\mathfrak{S}_{k} is the symmetric group on kk elements. It is endowed with the induced connection from (Kl2,)\left(\mathrm{Kl}_{2},\nabla\right). After twisting with the connection (𝒪𝔾m,d+12dzz)\left(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},\mathrm{d}+\frac{1}{2}\frac{{\mathrm{d}z}}{z}\right), we define

zSymkKl2=(𝒪𝔾m,d+12dzz)SymkKl2.\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}=\left(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},{\mathrm{d}}+\frac{1}{2}\frac{{\mathrm{d}}z}{z}\right)\otimes\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}.

The induced connection \nabla on zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} is given by

v0av1ka\displaystyle\nabla v_{0}^{a}v_{1}^{k-a} =(ka)v0a+1v1ka1dz+azv0a1v1ka+1dz+12zv0av1kadz.\displaystyle=(k-a)v_{0}^{a+1}v_{1}^{k-a-1}{\mathrm{d}}z+\dfrac{a}{z}v_{0}^{a-1}v_{1}^{k-a+1}{\mathrm{d}}z+\dfrac{1}{2z}v_{0}^{a}v_{1}^{k-a}{\mathrm{d}}z. (6)

Note that zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} is the same sheaf as SymkKl2\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} but endowed with a different connection.

Via the self-duality on Kl2{\mathrm{Kl}}_{2} and on (𝒪𝔾m,d+12dzz)\left(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z}\right), we have the perfect algebraic pairing ,alg\left\langle\ ,\ \right\rangle_{\mathrm{alg}} on zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}:

zSymkKl2×zSymkKl2{\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\times\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}}(𝒪𝔾m,d){{\left(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},{\mathrm{d}}\right)}},alg\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{alg}}}

given by

v0kav1a,v0kbv1balg=zδk,a+b(1)aa!b!k!=(2π1)ke0kae1a,e0kbe1balg.\left\langle v_{0}^{k-a}v_{1}^{a},v_{0}^{k-b}v_{1}^{b}\right\rangle_{\mathrm{alg}}=z\delta_{k,a+b}(-1)^{a}\dfrac{a!b!}{k!}=(2\pi\sqrt{-1})^{k}\left\langle e_{0}^{k-a}e_{1}^{a},e_{0}^{k-b}e_{1}^{b}\right\rangle_{\mathrm{alg}}.

Indeed,

v0kav1a,v0kbv1balg\displaystyle\left\langle v_{0}^{k-a}v_{1}^{a},v_{0}^{k-b}v_{1}^{b}\right\rangle_{\mathrm{alg}} =1|𝔖k|σ𝔖kσ(v0kav1a),1|𝔖k|τ𝔖kτ(v0kbv1b)alg\displaystyle=\left\langle\frac{1}{|\mathfrak{S}_{k}|}\sum_{\sigma\in\mathfrak{S}_{k}}\sigma\left(v_{0}^{\otimes k-a}\otimes v_{1}^{\otimes a}\right),\frac{1}{|\mathfrak{S}_{k}|}\sum_{\tau\in\mathfrak{S}_{k}}\tau\left(v_{0}^{\otimes k-b}\otimes v_{1}^{\otimes b}\right)\right\rangle_{\mathrm{alg}}
=1(k!)2σ,τ𝔖kσ(v0kav1a),τ(v0kbv1b)alg,Kl21,1alg,(𝒪𝔾m,d+12dzz)\displaystyle=\frac{1}{(k!)^{2}}\sum_{\sigma,\tau\in\mathfrak{S}_{k}}\langle\sigma(v_{0}^{\otimes k-a}\otimes v_{1}^{\otimes a}),\tau(v_{0}^{\otimes k-b}\otimes v_{1}^{\otimes b})\rangle_{\mathrm{alg},\mathrm{Kl}_{2}}\cdot\langle 1,1\rangle_{\mathrm{alg},(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},\mathrm{d}+\frac{1}{2}\frac{\mathrm{d}z}{z})}
=1(k!)2(δka,bk!a!b!(1)a)z=zδk,a+b(1)aa!b!k!.\displaystyle=\frac{1}{(k!)^{2}}(\delta_{k-a,b}k!a!b!(-1)^{a})\cdot z=z\delta_{k,a+b}(-1)^{a}\frac{a!b!}{k!}.

By the definition of e0,e1e_{0},e_{1} in (4) and Wronskian relation A0B1A1B0=2A_{0}B_{1}-A_{1}B_{0}=2, similar computation shows the formula for the algebraic pairing e0kae1a,e0kbe1balg\langle e_{0}^{k-a}e_{1}^{a},e_{0}^{k-b}e_{1}^{b}\rangle_{\mathrm{alg}}.

The local system (zSymkKl2)(\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2})^{\nabla} is a \mathbb{Q}-vector space generated by the horizontal sections

1ze0ae1ka=1zσ𝔖kσ(e0ae1ka),a=0,1,,k,\dfrac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a}=\dfrac{1}{\sqrt{z}}\sum_{\sigma\in\mathfrak{S}_{k}}\sigma\big{(}e_{0}^{\otimes a}\otimes e_{1}^{\otimes k-a}\big{)},\quad a=0,1,\cdots,k, (7)

which are the products of the horizontal sections of the connections (𝒪𝔾m,d+12dzz)\left(\mathcal{O}_{\mathbb{G}_{\mathrm{m}}},\mathrm{d}+\frac{1}{2}\frac{{\mathrm{d}}z}{z}\right) and SymkKl2\operatorname{Sym}^{k}\mathrm{Kl}_{2}. The topological pairing ,\langle\ ,\,\rangle on Kl2\mathrm{Kl}_{2}^{\nabla} induces a topological pairing on (SymkKl2)(\operatorname{Sym}^{k}\mathrm{Kl}_{2})^{\nabla} and thus on (zSymkKl2)(\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2})^{\nabla}:

(zSymkKl2)×(zSymkKl2){(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\times(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}},{\mathbb{Q},},top\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{top}}}

where \mathbb{Q} on the right hand side is the constant sheaf associated with the field \mathbb{Q} on 𝔾m\mathbb{G}_{\mathrm{m}}. This pairing reads

1ze0ae1ka,1ze0be1kbtop=δk,a+b(1)kaa!b!k!\left<\frac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a},\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}\right>_{\mathrm{top}}=\delta_{k,a+b}(-1)^{k-a}\dfrac{a!b!}{k!}

by the similar computation as above.

3 The de Rham cohomology

In this section, we study the de Rham cohomology of the twisted Kloosterman connection HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}(\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2}) and its dual, compact support de Rham cohomology HdR,c1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}(\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2}). We will write down certain elements in these cohomologies explicitly and compute the Poincaré pairing between these elements. Finally, we conclude the bases of these two cohomologies in the end of this section.

3.1 Dimension of HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}

Proposition 6.

For the connection zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} on 𝔾m\mathbb{G}_{\mathrm{m}}, we have

dimHdR1(𝔾m,zSymkKl2)=k+12.\dim H^{1}_{\mathrm{dR}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\Big{)}=\left\lfloor\dfrac{k+1}{2}\right\rfloor.
Proof.

In [11, Lemma 2.9.13], we have the following formula.

Lemma 7.

On 𝔾m\mathbb{G}_{\mathrm{m}} with parameter zz, let 𝒟=𝒪𝔾m[z]\mathcal{D}=\mathcal{O}_{\mathbb{G}_{\mathrm{m}}}[\partial_{z}] be the ring of all differential operators on 𝔾m\mathbb{G}_{\mathrm{m}}. Write θz=zz\theta_{z}=z\partial_{z}. For a non-zero element L𝒟L\in\mathcal{D}, write LL into a finite sum of the form iziPi(θz)\sum_{i}z^{i}P_{i}(\theta_{z}), where Pi(x)[x]P_{i}(x)\in\mathbb{Q}[x]. Define integers a,ba,b by

aL:=max{iPi0};bL:=min{iPi0}.a_{L}:=\max\left\{i\mid P_{i}\neq 0\right\};\ \ \ \ b_{L}:=\min\left\{i\mid P_{i}\neq 0\right\}.

Then the Euler characteristic of the 𝒟\mathcal{D}-module 𝒟/𝒟L\mathcal{D}/\mathcal{D}L is given by χ(𝔾m,𝒟/𝒟L)=(aLbL)\chi\left(\mathbb{G}_{\mathrm{m}},\mathcal{D}/\mathcal{D}L\right)=-\left(a_{L}-b_{L}\right).

In this proof, we will follow the notations as in this lemma. Now, the differential operator on 𝔾m\mathbb{G}_{\mathrm{m}} associated with the connection Kl2{\mathrm{Kl}}_{2} is given by θz2z\theta_{z}^{2}-z which annihilates v0v_{0} and has fundamental solutions A0(z)A_{0}(z) and B0(z)B_{0}(z). Then, the differential operator for SymkKl2\operatorname{Sym}^{k}{\mathrm{Kl}}_{2} is given by the kk-th symmetric power of θz2z\theta_{z}^{2}-z, i.e., the differential operator annihilates v0kv_{0}^{k} and has fundamental solutions A0iB0kiA_{0}^{i}B_{0}^{k-i} for i=0,,ki=0,\ldots,k. Denote this operator by L~k+1𝒟\widetilde{L}_{k+1}\in\mathcal{D}. For zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}, the corresponding differential operator reads 1zL~k+1z=:L\frac{1}{\sqrt{z}}\widetilde{L}_{k+1}\sqrt{z}=:L since the solution is now given by 1zA0iB0ki\frac{1}{\sqrt{z}}A_{0}^{i}B_{0}^{k-i} for i=0,1,,ki=0,1,\ldots,k.

Recall in subsection 2.1, L2=(tt)2t2L_{2}=(t\partial_{t})^{2}-t^{2} is the differential operator annihilates I0(t)I_{0}(t) and K0(t)K_{0}(t). Write Lk+1L_{k+1} to be the kk-th symmetric power of (tt)2t2(t\partial_{t})^{2}-t^{2}. That is, Lk+1L_{k+1} annihilates I0a(t)K0ka(t)I_{0}^{a}(t)K_{0}^{k-a}(t) for a=0,,ka=0,\ldots,k. As discussed in subsection 2.1, the change of variable z=t24z=\frac{t^{2}}{4} sends Lk+1L_{k+1} to L~k+1\widetilde{L}_{k+1}. By Proposition 37, we have that aLk+1=2k+12a_{L_{k+1}}=2\lfloor\frac{k+1}{2}\rfloor, bLk+1=0b_{L_{k+1}}=0. Therefore, by the degree 22 change of variable z=t24z=\frac{t^{2}}{4}, we conclude aL~k+1=k+12a_{\widetilde{L}_{k+1}}=\lfloor\frac{k+1}{2}\rfloor, bL~k+1=0b_{\widetilde{L}_{k+1}}=0.

Using the fact that 1zθzz=θz+12\frac{1}{\sqrt{z}}\theta_{z}\sqrt{z}=\theta_{z}+\frac{1}{2}, we have 1zL~k+1z=ziPi(θz+12)\frac{1}{\sqrt{z}}\widetilde{L}_{k+1}\sqrt{z}=\sum z^{i}P_{i}\left(\theta_{z}+\frac{1}{2}\right) whenever L~k+1=ziPi(θz)\widetilde{L}_{k+1}=\sum z^{i}P_{i}(\theta_{z}). This shows aL=aL~k+1a_{L}=a_{\widetilde{L}_{k+1}} and bL=bL~k+1b_{L}=b_{\widetilde{L}_{k+1}}. Therefore, by Lemma 7, we have

χ(𝔾m,zSymkKl2)=χ(𝔾m,𝒟/𝒟L)=k+12.\chi\left(\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\right)=\chi\left(\mathbb{G}_{\mathrm{m}},\mathcal{D}/\mathcal{D}L\right)=-\left\lfloor\frac{k+1}{2}\right\rfloor.

Similar to the behavior of I0I_{0} and K0K_{0} [12, §10.30(i)], A0A_{0} is holomorphic at 0 and has exponential growth near infinity, and B0B_{0} has a log pole at 0. These imply all of the solutions 1zA0iB0ki\frac{1}{\sqrt{z}}A_{0}^{i}B_{0}^{k-i} are not algebraic solutions, and thus HdR0(𝔾m,zSymkKl2)=0H^{0}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}=0. Hence, combining the fact that HdR2(𝔾m,SymkKl2)=0H^{2}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}=0 by Artin vanishing theorem, we conclude that HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} has dimension k+12\left\lfloor\frac{k+1}{2}\right\rfloor. ∎

Remark 8.

In [11, Lemma 2.9.13], Katz provides the proof of lemma 7 only in the case 𝔾m,\mathbb{G}_{\mathrm{m},\mathbb{C}} which is over \mathbb{C}. Yet, the same proof still works in our situation 𝔾m,\mathbb{G}_{\mathrm{m},\mathbb{Q}} which is over \mathbb{Q}.

3.2 Compactly supported de Rham cohomology

Write k=k12k^{\prime}=\lfloor\frac{k-1}{2}\rfloor. Consider the k+1k^{\prime}+1 elements {v0kzidzz}i=0k\left\{v_{0}^{k}z^{i}\frac{{\mathrm{d}z}}{z}\right\}_{i=0}^{k^{\prime}} in HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}. We will prove these elements form a \mathbb{Q}-basis. (See Corollary 14.) The Poincaré dual of the de Rham cohomology is the de Rham cohomology with compact support. An element in the de Rham cohomology with compact support HdR,c1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} is represented by a triple (ξ,η,ω)(\xi,\eta,\omega), where ωHdR1(𝔾m,zSymkKl2)\omega\in H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} and ξ,η\xi,\eta are formal solutions to ξ=η=ω\nabla\xi=\nabla\eta=\omega at 0 and \infty respectively (See [8, Corollary 3.5]). The solutions are provided by the following lemma.

Lemma 9.

Suppose that k0,1,3mod4k\equiv 0,1,3\bmod{4}. For 0ik0\leq i\leq k^{\prime}, there exists (ξi,ηi)(zSymkKl2)0^(zSymkKl2)^(\xi_{i},\eta_{i})\in\big{(}\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\widehat{0}}\oplus\big{(}\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\widehat{\infty}} such that ξi=ηi=v0kzidzz\nabla\xi_{i}=\nabla\eta_{i}=v_{0}^{k}z^{i}\frac{\mathrm{d}z}{z}.

On the other hand, let k2mod4k\equiv 2\bmod{4}, say k=4r+2k=4r+2. For 0ik0\leq i\leq k^{\prime} with iri\neq r, there exists (ξi,ηi)(zSymkKl2)0^(zSymkKl2)^(\xi_{i},\eta_{i})\in\big{(}\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\widehat{0}}\oplus\big{(}\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\widehat{\infty}} such that

ξi=ηi=v0kzidzzγk,irv0kzrdzz,\nabla\xi_{i}=\nabla\eta_{i}=v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}-\gamma_{k,i-r}v_{0}^{k}z^{r}\frac{{\mathrm{d}}z}{z},

where γk,n\gamma_{k,n}\in\mathbb{Q} are the coefficients in the asymptotic expansion of (A0(z)B0(z))k/2(-A_{0}\left(z\right)B_{0}\left(z\right))^{k/2} given by (11) below.

Proof.

Near 0, we want to find

ξi=a=0kξi,a(z)v0av1kaa=0kzv0av1ka\xi_{i}=\sum_{a=0}^{k}\xi_{i,a}(z)v_{0}^{a}v_{1}^{k-a}\in\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket z\rrbracket v_{0}^{a}v_{1}^{k-a}

such that ξi=v0kzidzz\nabla\xi_{i}=v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}. Using the connection formula (6) on zSymkKl2\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}, we need to solve:

ddzξi,k(z)+(k1)ξi,k1(z)+12zξi,k(z)\displaystyle\frac{{\mathrm{d}}}{\mathrm{d}z}\xi_{i,k}(z)+(k-1)\xi_{i,k-1}(z)+\frac{1}{2z}\xi_{i,k}(z) =zi1,\displaystyle=z^{i-1},
ddzξi,a(z)+(ka+1)ξi,a1(z)+a+1zξi,a+1(z)+12zξi,a(z)\displaystyle\frac{{\mathrm{d}}}{\mathrm{d}z}\xi_{i,a}(z)+(k-a+1)\xi_{i,a-1}(z)+\frac{a+1}{z}\xi_{i,a+1}(z)+\frac{1}{2z}\xi_{i,a}(z) =0 for a=1,2,,k1,\displaystyle=0{\text{ for }}a=1,2,\cdots,k-1,
ddzξi,0(z)+1zξi,1(z)+12zξi,0(z)\displaystyle\frac{{\mathrm{d}}}{\mathrm{d}z}\xi_{i,0}(z)+\frac{1}{z}\xi_{i,1}(z)+\frac{1}{2z}\xi_{i,0}(z) =0.\displaystyle=0.

Write ξi,a=n=0ξi,a,nzn\xi_{i,a}=\sum_{n=0}^{\infty}\xi_{i,a,n}z^{n}. We solve ξi,a,n\xi_{i,a,n} recursively on nn. Suppose that we have solved ξi,a,j\xi_{i,a,j} for j<nj<n. Compare the coefficient of zn1z^{n-1} of the above system of equations and get

(n+12n+12kn+12k1n+121)(ξi,0,nξi,1,nξi,2,nξi,k,n)=lower order combinations.\left(\begin{array}[]{ccccc}&&&&n+\frac{1}{2}\\ &&&n+\frac{1}{2}&k\\ &&n+\frac{1}{2}&k-1&\\ &\iddots&\iddots&&\\ n+\frac{1}{2}&1&&&\end{array}\right)\left(\begin{array}[]{c}\xi_{i,0,n}\\ \xi_{i,1,n}\\ \xi_{i,2,n}\\ \vdots\\ \xi_{i,k,n}\end{array}\right)={\text{lower order combinations}}.

Since the first square matrix is invertible, ξi,a,n\xi_{i,a,n} is determined uniquely. Thus, we find ξia=0kzv0av1ka\xi_{i}\in\displaystyle\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket z\rrbracket v_{0}^{a}v_{1}^{k-a} such that ξi=v0kzidzz\nabla\xi_{i}=v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}. In k2mod4k\equiv 2\bmod{4} case, we only need to replace ξi\xi_{i} by ξiγk,irξr\xi_{i}-\gamma_{k,i-r}\xi_{r}.

Next, we turn to investigate the formal solutions at \infty using horizontal frames. We have the modified Bessel functions have the asymptotic expansions at 1t\frac{1}{t} [13, §7.23]

I0(t)\displaystyle I_{0}(t) et12πtn=0((2n1)!!)223nn!1tn,\displaystyle\sim e^{t}\frac{1}{\sqrt{2\pi t}}\sum_{n=0}^{\infty}\frac{((2n-1)!!)^{2}}{2^{3n}n!}\frac{1}{t^{n}}, |argt|<12π\displaystyle|\arg t\,|<\frac{1}{2}\pi (8)
K0(t)\displaystyle K_{0}(t) etπ2tn=0(1)n((2n1)!!)223nn!1tn,\displaystyle\sim e^{-t}\sqrt{\frac{\pi}{2t}}\sum_{n=0}^{\infty}(-1)^{n}\frac{((2n-1)!!)^{2}}{2^{3n}n!}\frac{1}{t^{n}}, |argt|<32π\displaystyle|\arg t\,|<\frac{3}{2}\pi (9)
I0(t)K0(t)\displaystyle I_{0}(t)K_{0}(t) 12tn=0((2n1)!!)323nn!1t2n.\displaystyle\sim\frac{1}{2t}\sum_{n=0}^{\infty}\frac{((2n-1)!!)^{3}}{2^{3n}n!}\frac{1}{t^{2n}}. (10)

Here, the notation n!!n!! is the double factorial of a positive integer nn defined by

n!!=k=0n21(n2k).n!!=\prod_{k=0}^{\lceil\frac{n}{2}\rceil-1}(n-2k).

Let w=1zw=\frac{1}{z} be the local coordinate at z=z=\infty. For kk even, by the last asymptotic expansion, we have

(A0(z)B0(z))k/2w1/4n=0γk,nwn,\displaystyle(-A_{0}(z)B_{0}(z))^{k/2}\sim w^{1/4}\sum_{n=0}^{\infty}\gamma_{k,n}w^{n}, (11)

where γk,0=1\gamma_{k,0}=1 and γk,n>0\gamma_{k,n}>0 for all n>0n>0. For convenience, we set γk,j=0\gamma_{k,j}=0 for all j<0j<0.

Following the notation in (4), let us set e¯1=π1e1\overline{e}_{1}=\pi\sqrt{-1}e_{1}. Then 1ze0ae1¯ka\frac{1}{\sqrt{z}}e_{0}^{a}\overline{e_{1}}^{k-a} are horizontal sections. Using the Wronskian relation A0B1A1B0=2A_{0}B_{1}-A_{1}B_{0}=2, we have v0=B0e0A0e¯1v_{0}=B_{0}e_{0}-A_{0}\overline{e}_{1}. Then, we obtain

v0kzidzz\displaystyle v_{0}^{k}z^{i}\frac{\mathrm{d}z}{z} =a=0k(ka)(A0)aB0kawi+3/21ze0kae¯1adw.\displaystyle=-\sum_{a=0}^{k}\binom{k}{a}\frac{(-A_{0})^{a}B_{0}^{k-a}}{w^{i+3/2}}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}\mathrm{d}w.

To solve the formal solution ηi\eta_{i} of ηi=v0kzidzz\nabla\eta_{i}=v_{0}^{k}z^{i}\frac{\mathrm{d}z}{z}, We first solve ηi,a\eta_{i,a} for each i,ai,a such that

dηi,a=(A0)aB0kawi+3/2.\displaystyle\mathrm{d}\eta_{i,a}=\frac{(-A_{0})^{a}B_{0}^{k-a}}{w^{i+3/2}}.

Then, ηi=a=0k(ka)ηi,a1ze0kae¯1a\eta_{i}=-\sum_{a=0}^{k}\binom{k}{a}\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a} is the desired solution. Moreover, since the function ηi,a\eta_{i,a} have wk/4w^{k/4} and the exponential factor (see (12) below), we need to justify that ηi\eta_{i} lies in a=0kwv0av1ka\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket w\rrbracket v_{0}^{a}v_{1}^{k-a} (not just in a=0kw1/4,e1/wv0av1ka\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket w^{1/4},e^{-1/\sqrt{w}}\rrbracket v_{0}^{a}v_{1}^{k-a}).

Near \infty, we have the expansion

(A0)aB0kawi+3/2={πk2ae2(k2a)/wwk/4i3/2Fi,a,ak2wk/4i3/2(n=0((2n1)!!)325nn!wn)k/2,a=k2\frac{(-A_{0})^{a}B_{0}^{k-a}}{w^{i+3/2}}=\begin{cases}\sqrt{\pi}^{k-2a}e^{-2(k-2a)/\sqrt{w}}w^{k/4-i-3/2}\cdot F_{i,a},&a\neq\frac{k}{2}\\ w^{k/4-i-3/2}\left(\sum_{n=0}^{\infty}\frac{\left(\left(2n-1\right)!!\right)^{3}}{2^{5n}n!}w^{n}\right)^{k/2},&a=\frac{k}{2}\end{cases}

where Fi,a1+wwF_{i,a}\in 1+\sqrt{w}\mathbb{Q}\llbracket\sqrt{w}\rrbracket. When ak2a\neq\frac{k}{2}, we can find an antiderivative ηi,a\eta_{i,a} of (A0)aB0kawi+3/2\frac{(-A_{0})^{a}B_{0}^{k-a}}{w^{i+3/2}} with the expansion

ηi,a=πk2ak2ae2(k2a)/wwk/4iGi,a\displaystyle\eta_{i,a}=\frac{\sqrt{\pi}^{k-2a}}{k-2a}e^{-2(k-2a)/\sqrt{w}}w^{k/4-i}\cdot G_{i,a} (12)

for some Gi,a1+wwG_{i,a}\in 1+\sqrt{w}\mathbb{Q}\llbracket\sqrt{w}\rrbracket. We analyze ηi,a1ze0kae¯1a\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}. Write e0kae¯1ae_{0}^{k-a}\overline{e}_{1}^{a} back to the expression in basis v0bv1kbv_{0}^{b}v_{1}^{k-b}:

e0kae¯1a\displaystyle e_{0}^{k-a}\overline{e}_{1}^{a} =2k(A0v1A1v0)ka(B0v1B1v0)a\displaystyle=2^{-k}(A_{0}v_{1}-A_{1}v_{0})^{k-a}\cdot(B_{0}v_{1}-B_{1}v_{0})^{a}
=2ke2(ka)z1πka(F1v0F2v1)kae2azπa(G1v1G2v0)a\displaystyle=2^{-k}e^{2(k-a)\sqrt{z}}\frac{1}{\sqrt{\pi}^{k-a}}(F_{1}v_{0}-F_{2}v_{1})^{k-a}\cdot e^{-2a\sqrt{z}}\sqrt{\pi}^{a}(G_{1}v_{1}-G_{2}v_{0})^{a}
=2ke2(k2a)zπ2ak(F1v0F2v1)ka(G1v1G2v0)a,\displaystyle=2^{-k}e^{2(k-2a)\sqrt{z}}\sqrt{\pi}^{2a-k}(F_{1}v_{0}-F_{2}v_{1})^{k-a}(G_{1}v_{1}-G_{2}v_{0})^{a},

where F1,F2,G1,G2z1/4z1/4F_{1},F_{2},G_{1},G_{2}\in z^{1/4}\mathbb{Q}\llbracket z^{-1/4}\rrbracket. Thus,

ηi,a1ze0kae¯1a\displaystyle\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a} =2kk2awk/4i+1/2Gi,a(F1v0F2v1)ka(G1v1G2v0)a,\displaystyle=\frac{2^{-k}}{k-2a}w^{k/4-i+1/2}G_{i,a}(F_{1}v_{0}-F_{2}v_{1})^{k-a}(G_{1}v_{1}-G_{2}v_{0})^{a},

where F1,F2,G1,G2z1/4z1/4F_{1},F_{2},G_{1},G_{2}\in z^{1/4}\mathbb{Q}\llbracket z^{-1/4}\rrbracket. We conclude that the desired ηi=a=0k(ka)ηi,a1ze0kae¯1a\eta_{i}=-\sum_{a=0}^{k}\binom{k}{a}\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a} has no exponential factor as a combination of monomials v0kbv1bv_{0}^{k-b}v_{1}^{b}, that is, ηi\eta_{i} lies in a=0kw1/4v0av1ka\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket w^{1/4}\rrbracket v_{0}^{a}v_{1}^{k-a}.

Next, we will prove ηi\eta_{i} lies in a=0kwv0av1ka\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket w\rrbracket v_{0}^{a}v_{1}^{k-a} by showing ηi\eta_{i} is invariant under the Galois group action. Let σ:w1/41w1/4\sigma:w^{1/4}\mapsto\sqrt{-1}w^{1/4} be the generator of the Galois group of the extension (w1/4)\mathbb{C}(w^{1/4}) of (w)\mathbb{C}(w). From the monodromy action [12, 10.34.5] of I0,K0I_{0},K_{0}, the σ\sigma action on Ai,BiA_{i},B_{i} is given by

σ(Aj,Bj)\displaystyle\sigma\left(A_{j},B_{j}\right) =(1π1Bj,π1Aj) for j=0,1,\displaystyle=\left(\frac{1}{\pi\sqrt{-1}}B_{j},-\pi\sqrt{-1}A_{j}\right){\text{ for }}j=0,1,

and thus on e0,e¯1e_{0},\overline{e}_{1} by

σ(e0,e¯1)=(1π1e¯1,π1e0);\displaystyle\sigma\left(e_{0},\overline{e}_{1}\right)=\left(\frac{1}{\pi\sqrt{-1}}\overline{e}_{1},-\pi\sqrt{-1}e_{0}\right); σ(e0kae¯1a)=(1)kπ2ake0ae¯1ka.\displaystyle\quad\sigma\left(e_{0}^{k-a}\overline{e}_{1}^{a}\right)=(\sqrt{-1})^{-k}\pi^{2a-k}e_{0}^{a}\overline{e}_{1}^{k-a}.

Moreover, we have

σ(ηi,a1ze0kae¯1a)=ηi,ka1ze0ae¯1ka.\sigma\left(\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}\right)=\eta_{i,k-a}\frac{1}{\sqrt{z}}e_{0}^{a}\overline{e}_{1}^{k-a}.

Hence, when k1,3mod4k\equiv 1,3\bmod{4}, the element ηi\eta_{i} is fixed by σ\sigma and

ηi=a=0k(ka)ηi,a1ze0kae¯1aa=0kwv0av1ka.\displaystyle\eta_{i}=-\sum_{a=0}^{k}\binom{k}{a}\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}\in\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket w\rrbracket v_{0}^{a}v_{1}^{k-a}.

This gives ηi=v0kzidzz.\nabla\eta_{i}=v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}.

When k=4r+4k=4r+4 and a=2r+2a=2r+2, the exponents of ww of the expansion of (A0)aB0kawi+3/2\frac{(-A_{0})^{a}B_{0}^{k-a}}{w^{i+3/2}} are in 12+\frac{1}{2}+\mathbb{Z} and one takes

ηi,2r+2wri+1/2ri+1/2Gi\eta_{i,2r+2}\sim\frac{w^{r-i+1/2}}{r-i+1/2}G_{i}

where Gi1+wwG_{i}\in 1+w\mathbb{Q}\llbracket w\rrbracket. More precisely, we have

Gi=1+n=1ri+1/2ri+1/2+nγk,nwn.G_{i}=1+\sum_{n=1}^{\infty}\frac{r-i+1/2}{r-i+1/2+n}\gamma_{k,n}w^{n}.

Moreover, ηi,2r+21z(e0e¯1)2r+2\eta_{i,2r+2}\frac{1}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+2} has no exponential factor as a combination of monomials v0kbv1bv_{0}^{k-b}v_{1}^{b} and is invariant under σ\sigma. Hence, when k0mod4k\equiv 0\bmod{4}, we take an element

ηi=a=0k(ka)ηi,a1ze0kae¯1aa=0kzv0av1ka\eta_{i}=-\sum_{a=0}^{k}\binom{k}{a}\eta_{i,a}\frac{1}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}\in\displaystyle\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket z\rrbracket v_{0}^{a}v_{1}^{k-a}

This gives ηi=v0kzidzz\nabla\eta_{i}=v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}.

Now, suppose that k=4r+2k=4r+2, a positive integer congruent to 22 modulo 44, and a=2r+1a=2r+1. Using the expansion (11), we have the residue:

Resw(A0)aB0kawi+3/2=γk,ir,\operatorname{Res}_{w}\frac{(-A_{0})^{a}B_{0}^{k-a}}{w^{i+3/2}}=\gamma_{k,i-r},

which vanishes if and only if ir1i\leq r-1. Therefore, for iri\geq r, there exists

ηi,2r+11riwriHi\eta_{i,2r+1}\sim\frac{1}{r-i}w^{r-i}\cdot H_{i}

such that

dηi,2r+1=(wi3/2γk,irwr3/2)(A0B0)2r+1dw\mathrm{d}\eta_{i,2r+1}=\left(w^{-i-3/2}-\gamma_{k,i-r}w^{-r-3/2}\right)(-A_{0}B_{0})^{2r+1}\mathrm{d}w

where Hi1+wwH_{i}\in 1+w\mathbb{Q}\llbracket w\rrbracket. Also, ηi,2r+11z(e0e¯1)2r+1\eta_{i,2r+1}\frac{1}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1} is invariant under σ\sigma. Moreover, ηi,2r+11z(e0e¯1)2r+1\eta_{i,2r+1}\frac{1}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1} has no exponential factor as a combination of monomials v0kbv1bv_{0}^{k-b}v_{1}^{b}. Thus, we have

v0kzidzzγk,irv0kzrdzz\displaystyle v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}-\gamma_{k,i-r}v_{0}^{k}z^{r}\frac{\mathrm{d}z}{z} =(a=0ak/2k(ka)ηi,aγk,irηr,aze0kae¯1a(kk/2)ηi,2r+1ze02r+1e¯12r+1)\displaystyle=\nabla\left(-\sum_{\begin{subarray}{c}a=0\\ a\neq k/2\end{subarray}}^{k}\binom{k}{a}\frac{\eta_{i,a}-\gamma_{k,i-r}\eta_{r,a}}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}-\binom{k}{k/2}\frac{\eta_{i,2r+1}}{\sqrt{z}}e_{0}^{2r+1}\overline{e}_{1}^{2r+1}\right)

and hence we find an element ηi\eta_{i} in a=0kzv0av1ka\displaystyle\bigoplus_{a=0}^{k}\mathbb{Q}\llbracket z\rrbracket v_{0}^{a}v_{1}^{k-a} such that ηi=v0kzidzzγk,irv0kzrdzz\nabla\eta_{i}=v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}-\gamma_{k,i-r}v_{0}^{k}z^{r}\frac{{\mathrm{d}}z}{z}. ∎

Now, we define some elements in the de Rham cohomology and the de Rham cohomology with compact support. In next subsection, we will prove that these elements form bases of the corresponding cohomology spaces (see Corollary 14).

Definition 10.

In the de Rham cohomology HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}, the classes ωk,i\omega_{k,i} are given as follows.

  1. 1.

    When k0,1,3mod4k\equiv 0,1,3\bmod{4}, define the k+1k^{\prime}+1 elements:

    ωk,i=v0kzidzz for i=0,1,2,,k.\omega_{k,i}=v_{0}^{k}z^{i}\frac{\mathrm{d}z}{z}{\text{ for }}i=0,1,2,\cdots,k^{\prime}.
  2. 2.

    When k2mod4k\equiv 2\bmod{4}, write k=4r+2k=4r+2 and define the kk^{\prime} elements:

    ωk,i={v0kzidzz,0ir1;v0kzidzzγk,irv0kzrdzz,r+1i2r,\omega_{k,i}=\begin{cases}v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z},&0\leq i\leq r-1;\\ v_{0}^{k}z^{i}\frac{{\mathrm{d}}z}{z}-\gamma_{k,i-r}v_{0}^{k}z^{r}\frac{{\mathrm{d}}z}{z},&r+1\leq i\leq 2r,\end{cases}

    where γk,n\gamma_{k,n}\in\mathbb{Q} are the coefficients in the asymptotic expansion of (A0(z)B0(z))k/2(-A_{0}\left(z\right)B_{0}\left(z\right))^{k/2} given by (11) above.

From the Lemma 9, we define the elements in the compactly supported de Rham cohomology.

Definition 11.

We define certain elements in the compactly supported de Rham cohomology HdR,c1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} as follows.

  1. 1.

    When k0,1,3mod4k\equiv 0,1,3\bmod{4}, define k+1k^{\prime}+1 elements

    ω~k,i=(ξi,ηi,ωk,i) for 0ik,\widetilde{\omega}_{k,i}=(\xi_{i},\eta_{i},\omega_{k,i}){\text{ for }}0\leq i\leq k^{\prime},

    where ξi=ηi=ωk,i\nabla\xi_{i}=\nabla\eta_{i}=\omega_{k,i}.

  2. 2.

    When k2mod4k\equiv 2\bmod{4}, write k=4r+2k=4r+2 and define kk^{\prime} elements

    ω~k,i=(ξi,ηi,ωk,i) for 0ir1 and r+1ik,\widetilde{\omega}_{k,i}=(\xi_{i},\eta_{i},\omega_{k,i}){\text{ for }}0\leq i\leq r-1{\text{ and }}r+1\leq i\leq k^{\prime},

    where ξi=ηi=ωk,i\nabla\xi_{i}=\nabla\eta_{i}=\omega_{k,i}.

  3. 3.

    In the case that k2mod4k\equiv 2\bmod{4}, write k=4r+2k=4r+2 and further define

    m^2r+1=(0,2k1z(e0e¯1)2r+1,0)HdR,c1(𝔾m,zSymkKl2).\widehat{m}_{2r+1}=\left(0,2^{k}\frac{1}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1},0\right)\in H^{1}_{\mathrm{dR,c}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\Big{)}.

    Here, e¯1:=π1e1\overline{e}_{1}:=\pi\sqrt{-1}e_{1} and e0,e1e_{0},e_{1} are horizontal sections of Kl2\mathrm{Kl}_{2} defined in (4).

Remark 12.

The pair of the formal solutions (ξi,ηi)(\xi_{i},\eta_{i}) is unique except in the case that there are solutions (ξ,η)(\xi,\eta) to ξ=η=0\nabla\xi=\nabla\eta=0. The latter happens only when k2mod4k\equiv 2\bmod{4}. In this circumstance, we fix the choice of (ξi,ηi)(\xi_{i},\eta_{i}) to be the one constructed in the proof of Lemma 9. These expressions will be used in the computations of Poincaré pairing and period pairing in the rest of this paper.

Further, we define the middle part de Rham cohomology, Hmid1(𝔾m,zSymkKl2)H^{1}_{\mathrm{mid}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}, to be the image of the projection HdR,c1(𝔾m,zSymkKl2)HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}\rightarrow H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}, (ξ,η,ω)ω(\xi,\eta,\omega)\mapsto\omega. We therefore have

ωk,iHmid1(𝔾m,zSymkKl2)\displaystyle\omega_{k,i}\in H^{1}_{\mathrm{mid}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\Big{)} for 0ik when k0,1,3mod4;\displaystyle{\text{ for }}0\leq i\leq k^{\prime}{\text{ when }}k\equiv 0,1,3\bmod{4};
ωk,iHmid1(𝔾m,zSymkKl2)\displaystyle\omega_{k,i}\in H^{1}_{\mathrm{mid}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\Big{)} for 0ik,ir when k=4r+2.\displaystyle{\text{ for }}0\leq i\leq k^{\prime},\,i\neq r{\text{ when }}k=4r+2.

We may regard Hmid1(𝔾m,zSymkKl2)H^{1}_{\mathrm{mid}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} as a quotient of HdR,c1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} containing the class of elements ω~k,i\widetilde{\omega}_{k,i}.

3.3 Poincaré pairing

We have the following Poincaré pairing between the de Rham cohomology and the compactly supported de Rham cohomology. Recall the algebraic pairing ,alg\left\langle\ ,\ \right\rangle_{\mathrm{alg}} is introduced in section 2.3.

HdR,c1(𝔾m,zSymkKl2)HdR1(𝔾m,zSymkKl2){{H^{1}_{\text{dR},c}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\text{Kl}}_{2}\Big{)}\otimes H^{1}_{\text{dR}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\text{Kl}}_{2}\Big{)}}}(k1){\mathbb{Q}(-k-1)}(m^0,m^,ω)η{{(\widehat{m}_{0},\widehat{m}_{\infty},\omega)\otimes\eta}}Reszm^0,ηalg+Reswm^,ηalg{{\operatorname{Res}_{z}\left\langle\widehat{m}_{0},\eta\right\rangle_{{\mathrm{alg}}}+\operatorname{Res}_{w}\left\langle\widehat{m}_{\infty},\eta\right\rangle_{{\mathrm{alg}}}}},Poin\scriptstyle{\left\langle\ ,\ \right\rangle_{\text{Poin}}}

Here, a one-form η\eta occurs in m^,ηalg\langle\widehat{m},\eta\rangle_{\mathrm{alg}}. This algebraic pairing means m^,falgdz\langle\widehat{m},f\rangle_{\mathrm{alg}}\mathrm{d}z whenever η=fdz\eta=f\mathrm{d}z. The notation (k1)\mathbb{Q}(-k-1) is the (k+1)(k+1)-time tensor product of the Tate structures (1)\mathbb{Q}(-1). As a vector space, (k1)\mathbb{Q}(-k-1) is nothing but \mathbb{Q}. Here, in consideration of Hodge filtrations, we use (k1)\mathbb{Q}(-k-1) instead of \mathbb{Q} to indicate the Hodge filtrations on both sides respect the Poincaré pairing. Note that the Poincaré pairing induces on the middle part de Rham cohomology which we still call it ,Poin\langle\ ,\ \rangle_{\mathrm{Poin}}:

Hmid1(𝔾m,zSymkKl2)Hmid1(𝔾m,zSymkKl2){{H^{1}_{\mathrm{mid}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\text{Kl}}_{2}\Big{)}\otimes H^{1}_{\mathrm{mid}}\Big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\text{Kl}}_{2}\Big{)}}}(k1).{\mathbb{Q}(-k-1).},Poin\scriptstyle{\left\langle\ ,\ \right\rangle_{\text{Poin}}}
Proposition 13.

Under the notation as in Definition 11, for j0j\geq 0, we have the Poincaré pairing

ω~k,i,v0kzjdzzPoin\displaystyle\left\langle\widetilde{\omega}_{k,i},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{Poin}} ={0if i+jk1,k:arbitrary,(2)kk!k!!if i+j=k,k1,3mod4,(kk/2)2k(ri+1/2)if i+j=k,k0mod4,k=4r+4,(kk/2)2k(ri)if i+j=k,k2mod4,k=4r+2.\displaystyle=\begin{cases}0&{\text{if }}i+j\leq k^{\prime}-1,\ k:{\text{arbitrary,}}\\ (-2)^{k^{\prime}}\frac{k^{\prime}!}{k!!}&{\text{if }}i+j=k^{\prime},\ k\equiv 1,3\bmod{4},\\ \frac{\binom{k}{k/2}}{2^{k}(r-i+1/2)}&{\text{if }}i+j=k^{\prime},\ k\equiv 0\bmod{4},\ k=4r+4,\\ -\frac{\binom{k}{k/2}}{2^{k}(r-i)}&{\text{if }}i+j=k^{\prime},\ k\equiv 2\bmod{4},\ k=4r+2.\end{cases}

Moreover, if k=4r+2k=4r+2, we have

m^2r+1,v0kzjdzzPoin={0if j<r,γk,jrif jr.\left\langle\widehat{m}_{2r+1},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{Poin}}=\begin{cases}0&{\text{if }}j<r,\\ \gamma_{k,j-r}&{\text{if }}j\geq r.\end{cases}

In particular, the Poincaré pairing matrix between k+1k^{\prime}+1 elements in Definition 11 and k+1k^{\prime}+1 elements {v0kzidzz}i=0k\left\{v_{0}^{k}z^{i}\frac{{\mathrm{d}z}}{z}\right\}_{i=0}^{k^{\prime}} in HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} is non-degenerate.

Proof.

In this proof, we will follow the notations as in the proof of Lemma 9. We first discuss the residue at z=0z=0. For any kk and 0i,jk0\leq i,j\leq k^{\prime}, we compute

Reszξi,v0kzjdzzalg\displaystyle\operatorname{Res}_{z}\left\langle\xi_{i},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}} =a=0kReszξi,av0av1ka,v0kzjdzzalg\displaystyle=\sum_{a=0}^{k}\operatorname{Res}_{z}\left\langle\xi_{i,a}v_{0}^{a}v_{1}^{k-a},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}}
=Reszξi,0v1k,v0kzjdzzalg\displaystyle=\operatorname{Res}_{z}\left\langle\xi_{i,0}v_{1}^{k},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}}
=Resz((1)kξi,0zj)=0\displaystyle=\operatorname{Res}_{z}\left((-1)^{k}\xi_{i,0}z^{j}\right)=0

where ξi,0z\xi_{i,0}\in\mathbb{Q}\llbracket z\rrbracket.

Next, we discuss the residue at z=z=\infty. When k1,3mod4k\equiv 1,3\bmod{4} and for any 0i,jk0\leq i,j\leq k^{\prime}, we compute

Reswηi,v0kzjdzzalg\displaystyle\operatorname{Res}_{w}\left\langle\eta_{i},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}} =a=0k(ka)Reswηi,ae0kae¯1az,v0kzjdzzalg\displaystyle=-\sum_{a=0}^{k}\binom{k}{a}\operatorname{Res}_{w}\left\langle\eta_{i,a}\frac{e_{0}^{k-a}\overline{e}_{1}^{a}}{\sqrt{z}},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}}
=a,b=0k(ka)(kb)Reswηi,ae0kae¯1az,(A0)bB0kbwj+3/2e0kbe¯1bzdwalg\displaystyle=\sum_{a,b=0}^{k}\binom{k}{a}\binom{k}{b}\operatorname{Res}_{w}\left\langle\eta_{i,a}\frac{e_{0}^{k-a}\overline{e}_{1}^{a}}{\sqrt{z}},\frac{(-A_{0})^{b}B_{0}^{k-b}}{w^{j+3/2}}\frac{e_{0}^{k-b}\overline{e}_{1}^{b}}{\sqrt{z}}\mathrm{d}w\right\rangle_{\mathrm{alg}}
=12ka=0k(1)a(ka)Resw(ηi,a1wj+3/2(A0)kaB0adw)\displaystyle=\frac{1}{2^{k}}\sum_{a=0}^{k}(-1)^{a}\binom{k}{a}\operatorname{Res}_{w}\left(\eta_{i,a}\frac{1}{w^{j+3/2}}(-A_{0})^{k-a}B_{0}^{a}{\mathrm{d}}w\right)
=12ka=0k(1)a(ka)k2aResw(w(k1)/2ij1Gi,aFj,kadw)\displaystyle=\frac{1}{2^{k}}\sum_{a=0}^{k}\frac{(-1)^{a}\binom{k}{a}}{k-2a}\operatorname{Res}_{w}\left(w^{(k-1)/2-i-j-1}G_{i,a}F_{j,k-a}{\mathrm{d}}w\right)
={0 if i+jk1,(2)kk!k!! if i+j=k.\displaystyle=\begin{cases}0&{\text{ if }}i+j\leq k^{\prime}-1,\\ (-2)^{k^{\prime}}\frac{k^{\prime}!}{k!!}&{\text{ if }}i+j=k^{\prime}.\end{cases}

where Gi,a,Fj,ka1+wwG_{i,a},F_{j,k-a}\in 1+\sqrt{w}\mathbb{Q}\llbracket\sqrt{w}\rrbracket and the last equality follows from [2, lemma 3.18].

When k0mod4k\equiv 0\bmod{4}, write k=4r+4k=4r+4. For any 0i,jk0\leq i,j\leq k^{\prime}, we compute

Reswηi,v0kzjdzzalg\displaystyle\operatorname{Res}_{w}\left\langle\eta_{i},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}} =12ka=0ak2k(1)a(ka)k2aResw(w(k1)/2ij1Gi,aFj,kadw)\displaystyle=\frac{1}{2^{k}}\sum_{\begin{subarray}{c}a=0\\ a\neq\frac{k}{2}\end{subarray}}^{k}\frac{(-1)^{a}\binom{k}{a}}{k-2a}\operatorname{Res}_{w}\left(w^{(k-1)/2-i-j-1}G_{i,a}F_{j,k-a}{\mathrm{d}}w\right)
+(1)k/22k(kk/2)Resw(ηi,2r+21wj+3/2(A0B0)2r+2dw)\displaystyle\qquad+\frac{(-1)^{k/2}}{2^{k}}\binom{k}{k/2}\operatorname{Res}_{w}\left(\eta_{i,2r+2}\frac{1}{w^{j+3/2}}(-A_{0}B_{0})^{2r+2}{\mathrm{d}}w\right)
=12ka=0ak2k(1)a(ka)k2aResw(w(k1)/2ij1Gi,aFj,kadw)\displaystyle=\frac{1}{2^{k}}\sum_{\begin{subarray}{c}a=0\\ a\neq\frac{k}{2}\end{subarray}}^{k}\frac{(-1)^{a}\binom{k}{a}}{k-2a}\operatorname{Res}_{w}\left(w^{(k-1)/2-i-j-1}G_{i,a}F_{j,k-a}{\mathrm{d}}w\right)
+(kk/2)2k(ri+1/2)Resw(wkij1GiF2r+2dw)\displaystyle\qquad+\frac{\binom{k}{k/2}}{2^{k}(r-i+1/2)}\operatorname{Res}_{w}\left(w^{k^{\prime}-i-j-1}\cdot G_{i}F_{2r+2}{\mathrm{d}}w\right)
={0 if i+jk1,(kk/2)2k(ri+1/2) if i+j=k.\displaystyle=\begin{cases}0&{\text{ if }}i+j\leq k^{\prime}-1,\\ \frac{\binom{k}{k/2}}{2^{k}(r-i+1/2)}&{\text{ if }}i+j=k^{\prime}.\end{cases}

where Gi,a,Fj,ka1+wwG_{i,a},F_{j,k-a}\in 1+\sqrt{w}\mathbb{Q}\llbracket\sqrt{w}\rrbracket, Gi1+wwG_{i}\in 1+w\mathbb{Q}\llbracket w\rrbracket and F2r+2=(n=0((2n1)!!)325nn!wn)2r+2F_{2r+2}=\left(\sum_{n=0}^{\infty}\frac{\left(\left(2n-1\right)!!\right)^{3}}{2^{5n}n!}w^{n}\right)^{2r+2}.

When k2mod4k\equiv 2\bmod{4}, the computation is similar to the case k0mod4k\equiv 0\bmod{4}.

Finally, we compute

Resw2k(e0e¯1)2r+1z,v0kzjdzzalg\displaystyle\operatorname{Res}_{w}\left\langle\frac{2^{k}(e_{0}\overline{e}_{1})^{2r+1}}{\sqrt{z}},v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\rangle_{\mathrm{alg}} =b=0k(kb)Resw2k(e0e¯1)2r+1z,(A0)bB0kbwj+3/2e0kbe¯1bzdwalg\displaystyle=-\sum_{b=0}^{k}\binom{k}{b}\operatorname{Res}_{w}\left\langle\frac{2^{k}(e_{0}\overline{e}_{1})^{2r+1}}{\sqrt{z}},\frac{(-A_{0})^{b}B_{0}^{k-b}}{w^{j+3/2}}\frac{e_{0}^{k-b}\overline{e}_{1}^{b}}{\sqrt{z}}{\mathrm{d}}w\right\rangle_{\mathrm{alg}}
=(kk/2)Resw2k(e0e¯1)2r+1z,(A0B0)2r+1wj+3/2(e0e¯1)2r+1zdwalg\displaystyle=-\binom{k}{k/2}\operatorname{Res}_{w}\left\langle\frac{2^{k}(e_{0}\overline{e}_{1})^{2r+1}}{\sqrt{z}},\frac{(-A_{0}B_{0})^{2r+1}}{w^{j+3/2}}\frac{(e_{0}\overline{e}_{1})^{2r+1}}{\sqrt{z}}{\mathrm{d}}w\right\rangle_{\mathrm{alg}}
=(1)2r+2Resw((A0B0)2r+1wj+3/2dw)\displaystyle=(-1)^{2r+2}\operatorname{Res}_{w}\left(\frac{(-A_{0}B_{0})^{2r+1}}{w^{j+3/2}}{\mathrm{d}}w\right)
=Resw(wk/4j3/2n=0γk,nwndw)\displaystyle=\operatorname{Res}_{w}\left(w^{k/4-j-3/2}\sum_{n=0}^{\infty}\gamma_{k,n}w^{n}{\mathrm{d}}w\right)
={0if j<r,γk,jrif jr.\displaystyle=\begin{cases}0&{\text{if }}j<r,\\ \gamma_{k,j-r}&{\text{if }}j\geq r.\end{cases}

Combining these residues, we obtain this proposition. ∎

Corollary 14 (Bases in de Rham side).

Let kk be a positive integer.

  1. 1.

    HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} has basis {v0kzjdzz}j=0k\left\{v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\}_{j=0}^{k^{\prime}}.

  2. 2.

    HdR,c1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} has basis

    {{ω~k,j}j=0k if k0,1,3mod4,{ω~k,j}j=0r1{ω~k,j}j=r+1k{m^2r+1} if k2mod4 with k=4r+2.\begin{cases}\left\{\widetilde{\omega}_{k,j}\right\}_{j=0}^{k^{\prime}}&{\text{ if }}k\equiv 0,1,3\bmod{4},\\ \left\{\widetilde{\omega}_{k,j}\right\}_{j=0}^{r-1}\cup\left\{\widetilde{\omega}_{k,j}\right\}_{j=r+1}^{k^{\prime}}\cup\left\{\widehat{m}_{2r+1}\right\}&{\text{ if }}k\equiv 2\bmod{4}{\text{ with }}k=4r+2.\end{cases}
  3. 3.

    HdR,mid1(𝔾m,zSymkKl2)H_{{\mathrm{dR,mid}}}^{1}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} has basis

    {{ωk,i}i=0k if k0,1,3mod4,{ωk,i}i=0r1{ωk,i}i=r+1k if k2mod4 with k=4r+2.\begin{cases}\left\{\omega_{k,i}\right\}_{i=0}^{k^{\prime}}&{\text{ if }}k\equiv 0,1,3\bmod{4},\\ \left\{\omega_{k,i}\right\}_{i=0}^{r-1}\cup\left\{\omega_{k,i}\right\}_{i=r+1}^{k^{\prime}}&{\text{ if }}k\equiv 2\bmod{4}{\text{ with }}k=4r+2.\end{cases}
Proof.

Putting the dimension result in Proposition 6, the non-vanishing determinant of the Poincaré pairing matrix in Proposition 13 together, we obtain this corollary, thanks to the following simple observation in linear algebra. ∎

Fact 15.

Let VV and WW be two nn-dimensional vector spaces over a field FF. Suppose that ,:V×WF\langle\ ,\ \rangle:V\times W\rightarrow F is a bilinear pairing. If {v1,,vn}V\{v_{1},\ldots,v_{n}\}\subseteq V and {w1,,wn}W\{w_{1},\ldots,w_{n}\}\subseteq W are subsets of vectors such that the matrix

(vi,wj)i,j=1nMn(F)\big{(}\langle v_{i},w_{j}\rangle\big{)}_{i,j=1\ldots n}\in M_{n}(F)

is invertible, then {v1,,vn}\{v_{1},\ldots,v_{n}\} is a basis of VV and {w1,,wn}\{w_{1},\ldots,w_{n}\} is a basis of WW.

4 The local system and the associated homology

In this section, we study the rapid decay homology and moderate decay homology of the local system (zSymkKl2)(\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2})^{\nabla}. We write down the explicit cycles in these homologies and compute their Betti intersection pairing. In the end, we finish this section by concluding the bases of these two homologies.

In order to write down the cycles in the homology, we need to understand the monodromy action of the horizontal sections of zSymkKl2\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2}. Recall {e0,e1}\{e_{0},e_{1}\} is the basis of the local system Kl2{\mathrm{Kl}}_{2}^{\nabla} defined in (4). From [12, 10.25(ii)], the modified Bessel function I0(t)I_{0}(t) is entire. On the other hand, K0(t)K_{0}(t) extends analytically to a multivalued function on ×\mathbb{C}^{\times} satisfying the monodromy K0(eπ1t)=K0(t)π1I0(t)K_{0}(e^{\pi\sqrt{-1}}t)=K_{0}(t)-\pi\sqrt{-1}I_{0}(t) from [12, 10.34]. This implies e0,e1e_{0},e_{1} undergo the monodromy action T:(e0,e1)(e0,e1+e0)T:\left(e_{0},e_{1}\right)\mapsto\left(e_{0},e_{1}+e_{0}\right) near 0. Then the basis in (7) of the local system (zSymkKl2)(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla} satisfies T:1ze0ae1ka1ze0a(e1+e0)kaT:\frac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a}\mapsto\frac{-1}{\sqrt{z}}e_{0}^{a}(e_{1}+e_{0})^{k-a} near 0.

4.1 Rapid decay cycles

Write k=k12k^{\prime}=\left\lfloor\frac{k-1}{2}\right\rfloor. Denote the chains on ×\mathbb{C}^{\times}:

σ0\displaystyle\sigma_{0} =the unit circle, starting at 1 and oriented counterclockwise;\displaystyle={\text{the unit circle}},{\text{ starting at }}1{\text{ and oriented counterclockwise}};
σ+\displaystyle\sigma_{+} =the interval [1,), starting at 1 toward +.\displaystyle=\text{the interval $\left[1,\infty\right)$},{\text{ starting at }}1{\text{ toward }}+\infty.

By the asymptotic expansion (8), (9), the horizontal sections 1ze0ae1ka\frac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a} decay exponentially along σ+\sigma_{+} for a=0,1,,ka=0,1,\ldots,k^{\prime}. We have the following lemma describing some elements in the rapid decay homology.

Lemma 16.

For 0bk0\leq b\leq k^{\prime}, the elements

δb\displaystyle\delta_{b} =σ+1ze0be1kb12σ01ze0be1kb+n=1kbdkb(n)σ02n1ze0be1kb,\displaystyle=\sigma_{+}\otimes\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}-\dfrac{1}{2}\sigma_{0}\otimes\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}+\sum_{n=1}^{k-b}d_{k-b}(n)\sigma_{0}^{2n}\otimes\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}, (13)

are rapid decay cycles in H1rd(𝔾m,(zSymkKl2))H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}, where dn(i)d_{n}(i) are real numbers satisfying

i=1ndn(i)(2i)m\displaystyle\sum_{i=1}^{n}d_{n}(i)\left(2i\right)^{m} =12, for m=1,2,,n.\displaystyle=-\dfrac{1}{2},{\text{ for }}m=1,2,\cdots,n.

In fact, by Cramer’s rule, one can write dn(i)=(1)in!2n+1(ni)(2n1)!!2i1d_{n}(i)=\frac{(-1)^{i}}{n!2^{n+1}}\binom{n}{i}\frac{\left(2n-1\right)!!}{2i-1} uniquely.

Proof.

We need to prove that dn(i)d_{n}(i) makes δb\delta_{b} into a cycle, that is, δb=0\partial\delta_{b}=0. The boundaries of chains σ+\sigma_{+} and σ0\sigma_{0} in δb\delta_{b} support at the point 1×1\in\mathbb{C}^{\times}. It suffices to check that the coefficient of 1×1\in\mathbb{C}^{\times} in δb\partial\delta_{b} is 0. Indeed, considering the monodromy action TT describe above, a direct computation shows the coefficient of 1×1\in\mathbb{C}^{\times} in δb\partial\delta_{b} is

1ze0be1kb12(1ze0be1kb+1ze0b(e1+e0)kb)\displaystyle\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}-\dfrac{1}{2}\left(\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}+\frac{1}{\sqrt{z}}e_{0}^{b}(e_{1}+e_{0})^{k-b}\right)
+n=1kbdkb(n)(1ze0be1kb1ze0b(e1+2ne0)kb)\displaystyle\hskip 40.0pt+\sum_{n=1}^{k-b}d_{k-b}(n)\left(\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}-\frac{1}{\sqrt{z}}e_{0}^{b}(e_{1}+2ne_{0})^{k-b}\right)
=\displaystyle= 1ze0be1kb12(1ze0be1kb+j=0kb(kbj)1ze0b+je1kbj)\displaystyle\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}-\dfrac{1}{2}\left(\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}+\sum_{j=0}^{k-b}\binom{k-b}{j}\frac{1}{\sqrt{z}}e_{0}^{b+j}e_{1}^{k-b-j}\right)
+n=1kbdkb(n)(1ze0be1kbm=0kb(kbm)(2n)m1ze0b+me1kbm)\displaystyle\hskip 40.0pt+\sum_{n=1}^{k-b}d_{k-b}(n)\left(\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}-\sum_{m=0}^{k-b}\binom{k-b}{m}(2n)^{m}\frac{1}{\sqrt{z}}e_{0}^{b+m}e_{1}^{k-b-m}\right)
=\displaystyle= 12j=1kb(kbj)1ze0b+je1kbjm=1kb(kbm)n=1kbdkb(n)(2n)m1ze0b+me1kbm\displaystyle-\dfrac{1}{2}\sum_{j=1}^{k-b}\binom{k-b}{j}\frac{1}{\sqrt{z}}e_{0}^{b+j}e_{1}^{k-b-j}-\sum_{m=1}^{k-b}\binom{k-b}{m}\sum_{n=1}^{k-b}d_{k-b}(n)(2n)^{m}\frac{1}{\sqrt{z}}e_{0}^{b+m}e_{1}^{k-b-m}
=\displaystyle= j=1kb(kbj)(12n=1kbdkb(n)(2n)j)1ze0b+je1kbj=0,\displaystyle\sum_{j=1}^{k-b}\binom{k-b}{j}\left(-\dfrac{1}{2}-\sum_{n=1}^{k-b}d_{k-b}(n)(2n)^{j}\right)\frac{1}{\sqrt{z}}e_{0}^{b+j}e_{1}^{k-b-j}=0,

where the last equality is the assumption on real numbers dn(i)d_{n}(i). ∎

From this lemma, we have k+1k^{\prime}+1 elements {δb}b=0k\{\delta_{b}\}_{b=0}^{k^{\prime}} in the rapid decay homology H1rd(𝔾m,(zSymkKl2))H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}. At the end of this section, we will prove these elements form a basis (see Corollary 20).

4.2 Moderate decay cycles

Define one more chain

+\displaystyle\mathbb{R}_{+} =the half line [0,), starting at 0 toward +.\displaystyle=\text{the half line $\left[0,\infty\right)$, starting at $0$ toward $+\infty$}.

By [12, §10.30(i)], the modified Bessel function K0(t)K_{0}\left(t\right) has log pole at 0, so the horizontal sections 1ze0ae1ka\frac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a} decay moderately along +\mathbb{R}_{+} near 0 for a=0,1,,k2a=0,1,\ldots,\lfloor\frac{k}{2}\rfloor. Moreover, by the expression (10), (I0K0)a(I_{0}K_{0})^{a} decay polynomially along +\mathbb{R}_{+} near \infty. Then, we define the moderate decay cycles in H1mod(𝔾m,(zSymkKl2))H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}

γa=+1ze0ae1ka, for a=0,1,2,,k2.\displaystyle\gamma_{a}=\mathbb{R}_{+}\otimes\frac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a},{\text{ for }}a=0,1,2,\cdots,\left\lfloor\dfrac{k}{2}\right\rfloor. (14)

They are indeed a cycle. The proof is the same as the above lemma by taking the homotopy as the radius of σ0\sigma_{0} tends to 0 and σ+\sigma_{+} tends to +\mathbb{R}_{+}. Since a rapid decay cycle is a moderate decay cycle as well, we have the natural map

H1rd(𝔾m,(zSymkKl2)){{H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}}}H1mod(𝔾m,(zSymkKl2)).{{H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}{\operatorname{Sym}^{k}}{\mathrm{Kl}}_{2})^{\nabla}\big{)}}.}

This natural map sends δb\delta_{b} to γb\gamma_{b} for b=0,1,,kb=0,1,\cdots,k^{\prime} by the homotopy argument. The following lemma shows when k2mod4k\equiv 2\bmod{4}, j=0(k2)/4(k/22j)δ2j\displaystyle\sum_{j=0}^{\left(k-2\right)/4}\binom{k/2}{2j}\delta_{2j} belongs to the kernel of this map.

Lemma 17.

In H1mod(𝔾m,(zSymkKl2))H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}, one has

j=0k/4(k/22j)γ2j=0 if k0mod4;\displaystyle\sum_{j=0}^{k/4}\binom{k/2}{2j}\gamma_{2j}=0{\text{ if }}k\equiv 0\bmod{4};
j=0(k2)/4(k/22j)γ2j=0 if k2mod4.\displaystyle\sum_{j=0}^{\left(k-2\right)/4}\binom{k/2}{2j}\gamma_{2j}=0{\text{ if }}k\equiv 2\bmod{4}.
Proof.

Let ρ:{(x,y)20<x,y,x+y<1}\rho:\left\{\left(x,y\right)\in\mathbb{R}^{2}\mid 0<x,y,x+y<1\right\}\rightarrow\mathbb{C} be the open simplicial 22-chain

ρ(x,y)=tanπ(x+y)2exp(41tan1yx)\rho\left(x,y\right)=\tan\dfrac{\pi\left(x+y\right)}{2}\exp\left(4\sqrt{-1}\tan^{-1}\dfrac{y}{x}\right)

that covers \mathbb{C} once. If kk is even, by the asymptotic expansion (10), the singular chain

Δ=ρ(1z(e1e0)k/2e1k/2)\Delta=\rho\otimes\left(\frac{1}{\sqrt{z}}(e_{1}-e_{0})^{k/2}e_{1}^{k/2}\right)

has moderate growth. The boundary of ρ\rho consists of two positive real lines +\mathbb{R}_{+}. From the monodromy action T:(e0,e1)(e0,e1+e0)T:(e_{0},e_{1})\mapsto(e_{0},e_{1}+e_{0}), one computes Δ\partial\Delta:

Δ\displaystyle\partial\Delta =+(1z(e1e0)k/2e1k/2)++(1ze1k/2(e0+e1)k/2)\displaystyle=\mathbb{R}_{+}\otimes\left(\frac{1}{\sqrt{z}}(e_{1}-e_{0})^{k/2}e_{1}^{k/2}\right)+\mathbb{R}_{+}\otimes\left(\frac{1}{\sqrt{z}}e_{1}^{k/2}(e_{0}+e_{1})^{k/2}\right)
=i=0k/2(1)i(k/2i)+1ze0ie1ki+i=0k/2(k/2i)+1ze0ie1ki\displaystyle=\sum_{i=0}^{k/2}(-1)^{i}\binom{k/2}{i}\mathbb{R}_{+}\otimes\frac{1}{\sqrt{z}}e_{0}^{i}e_{1}^{k-i}+\sum_{i=0}^{k/2}\binom{k/2}{i}\mathbb{R}_{+}\otimes\frac{1}{\sqrt{z}}e_{0}^{i}e_{1}^{k-i}
=i=0k/2(1+(1)i)(k/2i)γi.\displaystyle=\sum_{i=0}^{k/2}(1+(-1)^{i})\binom{k/2}{i}\gamma_{i}.

When k0mod4k\equiv 0\bmod{4}, this reads

12Δ=j=0k/4(k/22j)γ2j.\frac{1}{2}\partial\Delta=\sum_{j=0}^{k/4}\binom{k/2}{2j}\gamma_{2j}.

Thus, j=0k/4(k/22j)γ2j\sum_{j=0}^{k/4}\binom{k/2}{2j}\gamma_{2j} is homologous to zero in H1mod(𝔾m,(zSymkKl2))H_{1}^{\mathrm{mod}}(\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}\mathrm{Kl}_{2})^{\nabla}). The case when k2mod4k\equiv 2\bmod{4} is similar. ∎

Here, we have written down the 1+k21+\lfloor\frac{k}{2}\rfloor elements {γa}a=0k/2\{\gamma_{a}\}_{a=0}^{\lfloor k/2\rfloor} in the moderate decay homology H1mod(𝔾m,(zSymkKl2))H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}. At the end of this section, we will prove that these elements form a basis modulo the linear relation given in the above lemma (see Corollary 20).

Similar to the middle part de Rham cohomology in the previous section, we define the middle part Betti homology H1mid(𝔾m,(zSymkKl2))H_{1}^{{\mathrm{mid}}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} to be the image of H1rd(𝔾m,(zSymkKl2))H_{1}^{\mathrm{rd}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} in H1mod(𝔾m,(zSymkKl2))H_{1}^{\mathrm{mod}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}. More precisely, we have

γiH1mid(𝔾m,(zSymkKl2))\displaystyle\gamma_{i}\in H_{1}^{{\mathrm{mid}}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} for 0ik when k0,1,3mod4;\displaystyle{\text{ for }}0\leq i\leq k^{\prime}{\text{ when }}k\equiv 0,1,3\bmod{4};
γiH1mid(𝔾m,(zSymkKl2))\displaystyle\gamma_{i}\in H_{1}^{{\mathrm{mid}}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} for 1ik when k2mod4.\displaystyle{\text{ for }}1\leq i\leq k^{\prime}{\text{ when }}k\equiv 2\bmod{4}.

Also, we may regard H1mid(𝔾m,(zSymkKl2))H_{1}^{{\mathrm{mid}}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} as the quotient of H1rd(𝔾m,(zSymkKl2))H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} containing the class of elements δb\delta_{b}. At the end of this section, we will prove these elements form a basis (see Corollary 20).

4.3 Betti intersection pairing

We use the topological pairing ,top\left\langle\ ,\ \right\rangle_{\mathrm{top}} introduced in section 2.3 to define the Betti intersection pairing

H1rd(𝔾m,(zSymkKl2))×H1mod(𝔾m,(zSymkKl2)){{H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}\times H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}}}{\mathbb{Q}}(δ=iσisσi,γ=jτjsτj){{\big{(}\delta=\sum_{i}\sigma_{i}\otimes s_{\sigma_{i}},\gamma=\sum_{j}\tau_{j}\otimes s_{\tau_{j}}\big{)}}}i,jσiτjsσi,sτjtop{{\displaystyle\sum_{i,j}\sum_{\sigma_{i}\cap\tau_{j}}\left\langle s_{\sigma_{i}},s_{\tau_{j}}\right\rangle_{\mathrm{top}}}},Betti\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{Betti}}}

Here, we need to find representatives of δ=σisσi\delta=\sum\sigma_{i}\otimes s_{\sigma_{i}} and γ=τjsτj\gamma=\sum\tau_{j}\otimes s_{\tau_{j}} in their homology classes respectively such that any two chains σi\sigma_{i} and τj\tau_{j} intersect transversally for all i,ji,j. Then, for each pair (i,j)(i,j), σiτj\sigma_{i}\cap\tau_{j} consists of only finitely many topological intersection points. The sum over σiτj\sigma_{i}\cap\tau_{j} is then the sum of the topological pairings of the corresponding sections at each intersection point. Note that the Betti intersection pairing induces on the middle part Betti homology which we still call it ,Betti\langle\ ,\ \rangle_{\mathrm{Betti}}:

H1mid(𝔾m,(zSymkKl2))×H1mid(𝔾m,(zSymkKl2)){{H_{1}^{\mathrm{mid}}(\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\text{Kl}}_{2})^{\nabla})\times H_{1}^{\mathrm{mid}}(\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\text{Kl}}_{2})^{\nabla})}}{\mathbb{Q}},Betti\scriptstyle{\left\langle\ ,\ \right\rangle_{\text{Betti}}}

To compute the topological pairing with respect to the elements we had written down, we need to introduce the Euler numbers and Euler polynomials. The Euler polynomials En(x)E_{n}(x) are given by the following power series, and we define the numbers EnE_{n} for n0n\geq 0 as in [14],

n=0En(x)znn!=2ez+1exz,En=En(0).\sum_{n=0}^{\infty}E_{n}(x)\frac{z^{n}}{n!}=\frac{2}{e^{z}+1}e^{xz},\ \ E_{n}=E_{n}(0).

The first few EnE_{n} are

E0E_{0} E1E_{1} E2E_{2} E3E_{3} E4E_{4} E5E_{5} E6E_{6}
11 1/2-1/2 0 1/41/4 0 1/2-1/2 0

We have the inversion formula for Euler polynomials,

xn=En(x)+12k=0n1(nk)Ek(x).x^{n}=E_{n}(x)+\dfrac{1}{2}\sum_{k=0}^{n-1}\binom{n}{k}E_{k}(x).

Evaluating at x=0x=0, we get

k=0n1(nk)Ek=2En.\displaystyle\sum_{k=0}^{n-1}\binom{n}{k}E_{k}=-2E_{n}. (15)
Proposition 18.

We have the Betti intersection pairing

δb,γaBetti=(1)a(kba)(ka)12Ekab=(1)a+12(ka)!(kb)!k!Ekab(kab)!\left<\delta_{b},\gamma_{a}\right>_{\mathrm{Betti}}=(-1)^{a}\frac{\binom{k-b}{a}}{\binom{k}{a}}\frac{-1}{2}E_{k-a-b}=\frac{(-1)^{a+1}}{2}\frac{(k-a)!(k-b)!}{k!}\frac{E_{k-a-b}}{(k-a-b)!}

for b=0,,kb=0,\cdots,k^{\prime} and a=0,,k2a=0,\cdots,\left\lfloor\frac{k}{2}\right\rfloor.

Proof.

Fix some π<θ0<0-\pi<\theta_{0}<0 and let x0=exp(1θ0)x_{0}=\exp(\sqrt{-1}\theta_{0}). To compute the pairing δb,γaBetti\left<\delta_{b},\gamma_{a}\right>_{\mathrm{Betti}}, we move the ray σ+\sigma_{+} by adding the scalar (x01)(x_{0}-1) and let the circle σ0\sigma_{0} start at x0x_{0}. Then the component σ0j1ze0be1kb\sigma_{0}^{j}\otimes\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b} in the deformed δb\delta_{b} meets γa\gamma_{a} topologically jj times at the same point +1×+1\in\mathbb{C}^{\times}. At the ii-th intersection, the factor 1ze0be1kb\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b} becomes (1)i11ze0b(e1+(i1)e0)kb(-1)^{i-1}\frac{1}{\sqrt{z}}e_{0}^{b}(e_{1}+(i-1)e_{0})^{k-b} and we have

(1)i11ze0b(e1+(i1)e0)kb,1ze0ae1katop=(1)i1(i1)kab(1)a(kba)(ka).\left\langle(-1)^{i-1}\frac{1}{\sqrt{z}}e_{0}^{b}(e_{1}+(i-1)e_{0})^{k-b},\frac{1}{\sqrt{z}}e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}=(-1)^{i-1}(i-1)^{k-a-b}(-1)^{a}\dfrac{\binom{k-b}{a}}{\binom{k}{a}}.

By adding these contributions, we obtain

δb,γaBetti=(kba)(ka)(1)an=1kbdkb(n)Tkab(2n),\left<\delta_{b},\gamma_{a}\right>_{\mathrm{Betti}}=\dfrac{\binom{k-b}{a}}{\binom{k}{a}}(-1)^{a}\sum_{n=1}^{k-b}d_{k-b}(n)T_{k-a-b}(2n),

where

Tn(k)==0k1(1)n=1+2n+(1)k1(k1)n.T_{n}(k)=\sum_{\ell=0}^{k-1}(-1)^{\ell}\ell^{n}=-1+2^{n}-\cdots+(-1)^{k-1}(k-1)^{n}.

Kim [14] gave the following relation for Tn(k)T_{n}(k):

Tn(k)=(1)k+12=0n1(n)Ekn+En2(1+(1)k+1).T_{n}(k)=\dfrac{(-1)^{k+1}}{2}\sum_{\ell=0}^{n-1}\binom{n}{\ell}E_{\ell}k^{n-\ell}+\dfrac{E_{n}}{2}\left(1+(-1)^{k+1}\right).

Now, we have the following computation

n=1kbdkb(n)Tkab(2n)\displaystyle\sum_{n=1}^{k-b}d_{k-b}(n)T_{k-a-b}(2n) =n=1kbdkb(n)[12=0kab1(kab)E(2n)kab]\displaystyle=\sum_{n=1}^{k-b}d_{k-b}(n)\left[\dfrac{-1}{2}\sum_{\ell=0}^{k-a-b-1}\binom{k-a-b}{\ell}E_{\ell}(2n)^{k-a-b-\ell}\right]
=12=0kab1(kab)En=1kbdkb(n)(2n)kab\displaystyle=\dfrac{-1}{2}\sum_{\ell=0}^{k-a-b-1}\binom{k-a-b}{\ell}E_{\ell}\sum_{n=1}^{k-b}d_{k-b}(n)(2n)^{k-a-b-\ell}
=14=0kab1(kab)E\displaystyle=\dfrac{1}{4}\displaystyle\sum_{\ell=0}^{k-a-b-1}\binom{k-a-b}{\ell}E_{\ell}
=12Ekab\displaystyle=\dfrac{-1}{2}E_{k-a-b}

where the last equality follows from (15). ∎

Consider the (k+1)×(k+1)(k^{\prime}+1)\times(k^{\prime}+1) pairing matrix

Bk={(δb,γaBetti) 0bk, 0ak2 if k is odd,(δb,γaBetti)0bk, 1ak2 if k is even.B_{k}=\begin{cases}\left(\left<\delta_{b},\gamma_{a}\right>_{\mathrm{Betti}}\right)_{\ 0\leq b\leq k^{\prime},\ 0\leq a\leq\lfloor\frac{k}{2}\rfloor}&{\text{ if }}k{\text{ is odd,}}\\ \left(\left<\delta_{b},\gamma_{a}\right>_{\mathrm{Betti}}\right)_{0\leq b\leq k^{\prime},\ 1\leq a\leq\frac{k}{2}}&{\text{ if }}k{\text{ is even.}}\end{cases}

By Proposition 18, when kk is even, we have

Bk=((1)22(k1)!k!k!Ek1(k1)!(1)k/2+12(k/2)!k!k!Ek/2(k/2)!(1)22(k1)!(k/2+1)!k!Ek/2(k/2)!(1)k/2+12(k/2)!(k/2+1)!k!E1(1)!)B_{k}=\left(\begin{array}[]{ccc}\frac{(-1)^{2}}{2}\frac{(k-1)!k!}{k!}\frac{E_{k-1}}{(k-1)!}&\cdots&\frac{(-1)^{k/2+1}}{2}\frac{(k/2)!k!}{k!}\frac{E_{k/2}}{(k/2)!}\\ \vdots&\ddots&\vdots\\ \frac{(-1)^{2}}{2}\frac{(k-1)!(k/2+1)!}{k!}\frac{E_{k/2}}{(k/2)!}&\cdots&\frac{(-1)^{k/2+1}}{2}\frac{(k/2)!(k/2+1)!}{k!}\frac{E_{1}}{\left(1\right)!}\end{array}\right)

and that

Bk1=((1)2(k1)!(k1)!(k1)!Ek1(k1)!(1)k/22(k/2)!(k1)!(k1)!Ek/2(k/2)!(1)2(k1)!(k/2)!(k1)!Ek/2(k/2)!(1)k/22(k/2)!(k/2)!(k1)!E1(1)!).B_{k-1}=\left(\begin{array}[]{ccc}\frac{(-1)}{2}\frac{(k-1)!(k-1)!}{(k-1)!}\frac{E_{k-1}}{(k-1)!}&\cdots&\frac{(-1)^{k/2}}{2}\frac{(k/2)!(k-1)!}{(k-1)!}\frac{E_{k/2}}{(k/2)!}\\ \vdots&\ddots&\vdots\\ \frac{(-1)}{2}\frac{(k-1)!(k/2)!}{(k-1)!}\frac{E_{k/2}}{(k/2)!}&\cdots&\frac{(-1)^{k/2}}{2}\frac{(k/2)!(k/2)!}{(k-1)!}\frac{E_{1}}{\left(1\right)!}\end{array}\right).

Then we obtain the relation

Bk=1kdiag(k,k1,,k/2+1)Bk1.B_{k}=-\frac{1}{k}\operatorname{diag}(k,k-1,\cdots,k/2+1)\cdot B_{k-1}. (16)

Thus, BkB_{k} and Bk1B_{k-1} have the same rank whenever kk is even. Moreover, we may compute the determinant of BkB_{k} explicitly as given in the following proposition.

Proposition 19.

The determinant of BkB_{k} is given by the following.

  1. 1.

    When kk is odd, we have

    detBk\displaystyle\det B_{k} =2k1a=1kak+12a(2a+1)k2a.\displaystyle=2^{-k-1}\prod_{a=1}^{k^{\prime}}a^{k^{\prime}+1-2a}(2a+1)^{k^{\prime}-2a}.
  2. 2.

    When kk is even, we have

    detBk\displaystyle\det B_{k} =(1)(k+1)(k+3)2ka=1k(a+1)k2a1(2a+1)k+12a.\displaystyle=(-1)^{(k^{\prime}+1)(k^{\prime}+3)}2^{-k}\prod_{a=1}^{k^{\prime}}(a+1)^{k^{\prime}-2a-1}(2a+1)^{k^{\prime}+1-2a}.

In particular, they are all non-vanishing.

Proof.

Set 2n1=(1)n22n1E2n1\mathcal{E}_{2n-1}=(-1)^{n}2^{2n-1}E_{2n-1}. Apply the result [15, Eq. H12] in the following computations.

When k=2k+1k=2k^{\prime}+1 is odd, we have

detBk\displaystyle\det B_{k} =(1)(k+1)(k+2)/2(2k!)k+1[i=k+1ki!]2det(Ekk!Ek+1(k+1)!Ek+1(k+1)!E1(1)!)\displaystyle=\frac{(-1)^{(k^{\prime}+1)(k^{\prime}+2)/2}}{(2\cdot k!)^{k^{\prime}+1}}\left[\prod_{i=k^{\prime}+1}^{k}i!\right]^{2}\det\left(\begin{array}[]{ccc}\frac{E_{k}}{k!}&\cdots&\frac{E_{k^{\prime}+1}}{(k^{\prime}+1)!}\\ \vdots&\ddots&\vdots\\ \frac{E_{k^{\prime}+1}}{(k^{\prime}+1)!}&\cdots&\frac{E_{1}}{\left(1\right)!}\end{array}\right)
=1(2k!)k+1[i=k+1ki!]212(k+1)2det(kk!k+1(k+1)!k+1(k+1)!1(1)!)\displaystyle=\frac{1}{(2\cdot k!)^{k^{\prime}+1}}\left[\prod_{i=k^{\prime}+1}^{k}i!\right]^{2}\frac{1}{2^{(k^{\prime}+1)^{2}}}\det\left(\begin{array}[]{ccc}\frac{\mathcal{E}_{k}}{k!}&\cdots&\frac{\mathcal{E}_{k^{\prime}+1}}{(k^{\prime}+1)!}\\ \vdots&\ddots&\vdots\\ \frac{\mathcal{E}_{k^{\prime}+1}}{(k^{\prime}+1)!}&\cdots&\frac{\mathcal{E}_{1}}{\left(1\right)!}\end{array}\right)
=12(k+1)(k+2)(k!)k+1[i=k+1ki!]22k2k!k!j=1k(j1)!2(2j1)!2\displaystyle=\frac{1}{2^{(k^{\prime}+1)(k^{\prime}+2)}(k!)^{k^{\prime}+1}}\left[\prod_{i=k^{\prime}+1}^{k}i!\right]^{2}2^{k^{\prime 2}}\frac{k^{\prime}!}{k!}\prod_{j=1}^{k^{\prime}}\frac{(j-1)!^{2}}{(2j-1)!^{2}}
=12k+1a=1kak+12a(2a+1)k2a.\displaystyle=\frac{1}{2^{k+1}}\prod_{a=1}^{k^{\prime}}a^{k^{\prime}+1-2a}(2a+1)^{k^{\prime}-2a}.

When k=2k+2k=2k^{\prime}+2 is even, we have

detBk\displaystyle\det B_{k} =(1)(k+1)(k+4)/2(2k!)k+1[i=k+1k1i!(i+1)!]det(Ek1(k1)!Ek+1(k+1)!Ek+1(k+1)!E1(1)!)\displaystyle=\frac{(-1)^{(k^{\prime}+1)(k^{\prime}+4)/2}}{(2\cdot k!)^{k^{\prime}+1}}\left[\prod_{i=k^{\prime}+1}^{k-1}i!\left(i+1\right)!\right]\det\left(\begin{array}[]{ccc}\frac{E_{k-1}}{(k-1)!}&\cdots&\frac{E_{k^{\prime}+1}}{(k^{\prime}+1)!}\\ \vdots&\ddots&\vdots\\ \frac{E_{k^{\prime}+1}}{(k^{\prime}+1)!}&\cdots&\frac{E_{1}}{\left(1\right)!}\end{array}\right)
=(1)(k+1)(k+4)/2(2k!)k+1[i=k+1k1i!(i+1)!]\displaystyle=\frac{(-1)^{(k^{\prime}+1)(k^{\prime}+4)/2}}{(2\cdot k!)^{k^{\prime}+1}}\left[\prod_{i=k^{\prime}+1}^{k-1}i!\left(i+1\right)!\right]
(1)(k+1)(k+2)2(k+1)2det(k1(k1)!k+1(k+1)!k+1(k+1)!1(1)!)\displaystyle\hskip 100.0pt\cdot\frac{(\sqrt{-1})^{(k^{\prime}+1)(k^{\prime}+2)}}{2^{(k^{\prime}+1)^{2}}}\det\left(\begin{array}[]{ccc}\frac{\mathcal{E}_{k-1}}{(k-1)!}&\cdots&\frac{\mathcal{E}_{k^{\prime}+1}}{(k^{\prime}+1)!}\\ \vdots&\ddots&\vdots\\ \frac{\mathcal{E}_{k^{\prime}+1}}{(k^{\prime}+1)!}&\cdots&\frac{\mathcal{E}_{1}}{\left(1\right)!}\end{array}\right)
=(1)(k+1)(k+3)2(k+1)(k+2)(k!)k+1[i=k+1k1i!(i+1)!]2k2k!(k1)!j=1k(j1)!2(2j1)!2\displaystyle=\frac{(-1)^{(k^{\prime}+1)(k^{\prime}+3)}}{2^{(k^{\prime}+1)(k^{\prime}+2)}(k!)^{k^{\prime}+1}}\left[\prod_{i=k^{\prime}+1}^{k-1}i!\left(i+1\right)!\right]2^{k^{\prime 2}}\frac{k^{\prime}!}{(k-1)!}\prod_{j=1}^{k^{\prime}}\frac{(j-1)!^{2}}{(2j-1)!^{2}}
=(1)(k+1)(k+3)2ka=1k(a+1)k2a1(2a+1)k+12a.\displaystyle=\frac{(-1)^{(k^{\prime}+1)(k^{\prime}+3)}}{2^{k}}\prod_{a=1}^{k^{\prime}}(a+1)^{k^{\prime}-2a-1}(2a+1)^{k^{\prime}+1-2a}.\qed

Finally, before we conclude the basis of Betti homologies, we need to introduce the period pairings here. However, the details of the pairings will be given in the next section. By [2, Corollary 2.11], there exist two perfect pairings

H1rd(𝔾m,(zSymkKl2))×HdR1(𝔾m,zSymkKl2){{H_{1}^{\mathrm{rd}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}_{\mathbb{C}}\times H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\mathbb{C}}}}{\mathbb{C}},per\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{per}}}
HdR,c1(𝔾m,zSymkKl2)×H1mod(𝔾m,(zSymkKl2)){{H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\mathbb{C}}\times H_{1}^{\mathrm{mod}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}_{\mathbb{C}}}}{\mathbb{C}},per,c\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{per,c}}}

Here, the notation VV_{\mathbb{C}} means VV\otimes_{\mathbb{Q}}\mathbb{C}. For the next corollary, we just need to use the fact that these pairings are perfect. In the next section, we will compute these two pairings explicitly.

Corollary 20.

The natural map

H1rd(𝔾m,(zSymkKl2)){{H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}{\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}}\big{)}}}H1mod(𝔾m,(zSymkKl2)){{H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}{\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}}\big{)}}}

sending δb\delta_{b} to γb\gamma_{b} is an isomorphism when k0,1,3mod4k\equiv 0,1,3\bmod 4 and has a one-dimensional kernel when k2mod4k\equiv 2\bmod 4. Moreover, we find the following.

  1. 1.

    H1rd(𝔾m,(zSymkKl2))H^{\mathrm{rd}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} has basis {δb}b=0k\{\delta_{b}\}_{b=0}^{k^{\prime}}.

  2. 2.

    H1mod(𝔾m,(zSymkKl2))H^{\mathrm{mod}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} has basis

    {{γa}a=0k if k is odd;{γa}a=1k/2 if k is even.\begin{cases}\left\{\gamma_{a}\right\}_{a=0}^{k^{\prime}}&{\text{ if }}k{\text{ is odd;}}\\ \left\{\gamma_{a}\right\}_{a=1}^{k/2}&{\text{ if }}k{\text{ is even.}}\end{cases}
  3. 3.

    H1mid(𝔾m,(zSymkKl2))H^{\mathrm{mid}}_{1}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)} has basis

    {{γa}a=0k if k0,1,3mod4;{γa}a=1k if k2mod4.\begin{cases}\left\{\gamma_{a}\right\}_{a=0}^{k^{\prime}}&{\text{ if }}k\equiv 0,1,3\bmod{4};\\ \left\{\gamma_{a}\right\}_{a=1}^{k^{\prime}}&{\text{ if }}k\equiv 2\bmod{4}.\\ \end{cases}
Proof.

From the perfect period pairings, the dimension of rapid decay homology and moderate decay homology are both k+1k^{\prime}+1 by Proposition 6. Then, by the Fact 15 and the non-vanishing determinant of BkB_{k} in Proposition 19, we conclude 1.1. and 2.2. This also shows the natural map which sends δb\delta_{b} to γb\gamma_{b} for b=0,,kb=0,\ldots,k^{\prime} is an isomorphism when k1,3mod4k\equiv 1,3\bmod 4. When k2mod4k\equiv 2\bmod{4}, Lemma 17 describes the one-dimensional kernel of the natural map. Moreover, BkB_{k} has full rank k/2k/2 when kk is even by the relation (16). Hence, we conclude that the natural map is an isomorphism when k0mod4k\equiv 0\bmod{4}. ∎

5 Twisted moments as periods

In this section, we compute the period pairing of the basis of de Rham cohomology and Betti homology in Corollary 14 and Corollary 20. Also, we interpret these periods as the Bessel moments and regularized Bessel moments.

5.1 Bessel moments and regularized Bessel moments

The Bessel moments are defined by

IKMk(a,b)=0I0a(t)K0ka(t)tbdt.\displaystyle{\mathrm{IKM}}_{k}(a,b)=\int_{0}^{\infty}I_{0}^{a}(t)K_{0}^{k-a}(t)t^{b}\mathrm{d}t.

provided the convergence of the integral, that is, for non-negative integers k,a,bk,a,b satisfying aka\leq k^{\prime}, b0b\geq 0 or a=k2a=\frac{k}{2}, 0b<k0\leq b<k^{\prime}. The justification is given in the following lemma. Moreover, if a=k2a=\frac{k}{2} and bkb\geq k^{\prime}, by analyzing the singular integral, we could define the regularized Bessel moments IKMkreg(k2,b)\mathrm{IKM}_{k}^{\mathrm{reg}}\left(\frac{k}{2},b\right) by subtracting the singular part of the integral. The precise definition is also given in the following lemma.

Lemma 21.

The integral expression of Bessel moments

IKMk(a,b)=0I0a(t)K0ka(t)tbdt{\mathrm{IKM}}_{k}(a,b)=\int_{0}^{\infty}I_{0}^{a}(t)K_{0}^{k-a}(t)t^{b}{\mathrm{d}}t

converges for non-negative integers k,a,bk,a,b satisfying aka\leq k^{\prime}, b0b\geq 0 or a=k2a=\frac{k}{2}, 0b<k0\leq b<k^{\prime}. Moreover, in the case that kk is even, a=k2a=\frac{k}{2}, and bkb\geq k^{\prime} with bb even, the following two limits exist for 2jk2j\geq k^{\prime}:

IKMkreg(k2,2j)\displaystyle{\mathrm{IKM}}^{\mathrm{reg}}_{k}\left(\frac{k}{2},2j\right) :=limt(0t(I0K0)2r+2s2jdsm=0jr1γk,jr1mt2m+12k2j+2m(2m+1))if k=4r+4,\displaystyle:=\lim_{t\rightarrow\infty}\left(\int_{0}^{t}(I_{0}K_{0})^{2r+2}s^{2j}{\mathrm{d}}s-\sum_{m=0}^{j-r-1}\frac{\gamma_{k,j-r-1-m}t^{2m+1}}{2^{k-2j+2m}(2m+1)}\right)\ \ {\text{if }}k=4r+4,
IKMkreg(k2,2j)\displaystyle{\mathrm{IKM}}^{\mathrm{reg}}_{k}\left(\frac{k}{2},2j\right) :=limtε0+(εt(I0K0)2r+1s2jds2γk,jr2k2jεtdssm=0jr1γk,jr1mt2m+22k2j+2m+1(2m+2))\displaystyle:=\lim_{\begin{subarray}{c}t\rightarrow\infty\\ \varepsilon\rightarrow 0^{+}\end{subarray}}\left(\int_{\varepsilon}^{t}(I_{0}K_{0})^{2r+1}s^{2j}{\mathrm{d}}s-\frac{2\gamma_{k,j-r}}{2^{k-2j}}\int_{\varepsilon}^{t}\frac{\mathrm{d}s}{s}-\sum_{m=0}^{j-r-1}\frac{\gamma_{k,j-r-1-m}t^{2m+2}}{2^{k-2j+2m+1}(2m+2)}\right)
if k=4r+2.\displaystyle\hskip 284.52756pt{\text{if }}k=4r+2.
Proof.

Near 0, by [12, §10.30(i)] we have the asymptotics

I0(t)\displaystyle I_{0}(t) =1+O(t2);\displaystyle=1+O(t^{2}); (17)
K0(t)\displaystyle K_{0}(t) =(γ+logt2)+O(t2logt),\displaystyle=-\left(\gamma+\log\frac{t}{2}\right)+O(t^{2}\log t), (18)

where γ\gamma is the Euler constant. Then, the integral 01I0a(t)K0ka(t)tbdt\int_{0}^{1}I_{0}^{a}(t)K_{0}^{k-a}(t)t^{b}{\mathrm{d}}t converges for all 0ak20\leq a\leq\frac{k}{2} and any b0b\geq 0.

Near \infty, from (8) and (9), when 0ak0\leq a\leq k^{\prime}, I0a(t)K0ka(t)I_{0}^{a}(t)K_{0}^{k-a}(t) decays exponentially and hence the integral 1I0a(t)K0ka(t)tbdt\int_{1}^{\infty}I_{0}^{a}(t)K_{0}^{k-a}(t)t^{b}{\mathrm{d}}t converges.

When kk is even and a=k2a=\frac{k}{2}, near \infty, by (11), we have the asymptotic expansion

(I0K0)k/2t2j=12k/2n=0γk,n4nt2nk/2+2j.\left(I_{0}K_{0}\right)^{k/2}t^{2j}=\frac{1}{2^{k/2}}\sum_{n=0}^{\infty}\gamma_{k,n}4^{n}t^{-2n-k/2+2j}.

Taking integration, we have

εt(I0K0)k/2s2jds=12k/2n=0γk,n4nεts2nk/2+2jds.\int_{\varepsilon}^{t}(I_{0}K_{0})^{k/2}s^{2j}\mathrm{d}s=\frac{1}{2^{k/2}}\sum_{n=0}^{\infty}\gamma_{k,n}4^{n}\int_{\varepsilon}^{t}s^{-2n-k/2+2j}\mathrm{d}s.

Using the fact that 1tαdt\int_{1}^{\infty}t^{\alpha}{\mathrm{d}}t converges if and only if α<1\alpha<-1 and 01tαdt\int_{0}^{1}t^{\alpha}{\mathrm{d}}t converges if and only if α>1\alpha>-1, the divergent part of the integral εt(I0K0)k/2s2jds\int_{\varepsilon}^{t}\left(I_{0}K_{0}\right)^{k/2}s^{2j}{\mathrm{d}}s as tt\rightarrow\infty, ε0+\varepsilon\rightarrow 0^{+} is

m=0jr1γk,jr1m2k2j+2mt2m+1(2m+1)\displaystyle\sum_{m=0}^{j-r-1}\frac{\gamma_{k,j-r-1-m}}{2^{k-2j+2m}}\frac{t^{2m+1}}{\left(2m+1\right)} if k=4r+4\displaystyle{\text{ if }}k=4r+4 (19)
2γk,jr2k2jεtdss+m=0jr1γk,jr1m2k2j+2m+1t2m+2(2m+2)\displaystyle\frac{2\gamma_{k,j-r}}{2^{k-2j}}\int_{\varepsilon}^{t}\frac{\mathrm{d}s}{s}+\sum_{m=0}^{j-r-1}\frac{\gamma_{k,j-r-1-m}}{2^{k-2j+2m+1}}\frac{t^{2m+2}}{(2m+2)} if k=4r+2\displaystyle{\text{ if }}k=4r+2 (20)

Hence, after subtracting the divergent part of the integral, we conclude that the limits IKMkreg(k2,2j)\mathrm{IKM}_{k}^{\mathrm{reg}}(\frac{k}{2},2j) exist. ∎

Remark 22.

For a=k2a=\frac{k}{2} and bkb\geq k^{\prime} with odd bb, the integral 0I0a(t)K0ka(t)tbdt\int_{0}^{\infty}I_{0}^{a}(t)K_{0}^{k-a}(t)t^{b}{\mathrm{d}}t also diverges. We may similarly define the regularized Bessel moments in this case. See [2, Definitions 6.1 and 6.4].

5.2 Period pairing and compactly supported period pairing

By [2, Corollary 2.11], there exist the following two perfect pairings. The period pairing is defined to be

H1rd(𝔾m,(zSymkKl2))×HdR1(𝔾m,zSymkKl2){{H_{1}^{\mathrm{rd}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}_{\mathbb{C}}\times H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\mathbb{C}}}}{\mathbb{C}},per\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{per}}}

by

σ1ze0be1kb,ωper=σ1ze0be1kb,ωtop.\left\langle\sigma\otimes\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b},\omega\right\rangle_{\mathrm{per}}=\int_{\sigma}\frac{1}{\sqrt{z}}\left\langle e_{0}^{b}e_{1}^{k-b},\omega\right\rangle_{\mathrm{top}}.

Here, the notation VV_{\mathbb{C}} means VV\otimes_{\mathbb{Q}}\mathbb{C}. There is a one-form ω\omega occurs in e0be1kb,ωtop\langle e_{0}^{b}e_{1}^{k-b},\omega\rangle_{\mathrm{top}}. This topological pairing means e0be1kb,ftopdz\langle e_{0}^{b}e_{1}^{k-b},f\rangle_{\mathrm{top}}\mathrm{d}z whenever ω=fdz\omega=f\mathrm{d}z. That is, we take the pairing ,top\langle\ ,\ \rangle_{\mathrm{top}} only on the coefficients. Note that the period pairing induces on the middle part Betti homology H1mid(𝔾m,(zSymkKl2))H_{1}^{\mathrm{mid}}(\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla})_{\mathbb{C}} and middle part de Rham cohomology Hmid1(𝔾m,zSymkKl2)H^{1}_{\mathrm{mid}}(\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})_{\mathbb{C}} by the restriction:

H1mid(𝔾m,(zSymkKl2))×Hmid1(𝔾m,zSymkKl2){{H_{1}^{\mathrm{mid}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}_{\mathbb{C}}\times H^{1}_{\mathrm{mid}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\mathbb{C}}}}.{\mathbb{C}.},per\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{per}}}

Moreover, the compactly supported period pairing is defined to be

HdR,c1(𝔾m,zSymkKl2)×H1mod(𝔾m,(zSymkKl2)){{H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}_{\mathbb{C}}\times H_{1}^{\mathrm{mod}}\big{(}\mathbb{G}_{\mathrm{m}},(\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2})^{\nabla}\big{)}_{\mathbb{C}}}}{\mathbb{C}},per,c\scriptstyle{\left\langle\ ,\ \right\rangle_{\mathrm{per,c}}}

by

(ξ,η,ω),σ1ze0be1kbper,c\displaystyle\left\langle(\xi,\eta,\omega),\sigma\otimes\frac{1}{\sqrt{z}}e_{0}^{b}e_{1}^{k-b}\right\rangle_{\mathrm{per,c}} =σ1zω,e0be1kbtop1zη,e0be1kbtop+1zξ,e0be1kbtop.\displaystyle=\int_{\sigma}\frac{1}{\sqrt{z}}\left\langle\omega,e_{0}^{b}e_{1}^{k-b}\right\rangle_{\mathrm{top}}-\frac{1}{\sqrt{z}}\left\langle\eta,e_{0}^{b}e_{1}^{k-b}\right\rangle_{\mathrm{top}}+\frac{1}{\sqrt{z}}\left\langle\xi,e_{0}^{b}e_{1}^{k-b}\right\rangle_{\mathrm{top}}.
Remark 23.

Note that the order of homology and cohomology in these two pairing are different. This is because we want to write down the matrix expression of quadratic relation (22) preventing the transpose notation.

Proposition 24.

The period pairing of the rapid decay cycle δb\delta_{b} in (13) and the de Rham cohomology class ωk,j\omega_{k,j} in Definition 10 is given by

δb,v0kzjdzzper=(π1)b(1)kb2k2jIKMk(b,2j)\left\langle\delta_{b},v_{0}^{k}z^{j}\frac{\mathrm{d}z}{z}\right\rangle_{\mathrm{per}}=(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k-2j}{\mathrm{IKM}}_{k}(b,2j)

for 0bk0\leq b\leq k^{\prime} and 0jk0\leq j\leq k^{\prime}.

Proof.

Denote εσ0\varepsilon\sigma_{0} to be the scaling of the chain σ0\sigma_{0}, that is, εσ0\varepsilon\sigma_{0} is a chain of a circle of radius ε\varepsilon. Similarly, denote εσ+\varepsilon\sigma_{+} to be the chain of the ray [ε,)[\varepsilon,\infty). Then, since σ0\sigma_{0} and σ+\sigma_{+} are homotopy to εσ0\varepsilon\sigma_{0} and σ+\sigma_{+} respectively, we may replace σ0\sigma_{0} and σ+\sigma_{+} in δb\delta_{b} by εσ0\varepsilon\sigma_{0} and εσ+\varepsilon\sigma_{+} respectively in the following computation. We compute

δb,v0kzjdzzper\displaystyle\left\langle\delta_{b},v_{0}^{k}z^{j}\frac{\mathrm{d}z}{z}\right\rangle_{\mathrm{per}} =ε(σ+12σ0+n=1kbdkb(n)σ02n)1ze0be1kb,v0kzjdzztop\displaystyle=\int_{\varepsilon\left(\sigma_{+}-\frac{1}{2}\sigma_{0}+\sum_{n=1}^{k-b}d_{k-b}(n)\sigma_{0}^{2n}\right)}\frac{1}{\sqrt{z}}\left\langle e_{0}^{b}e_{1}^{k-b},v_{0}^{k}z^{j}\frac{\mathrm{d}z}{z}\right\rangle_{\mathrm{top}}
=(1)kb(π1)bε(σ+12σ0+n=1kbdkb(n)σ02n)z(A0)bB0kbzj1dz\displaystyle=(-1)^{k-b}(\pi\sqrt{-1})^{b}\int_{\varepsilon\left(\sigma_{+}-\frac{1}{2}\sigma_{0}+\sum_{n=1}^{k-b}d_{k-b}(n)\sigma_{0}^{2n}\right)}\sqrt{z}(-A_{0})^{b}B_{0}^{k-b}z^{j-1}\mathrm{d}z
=(π1)b(1)kb2kε(σ+12σ0+n=1kbdkb(n)σ02n)zj1zI0(2z)bK0(2z)kbdz\displaystyle=(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k}\int_{\varepsilon\left(\sigma_{+}-\frac{1}{2}\sigma_{0}+\sum_{n=1}^{k-b}d_{k-b}(n)\sigma_{0}^{2n}\right)}z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\mathrm{d}z
=(π1)b(1)kb2kεzj1zI0(2z)bK0(2z)kbdz\displaystyle=(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k}\int_{\varepsilon}^{\infty}z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\mathrm{d}z
12(π1)b(1)kb2kεσ0zj1zI0(2z)bK0(2z)kbdz\displaystyle\hskip 20.0pt-\frac{1}{2}(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k}\int_{\varepsilon\sigma_{0}}z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\mathrm{d}z
+n=1kbdkb(n)(π1)b(1)kb2kεσ02nzj1zI0(2z)bK0(2z)kbdz.\displaystyle\hskip 20.0pt+\sum_{n=1}^{k-b}d_{k-b}(n)(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k}\int_{\varepsilon\sigma_{0}^{2n}}z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\mathrm{d}z.

Changing the coordinate by z=t24z=\frac{t^{2}}{4}, the first term becomes

(π1)b(1)kb2k2j2εI0(t)bK0(t)kbt2jdt.(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k-2j}\int_{2\sqrt{\varepsilon}}^{\infty}I_{0}(t)^{b}K_{0}(t)^{k-b}t^{2j}\mathrm{d}t.

When ε0+\varepsilon\rightarrow 0^{+}, this term tends to (π1)b(1)kb2k2jIKMk(b,2j)(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k-2j}{\mathrm{IKM}}_{k}(b,2j).

For the other two terms, εσ0pzj1zI0(2z)bK0(2z)kbdz\int_{\varepsilon\sigma_{0}^{p}}z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\mathrm{d}z tends to zero as ε0+\varepsilon\rightarrow 0^{+} for the following reason. As s0+s\rightarrow 0^{+}, we have the asymptotic expansions (17) and (18). Then, as ε0+\varepsilon\rightarrow 0^{+} for all j0j\geq 0, we have the estimate

|εσ0pzj1zI0(2z)bK0(2z)kbdz|\displaystyle\left|\int_{\varepsilon\sigma_{0}^{p}}z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\mathrm{d}z\right| εσ0p|zj1zI0(2z)bK0(2z)kb||dz|\displaystyle\leq\int_{\varepsilon\sigma_{0}^{p}}\left|z^{j-1}\sqrt{z}I_{0}(2\sqrt{z})^{b}K_{0}(2\sqrt{z})^{k-b}\right|\left|\mathrm{d}z\right|
02πpεj1ε|I0(2εeiθ)bK0(2εeiθ)kb|εdθ\displaystyle\leq\int_{0}^{2\pi p}\varepsilon^{j-1}\sqrt{\varepsilon}\left|I_{0}\left(2\sqrt{\varepsilon e^{i\theta}}\right)^{b}K_{0}\left(2\sqrt{\varepsilon e^{i\theta}}\right)^{k-b}\right|\varepsilon{\mathrm{d}}\theta
εjε02πp|γ+logεeiθ|kbdθ\displaystyle\leq\varepsilon^{j}\sqrt{\varepsilon}\int_{0}^{2\pi p}\left|\gamma+\log\sqrt{\varepsilon e^{i\theta}}\right|^{k-b}{\mathrm{d}}\theta
=εjε02πp|γ+logε+12iθ|kbdθ0.\displaystyle=\varepsilon^{j}\sqrt{\varepsilon}\int_{0}^{2\pi p}\left|\gamma+\log\sqrt{\varepsilon}+\frac{1}{2}i\theta\right|^{k-b}{\mathrm{d}}\theta\rightarrow 0.\qed
Proposition 25.

The compactly supported period pairing of the compactly supported de Rham cohomology ω~k,j\widetilde{\omega}_{k,j} in Definition 11 and moderate decay cycle γa\gamma_{a} in (14) is given by

ω~k,j,γaper,c=2k2j(1)ka(π1)aIKM,\left\langle\widetilde{\omega}_{k,j},\gamma_{a}\right\rangle_{\mathrm{per,c}}=2^{k-2j}(-1)^{k-a}(\pi\sqrt{-1})^{a}\cdot{\mathrm{IKM}},

where 0ak/2,0jk0\leq a\leq\lfloor k/2\rfloor,0\leq j\leq k^{\prime} with jrj\neq r if k2mod4k\equiv 2\bmod{4}, and

IKM={IKMkreg(a,2j)if 4k,a=k/2,r+1jk,IKMk(a,2j)γk,jk/222jkIKMk(a,k)if 4(k+2),0ak,r+1jk,IKMkreg(a,2j)γk,jk/222jkIKMkreg(a,k)if 4(k+2),a=k+1,r+1jk,IKMk(a,2j)otherwise.{\mathrm{IKM}}=\begin{cases}{\mathrm{IKM}}^{\text{reg}}_{k}(a,2j)&\text{if $4\mid k,a=k/2,r+1\leq j\leq k^{\prime}$},\\ {\mathrm{IKM}}_{k}(a,2j)-\gamma_{k,j-k^{\prime}/2}2^{2j-k^{\prime}}{\mathrm{IKM}}_{k}(a,k^{\prime})&\text{if $4\mid(k+2),0\leq a\leq k^{\prime},r+1\leq j\leq k^{\prime}$},\\ {\mathrm{IKM}}^{\text{reg}}_{k}(a,2j)-\gamma_{k,j-k^{\prime}/2}2^{2j-k^{\prime}}{\mathrm{IKM}}^{\text{reg}}_{k}(a,k^{\prime})&\text{if $4\mid(k+2),a=k^{\prime}+1,r+1\leq j\leq k^{\prime}$},\\ {\mathrm{IKM}}_{k}(a,2j)&\text{otherwise}.\end{cases}

Moreover, when k=4r+2k=4r+2, we have

m^2r+1,γaper,c=δa,2r+1(π1)a2k1(kk/2).\left\langle\widehat{m}_{2r+1},\gamma_{a}\right\rangle_{\mathrm{per,c}}=\delta_{a,2r+1}(\pi\sqrt{-1})^{a}2^{k}\frac{1}{\binom{k}{k/2}}.
Proof.

When k1,3mod4k\equiv 1,3\bmod{4}. We compute the compactly supported period pairing

(ξj,ηj,ωk,j),γaper,c\displaystyle\left\langle\left(\xi_{j},\eta_{j},\omega_{k,j}\right),\gamma_{a}\right\rangle_{\mathrm{per,c}} =+1zv0k,e0ae1katopzjdzz1zc=0k(kc)ηj,cze0kce¯1c,e0ae1katop\displaystyle=\int_{\mathbb{R}_{+}}\frac{1}{\sqrt{z}}\left\langle v_{0}^{k},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}z^{j}\frac{{\mathrm{d}}z}{z}-\frac{1}{\sqrt{z}}\left\langle-\sum_{c=0}^{k}\binom{k}{c}\frac{\eta_{j,c}}{\sqrt{z}}e_{0}^{k-c}\overline{e}_{1}^{c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
+1zc=0kξj,cv0cv1kc,e0ae1katop\displaystyle\hskip 20.0pt+\frac{1}{\sqrt{z}}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
=+1zc=0k(kc)(A0(z))cB0(z)kce0kce¯1c,e0ae1katopzj1dz\displaystyle=\int_{\mathbb{R}_{+}}\frac{1}{\sqrt{z}}\left\langle\sum_{c=0}^{k}\binom{k}{c}\left(-A_{0}(z)\right)^{c}B_{0}(z)^{k-c}e_{0}^{k-c}\overline{e}_{1}^{c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}z^{j-1}{\mathrm{d}}z
+(π1)a(1)aηj,a+1zc=0kξj,cv0cv1kc,e0ae1katop\displaystyle\hskip 20.0pt+(\pi\sqrt{-1})^{a}(-1)^{a}\eta_{j,a}+\frac{1}{\sqrt{z}}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
=(1)a(π1)a+(A0(z))aB0(z)kazzj1dz\displaystyle=(-1)^{a}(\pi\sqrt{-1})^{a}\int_{\mathbb{R}_{+}}\left(-A_{0}(z)\right)^{a}B_{0}(z)^{k-a}\sqrt{z}\,z^{j-1}{\mathrm{d}}z
+(1)a(π1)aηj,a+1zc=0kξj,cv0cv1kc,e0ae1katop\displaystyle\hskip 20.0pt+(-1)^{a}(\pi\sqrt{-1})^{a}\eta_{j,a}+\frac{1}{\sqrt{z}}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
=(1)a(π1)a2k2j+I0(s)aK0(s)kas2jds+(1)a(π1)aηj,a\displaystyle=(-1)^{a}(\pi\sqrt{-1})^{a}2^{k-2j}\int_{\mathbb{R}_{+}}I_{0}(s)^{a}K_{0}(s)^{k-a}s^{2j}{\mathrm{d}}s+(-1)^{a}(\pi\sqrt{-1})^{a}\eta_{j,a}
+2sc=0kξj,cv0cv1kc,e0ae1katop\displaystyle\hskip 20.0pt+\frac{2}{s}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}

where the last equality follows by the change of variable z=s24z=\frac{s^{2}}{4}. The first term converges by Lemma 21. Since k>2ak>2a, by (12), the second term tends to zero as ss\rightarrow\infty. The third term tends to zero as s0s\rightarrow 0 since all ξj,cs24\xi_{j,c}\in\mathbb{Q}\llbracket\frac{s^{2}}{4}\rrbracket and the topological pairing gives a factor s24\frac{s^{2}}{4}.

When k0mod4k\equiv 0\bmod{4}, write k=4r+4k=4r+4. We compute the compactly supported period pairing

(ξj,ηj,ωk,j),γaper,c\displaystyle\left\langle\left(\xi_{j},\eta_{j},\omega_{k,j}\right),\gamma_{a}\right\rangle_{\mathrm{per,c}} =+1zv0k,e0ae1katopzjdzz1zc=0k(kc)ηj,cze0kce¯1c,e0ae1katop\displaystyle=\int_{\mathbb{R}_{+}}\frac{1}{\sqrt{z}}\left\langle v_{0}^{k},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}z^{j}\frac{{\mathrm{d}}z}{z}-\frac{1}{\sqrt{z}}\left\langle-\sum_{c=0}^{k}\binom{k}{c}\frac{\eta_{j,c}}{\sqrt{z}}e_{0}^{k-c}\overline{e}_{1}^{c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
+1zc=0kξj,cv0cv1kc,e0ae1katop\displaystyle\hskip 20.0pt+\frac{1}{\sqrt{z}}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
=2k2j(1)a(π1)a+I0(s)aK0(s)kas2jds\displaystyle=2^{k-2j}(-1)^{a}(\pi\sqrt{-1})^{a}\int_{\mathbb{R}_{+}}I_{0}(s)^{a}K_{0}(s)^{k-a}s^{2j}{\mathrm{d}}s
+(1)a(π1)aηj,a+2sc=0kξj,cv0cv1kc,e0ae1katop\displaystyle\hskip 20.0pt+(-1)^{a}(\pi\sqrt{-1})^{a}\eta_{j,a}+\frac{2}{s}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}

where the last equality is the change of variable z=s24z=\frac{s^{2}}{4}. The third term tends to zero as s0s\rightarrow 0 since all ξj,cs24\xi_{j,c}\in\mathbb{Q}\llbracket\frac{s^{2}}{4}\rrbracket and the topological pairing gives a factor s24\frac{s^{2}}{4}. By the same argument above, when a=0,1,,k22=ka=0,1,\cdots,\frac{k-2}{2}=k^{\prime}, that is, k>2ak>2a, we have that the first term converges and the second term tends to zero as ss\rightarrow\infty.

Now, we turn to analyze the case that a=k2a=\frac{k}{2}. The pairing becomes

(ξj,ηj,ωk,j),γaper,c=2k2j(π1)a0sI0(t)aK0(t)kas2jds+(π1)aηj,2r+2\left\langle\left(\xi_{j},\eta_{j},\omega_{k,j}\right),\gamma_{a}\right\rangle_{\mathrm{per,c}}=2^{k-2j}(-\pi\sqrt{-1})^{a}\int_{0}^{s}I_{0}(t)^{a}K_{0}(t)^{k-a}s^{2j}{\mathrm{d}}s+(-\pi\sqrt{-1})^{a}\eta_{j,2r+2}

This term converges as ss\rightarrow\infty for the following reason:

The singular part of the integral (I0K0)2r+2s2j\left(I_{0}K_{0}\right)^{2r+2}s^{2j} is given by (19) and ηj,2r+2\eta_{j,2r+2} has expansion

ηj,2r+222r2j+1rj+1/2s2j2r1Gi22r2j+1n=022nγk,nrj+1/2+ns2j2r12n\eta_{j,2r+2}\sim\frac{2^{2r-2j+1}}{r-j+1/2}s^{2j-2r-1}\cdot G_{i}\sim 2^{2r-2j+1}\sum_{n=0}^{\infty}\frac{2^{2n}\gamma_{k,n}}{r-j+1/2+n}s^{2j-2r-1-2n}

Thus, both of the singular terms cancel.

When k2mod4k\equiv 2\bmod{4}, write k=4r+2k=4r+2. Recall from Definition 11 the elements ω~k,j\widetilde{\omega}_{k,j} and m^2r+1\widehat{m}_{2r+1} in HdR,c1(𝔾m,zSymkKl2)H^{1}_{{\mathrm{dR,c}}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}. If we use the convention that γk,p=0\gamma_{k,p}=0 whenever p<0p<0, we rewrite

ω~k,i\displaystyle\widetilde{\omega}_{k,i} =(ξi,ηi,ωk,i)\displaystyle=\left(\xi_{i},\eta_{i},\omega_{k,i}\right)
=(a=0kξi,a(z)v0av1kaγk,ira=0kξr,a(z)v0av1ka,\displaystyle=\left(\sum_{a=0}^{k}\xi_{i,a}(z)v_{0}^{a}v_{1}^{k-a}-\gamma_{k,i-r}\sum_{a=0}^{k}\xi_{r,a}(z)v_{0}^{a}v_{1}^{k-a},\right.
a=0ak/2k(ka)ηi,aγk,irηr,aze0kae¯1a(kk/2)ηi,2r+1z(e0e¯1)2r+1,\displaystyle\hskip 20.0pt-\sum_{\begin{subarray}{c}a=0\\ a\neq k/2\end{subarray}}^{k}\binom{k}{a}\frac{\eta_{i,a}-\gamma_{k,i-r}\eta_{r,a}}{\sqrt{z}}e_{0}^{k-a}\overline{e}_{1}^{a}-\binom{k}{k/2}\frac{\eta_{i,2r+1}}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1},
v0kzidzzγk,irv0kzrdzz)\displaystyle\hskip 20.0pt\left.v_{0}^{k}z^{i}\frac{\mathrm{d}z}{z}-\gamma_{k,i-r}v_{0}^{k}z^{r}\frac{\mathrm{d}z}{z}\right)

In the pairing ω~k,j,γaper,c\left\langle\widetilde{\omega}_{k,j},\gamma_{a}\right\rangle_{\mathrm{per,c}}, the third term

2sc=0kξj,cv0cv1kc,e0ae1katop\frac{2}{s}\left\langle\sum_{c=0}^{k}\xi_{j,c}v_{0}^{c}v_{1}^{k-c},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}

tends to zero as s0s\rightarrow 0 since all ξj,cs24\xi_{j,c}\in\mathbb{C}\llbracket\frac{s^{2}}{4}\rrbracket and the topological pairing gives a factor s24\frac{s^{2}}{4}. The other two terms are equal to

(+1zv0k,e0ae1katopzjdzzγk,jr+1zv0k,e0ae1katopzrdzz)\displaystyle\left(\int_{\mathbb{R}_{+}}\frac{1}{\sqrt{z}}\left\langle v_{0}^{k},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}z^{j}\frac{{\mathrm{d}}z}{z}-\gamma_{k,j-r}\int_{\mathbb{R}_{+}}\frac{1}{\sqrt{z}}\left\langle v_{0}^{k},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}z^{r}\frac{{\mathrm{d}}z}{z}\right)
1zb=0bk/2k(kb)ηj,bγk,jrηr,bze0kbe¯1b(kk/2)ηj,2r+1z(e0e¯1)2r+1,e0ae1katop\displaystyle-\frac{1}{\sqrt{z}}\left\langle-\sum_{\begin{subarray}{c}b=0\\ b\neq k/2\end{subarray}}^{k}\binom{k}{b}\frac{\eta_{j,b}-\gamma_{k,j-r}\eta_{r,b}}{\sqrt{z}}e_{0}^{k-b}\overline{e}_{1}^{b}-\binom{k}{k/2}\frac{\eta_{j,2r+1}}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
=(1)a(π1)a(2k2j+I0(s)aK0(s)kas2jds2k2rγk,jr+I0(s)aK0(s)kas2rds)\displaystyle=(-1)^{a}(\pi\sqrt{-1})^{a}\left(2^{k-2j}\int_{\mathbb{R}_{+}}I_{0}(s)^{a}K_{0}(s)^{k-a}s^{2j}{\mathrm{d}}s-2^{k-2r}\gamma_{k,j-r}\int_{\mathbb{R}_{+}}I_{0}(s)^{a}K_{0}(s)^{k-a}s^{2r}{\mathrm{d}}s\right)
+1zb=0bk/2k(kb)(ηj,bγk,jrηr,b)e0kbe¯1b+(kk/2)ηj,2r+1(e0e¯1)2r+1,e0ae1katop\displaystyle\hskip 20.0pt+\frac{1}{z}\left\langle\sum_{\begin{subarray}{c}b=0\\ b\neq k/2\end{subarray}}^{k}\binom{k}{b}\left(\eta_{j,b}-\gamma_{k,j-r}\eta_{r,b}\right)e_{0}^{k-b}\overline{e}_{1}^{b}+\binom{k}{k/2}\eta_{j,2r+1}(e_{0}\overline{e}_{1})^{2r+1},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}

We analyze the convergence of these terms. When 2a<k2a<k or j<rj<r, the integral 0tI0(s)aK0(s)kas2jds\int_{0}^{t}I_{0}(s)^{a}K_{0}(s)^{k-a}s^{2j}{\mathrm{d}}s converges as tt\rightarrow\infty by Lemma 21. The second term is equal to

(1)a(π1)a(ηj,aγk,jrηr,a).\displaystyle(-1)^{a}(\pi\sqrt{-1})^{a}\left(\eta_{j,a}-\gamma_{k,j-r}\eta_{r,a}\right).

By the expansion of ηi,a\eta_{i,a}:

ηi,aπk2ak2ae(k2a)s(4s2)k/4iGi,a,\eta_{i,a}\sim\frac{\sqrt{\pi}^{k-2a}}{k-2a}e^{-(k-2a)s}\left(\frac{4}{s^{2}}\right)^{k/4-i}\cdot G_{i,a},

where Gi,a1+2s2sG_{i,a}\in 1+\frac{2}{s}\mathbb{Q}\llbracket\frac{2}{s}\rrbracket, this term tends to 0 as ss\rightarrow\infty.

When a=k2a=\frac{k}{2} and jrj\geq r, the integral 0tI0(s)aK0(s)kas2jds\int_{0}^{t}I_{0}(s)^{a}K_{0}(s)^{k-a}s^{2j}{\mathrm{d}}s has the singular part (20). The second term is equal to

(1)2r+1(π1)2r+1ηj,2r+1=(1)2r+1(π1)2r+11rj(4s2)rjHi(-1)^{2r+1}(\pi\sqrt{-1})^{2r+1}\eta_{j,2r+1}=(-1)^{2r+1}(\pi\sqrt{-1})^{2r+1}\frac{1}{r-j}\left(\frac{4}{s^{2}}\right)^{r-j}\cdot H_{i}

where Hi1+4s24s2H_{i}\in 1+\frac{4}{s^{2}}\mathbb{Q}\llbracket\frac{4}{s^{2}}\rrbracket. Thus, the singular part of this term is

(1)2r+1(π1)2r+1n=1jrγk,jrnn(4s2)n.(-1)^{2r+1}(\pi\sqrt{-1})^{2r+1}\sum_{n=1}^{j-r}\frac{-\gamma_{k,j-r-n}}{n}\left(\frac{4}{s^{2}}\right)^{-n}.

In consequence, the singular parts cancel.

Finally, for a=0,1,,k2a=0,1,\cdots,\frac{k}{2}, we have

(0,2kz(e0e¯1)2r+1,0),γaper,c\displaystyle\left\langle\left(0,\frac{2^{k}}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1},0\right),\gamma_{a}\right\rangle_{\mathrm{per,c}} =2kz1z(e0e¯1)2r+1,e0ae1katop\displaystyle=-\frac{2^{k}}{\sqrt{z}}\left\langle\frac{1}{\sqrt{z}}(e_{0}\overline{e}_{1})^{2r+1},e_{0}^{a}e_{1}^{k-a}\right\rangle_{\mathrm{top}}
=δa,2r+1(π1)a2k1(kk/2).\displaystyle=\delta_{a,2r+1}(\pi\sqrt{-1})^{a}2^{k}\frac{1}{\binom{k}{k/2}}.\qed
Corollary 26.

The period matrix of the period pairing with respective to the bases {δb}b=0k\left\{\delta_{b}\right\}_{b=0}^{k^{\prime}} of H1rdH_{1}^{\mathrm{rd}} and {ωk,j}j=0k\left\{\omega_{k,j}\right\}_{j=0}^{k^{\prime}} of HdR1H^{1}_{\mathrm{dR}} is P=(Pbj)P=(P_{bj}), where

Pbj=δb,v0kzjdzzper=(π1)b(1)kb2k2jIKMk(b,2j)P_{bj}=\left\langle\delta_{b},v_{0}^{k}z^{j}\frac{\mathrm{d}z}{z}\right\rangle_{\mathrm{per}}=(\pi\sqrt{-1})^{b}(-1)^{k-b}2^{k-2j}{\mathrm{IKM}}_{k}(b,2j)

for 0bk0\leq b\leq k^{\prime} and 0jk0\leq j\leq k^{\prime}. Moreover, PP is invertible.

Remark 27 (determinant of the period matrix).

In fact,

detP\displaystyle\det P =(π1)k(k+1)/2(1)(2kk)(k+1)/22(kk)(k+1)det(IKMk(b,2j)),\displaystyle=(\pi\sqrt{-1})^{k^{\prime}(k^{\prime}+1)/2}(-1)^{(2k-k^{\prime})(k^{\prime}+1)/2}2^{(k-k^{\prime})(k^{\prime}+1)}\det\left({\mathrm{IKM}}_{k}(b,2j)\right),

where

det(IKMk(b,2j))={det(Mk+1)if k is odd;det(Nk+1)if k is even.\displaystyle\det\left({\mathrm{IKM}}_{k}(b,2j)\right)=\begin{cases}\det\left(M_{k^{\prime}+1}\right)&{\text{if }}k{\text{ is odd;}}\\ \det\left(N_{k^{\prime}+1}\right)&{\text{if }}k{\text{ is even.}}\\ \end{cases}

The definition of MrM_{r} and NrN_{r} are given in Appendix A.2 and their determinants are given in Corollary 39 explicitly.

5.3 \mathbb{Q}-linear and quadratic relations on Bessel moments

We have now developed all the tools and computations to see the wonderful results in \mathbb{Q}-linear and quadratic relations on Bessel moments.

Corollary 28.

For k=4r+4k=4r+4,

j=0r(k/22j)(1)jπ2jIKMk(2j,2i)={(1)rπ2r+2IKMk(2r+2,2i)if 0ir,(1)rπ2r+2IKMkreg(2r+2,2i)if r+1ik.\displaystyle\sum_{j=0}^{r}\binom{k/2}{2j}(-1)^{j}\pi^{2j}{\mathrm{IKM}}_{k}(2j,2i)=\begin{cases}(-1)^{r}\pi^{2r+2}{\mathrm{IKM}}_{k}(2r+2,2i)&\text{if $0\leq i\leq r$},\\ (-1)^{r}\pi^{2r+2}{\mathrm{IKM}}_{k}^{\mathrm{reg}}(2r+2,2i)&\text{if $r+1\leq i\leq k^{\prime}$}.\end{cases}

For k=4r+2k=4r+2,

j=0r(k/22j)(1)jπ2jIKMk(2j,2i)={0if 0ir1,γk,ir22i2rj=0r(k/22j)(1)jπ2jIKMk(2j,2r)if r+1i2r.\displaystyle\sum_{j=0}^{r}\binom{k/2}{2j}(-1)^{j}\pi^{2j}{\mathrm{IKM}}_{k}(2j,2i)=\begin{cases}0&\text{if $0\leq i\leq r-1$},\\ \gamma_{k,i-r}2^{2i-2r}\displaystyle\sum_{j=0}^{r}\binom{k/2}{2j}(-1)^{j}\pi^{2j}{\mathrm{IKM}}_{k}(2j,2r)&\text{if $r+1\leq i\leq 2r$}.\end{cases} (21)
Proof.

By Lemma 17, we know that

j=0k/4(k/22j)γ2j\displaystyle\sum_{j=0}^{k/4}\binom{k/2}{2j}\gamma_{2j} =0 if k0mod4;\displaystyle=0{\text{ if }}k\equiv 0\bmod{4};
j=0(k2)/4(k/22j)γ2j\displaystyle\sum_{j=0}^{\left(k-2\right)/4}\binom{k/2}{2j}\gamma_{2j} =0 if k2mod4.\displaystyle=0{\text{ if }}k\equiv 2\bmod{4}.

Then take the pairing with ω~k,i\widetilde{\omega}_{k,i} in the compactly supported de Rham cohomology. Combining with the result of Proposition 25, we obtain the desired algebraic relation. ∎

Remark 29.

The above linear algebraic relations for ii in the range 0ir0\leq i\leq r, under the name sum rule identities, are previously proved by analytic method in [6] (see [6, (1.3)] for k2mod4k\equiv 2\bmod{4} and [6, (1.5)] for k0mod4k\equiv 0\bmod{4}).

Corollary 30.

For any kk and any 0ak0\leq a\leq k^{\prime}, the dimension of the \mathbb{Q}-vector space generated by the Bessel moments has an upper bound:

dimspan{IKMk(a,2j)j{0}}k+1.\dim{\mathrm{span}}_{\mathbb{Q}}\left\{{\mathrm{IKM}}_{k}(a,2j)\mid j\in\left\{0\right\}\cup\mathbb{N}\right\}\leq k^{\prime}+1.

If kk is even, the dimension of the \mathbb{Q}-vector space generated by the regularized Bessel moments has an upper bound:

dimspan{IKMkreg(k/2,2j)j{0}}k+1.\dim{\mathrm{span}}_{\mathbb{Q}}\left\{{\mathrm{IKM}}^{\mathrm{reg}}_{k}(k/2,2j)\mid j\in\left\{0\right\}\cup\mathbb{N}\right\}\leq k^{\prime}+1.

Here when 0jk14=r0\leq j\leq\left\lfloor\frac{k-1}{4}\right\rfloor=r, we do not need to regularize the Bessel moments, that is, IKMkreg(k/2,2j)=IKMk(k/2,2j){\mathrm{IKM}}^{\mathrm{reg}}_{k}(k/2,2j)={\mathrm{IKM}}_{k}(k/2,2j) (see Lemma 21).

Proof.

We know that the dimensions of HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} and HdR,c1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} are k+1k^{\prime}+1.

For each integer s>ks>k^{\prime}, since {v0kzjdzz}j=0,,k\left\{v_{0}^{k}z^{j}\frac{{\mathrm{d}}z}{z}\right\}_{j=0,\cdots,k^{\prime}} form a basis of HdR1(𝔾m,zSymkKl2)H^{1}_{\mathrm{dR}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)}, we may express v0kzsdzzv_{0}^{k}z^{s}\frac{{\mathrm{d}}z}{z} as the \mathbb{Q}-linear combination of the basis. Then after we take the period pairing between v0kzsdzzv_{0}^{k}z^{s}\frac{{\mathrm{d}}z}{z} and the rapid decay cycle δa\delta_{a} (see Proposition 24), the \mathbb{Q}-linear relation becomes a \mathbb{Q}-linear relation for the Bessel moments

{IKMk(a,2j)j=0,,k}{IKMk(a,2s)}.\left\{{\mathrm{IKM}}_{k}(a,2j)\mid j=0,\cdots,k^{\prime}\right\}\cup\left\{{\mathrm{IKM}}_{k}(a,2s)\right\}.

If kk is even, similarly, when s>ks>k^{\prime}, express ω~k,sHdR,c1(𝔾m,zSymkKl2)\widetilde{\omega}_{k,s}\in H^{1}_{\mathrm{dR,c}}\big{(}\mathbb{G}_{\mathrm{m}},\sqrt{z}\operatorname{Sym}^{k}{\mathrm{Kl}}_{2}\big{)} as the \mathbb{Q}-linear combination of the basis describe in Corollary 14. Then taking the compactly supported period pairing (see Proposition 25), the \mathbb{Q}-linear equivalence become the \mathbb{Q}-linear relation for the regularized Bessel moments

{IKMk(a,2j)j=0,,r1}{IKMkreg(a,2j)j=r,,k}{IKMkreg(a,2s)}.\left\{{\mathrm{IKM}}_{k}(a,2j)\mid j=0,\cdots,r-1\right\}\cup\left\{{\mathrm{IKM}}^{\mathrm{reg}}_{k}(a,2j)\mid j=r,\cdots,k^{\prime}\right\}\cup\left\{{\mathrm{IKM}}^{\mathrm{reg}}_{k}(a,2s)\right\}.\qed
Remark 31.

In [16], Borwein and Salvy provide a recurrence to find out the \mathbb{Q}-linear combination for Bessel moments by analyzing the symmetric power of the modified Bessel differential operator. Moreover, Zhou proves a similar result in [7] for the \mathbb{Q}-linear dependence for Bessel moments IKMk(a,2j1){\mathrm{IKM}}_{k}(a,2j-1). Our result is parallel to Zhou’s result.

Proposition 32.

With respect to the bases of H1rdH^{{\mathrm{rd}}}_{1}, H1modH^{{\mathrm{mod}}}_{1}, HdR1H_{\mathrm{dR}}^{1}, and HdR,c1H_{\mathrm{dR,c}}^{1} described in Corollaries 14, 20, we form the pairing matrices:

  1. 1.

    BB, the Betti intersection pairing matrix between H1rdH^{{\mathrm{rd}}}_{1} and H1modH^{{\mathrm{mod}}}_{1} in Proposition 18.

  2. 2.

    DD, the Poincaré pairing matrix between HdR,c1H_{\mathrm{dR,c}}^{1} and HdR1H_{\mathrm{dR}}^{1} in Proposition 13.

  3. 3.

    PP, the period pairing matrix between H1rdH^{{\mathrm{rd}}}_{1} and HdR1H_{\mathrm{dR}}^{1} in Proposition 24.

  4. 4.

    PcP_{\mathrm{c}}, the period pairing matrix between HdR,c1H_{\mathrm{dR,c}}^{1} and H1modH^{{\mathrm{mod}}}_{1} in Proposition 25.

  5. 5.

    BmidB_{\mathrm{mid}}, the Betti pairing matrix on H1midH^{\mathrm{mid}}_{1}.

  6. 6.

    DmidD_{\mathrm{mid}}, the Poincaré pairing matrix on Hmid1H_{\mathrm{mid}}^{1}.

  7. 7.

    PmidP_{\mathrm{mid}}, the period pairing matrix between H1midH^{\mathrm{mid}}_{1} and Hmid1H_{\mathrm{mid}}^{1}.111B,D,P,PcB,D,P,P_{\mathrm{c}} are square matrices of size k+1k^{\prime}+1 and that Bmid,Dmid,PmidB_{\mathrm{mid}},D_{\mathrm{mid}},P_{\mathrm{mid}} are of size k+1δ4+2,kk^{\prime}+1-\delta_{4\mathbb{Z}+2,k}. When k0,1,3mod4k\equiv 0,1,3\bmod{4}, we have B=BmidB=B_{\mathrm{mid}}, D=DmidD=D_{\mathrm{mid}}, and Pmid=P=PctP_{\mathrm{mid}}=P=P_{\mathrm{c}}^{t}.

Then we have the algebraic quadratic relations

PD1Pc\displaystyle PD^{-1}P_{\mathrm{c}} =(1)k(2π1)k+1B,\displaystyle=(-1)^{k}(2\pi\sqrt{-1})^{k+1}B, (22)
PmidDmid1Pmidt\displaystyle P_{\mathrm{mid}}D^{-1}_{\mathrm{mid}}P_{\mathrm{mid}}^{t} =(1)k(2π1)k+1Bmid.\displaystyle=(-1)^{k}(2\pi\sqrt{-1})^{k+1}B_{\mathrm{mid}}. (23)
Proof.

This quadratic relation is a general phenomenon on periods of meromorphic flat connection on complex manifolds. We refer to [8, Corollaries 2.14, 2.16] for more details. ∎

From this proposition, when k0,1,3mod4k\equiv 0,1,3\bmod{4}, we see the Bessel moments have quadratic relation given by (22). On the other hand, when k2mod4k\equiv 2\bmod{4}, the relation involves some combination of Bessel moments and regularized Bessel moments in the matrix PcP_{\mathrm{c}}. In the following discussion, we provide another expression of this relation, and we will see the pure quadratic relation involving only Bessel moments.

When k2mod4k\equiv 2\bmod{4}, write k=4r+2k=4r+2 and define two (k+1)×k(k^{\prime}+1)\times k^{\prime} matrices with rational coefficients:

Rk=(Ir00γk,1γk,kr0Ir),Lk=(0(k/22)0(k/24)0(k/2k)Ik).R_{k}=\left(\begin{array}[]{cccc}I_{r}&&0&\\ 0&-\gamma_{k,1}&\cdots&-\gamma_{k,k^{\prime}-r}\\ 0&&I_{r}&\end{array}\right),\quad L_{k}=\left(\begin{array}[]{ccccccc}0&-\binom{k/2}{2}&0&-\binom{k/2}{4}&\cdots&0&-\binom{k/2}{k^{\prime}}\\ &&&I_{k^{\prime}}&\end{array}\right).

By the linear relations (21) in Corollary 28, we have

PRk=LkPmid.PR_{k}=L_{k}P_{\mathrm{mid}}.

Also, PmidP_{\mathrm{mid}} is obtained by deleting the first row of LkPmidL_{k}P_{\mathrm{mid}}. Set B~=LkBmidLkt\widetilde{B}=L_{k}B_{\mathrm{mid}}L_{k}^{t} and D~=RkDmid1Rkt\widetilde{D}=R_{k}D^{-1}_{\mathrm{mid}}R_{k}^{t} which are square matrices of size k+1k^{\prime}+1 with rational coefficients. Then BmidB_{\mathrm{mid}} is obtained by deleting the first row and column from B~\widetilde{B}. Therefore, the quadratic relation (23) (involving linear combinations of Bessel moments) now becomes

PD~Pt=(1)k(2π1)k+1B~P\widetilde{D}P^{t}=(-1)^{k}(2\pi\sqrt{-1})^{k+1}\widetilde{B}

(involving pure Bessel moments).

Remark 33.

The matrices B~\widetilde{B} and D~\widetilde{D} in the above expression are singular because of the linear relations (21) in Corollary 28. This expression is equivalent to the middle part quadratic relation (23) together with linear relations (21).

Proposition 34.

When k=4r+2k=4r+2, the middle part period matrix is a k×kk^{\prime}\times k^{\prime} matrix given by

Pmid=(δb,ωk,iper)b=1,,k,i=0,,r^,,k.P_{\mathrm{mid}}=\left(\left\langle\delta_{b},\omega_{k,i}\right\rangle_{\mathrm{per}}\right)_{b=1,\cdots,k^{\prime},\ i=0,\cdots,\hat{r},\cdots,k^{\prime}}.

The determinant of this matrix PmidP_{\mathrm{mid}} is given by

detPmid=πr(k+1)1r(k1)2r(2r+1)r!a=1k(2a+1)k+1a(a+1)a+1.\det P_{\mathrm{mid}}=\pi^{r(k+1)}\sqrt{-1}^{r(k^{\prime}-1)}\frac{2^{r(2r+1)}}{r!}\prod_{a=1}^{k^{\prime}}\frac{(2a+1)^{k^{\prime}+1-a}}{(a+1)^{a+1}}.
Proof.

The matrix PmidP_{\mathrm{mid}} appears in the upper left of the compactly supported period pairing matrix PcP_{\mathrm{c}}. Just take determinant on (22) and then use the results of Propositions 13, 19, and Remark 27. ∎

Appendix A The Bessel operator and determinants of Bessel moments

A.1 Symmetric power of the modified Bessel differential operator

Consider the Weyl algebra t,t\mathbb{Q}\langle t,\partial_{t}\rangle consisting of ordinary differential operators. Write θ=tt\theta=t\partial_{t}. The modified Bessel differential operator is an element in the subalgebra t2,θ\mathbb{Q}\langle t^{2},\theta\rangle given by L2=θ2t2L_{2}=\theta^{2}-t^{2}. The corresponding solutions are the modified Bessel functions I0(t)I_{0}(t) and K0(t)K_{0}(t). The nn-th symmetric power Ln+1θ,t2L_{n+1}\in\mathbb{Q}\left\langle\theta,t^{2}\right\rangle of L2L_{2} has order n+1n+1 and the corresponding solutions are I0a(t)K0na(t)I_{0}^{a}(t)K_{0}^{n-a}(t) for 0an0\leq a\leq n. By [17, 16], the operator Ln+1=Ln+1,nL_{n+1}=L_{n+1,n} can be obtained by the recurrence relation as follows:

L0,n=1,L1,n=θ,Lk+1,n=θLk,nt2k(n+1k)Lk1,n, 1kn.\displaystyle\begin{split}L_{0,n}&=1,\\ L_{1,n}&=\theta,\\ L_{k+1,n}&=\theta L_{k,n}-t^{2}k\left(n+1-k\right)L_{k-1,n},\ 1\leq k\leq n.\end{split} (24)

Here we provide two more concrete results about the operator Ln+1L_{n+1}.

Put the degree on t,θ\mathbb{Q}\langle t,\theta\rangle as degt=degθ=1\deg t=\deg\theta=1. The associated graded ring grt,θ=[t¯,θ¯]\operatorname{gr}\mathbb{Q}\langle t,\theta\rangle=\mathbb{Q}[\overline{t},\overline{\theta}] is a polynomial ring where t¯\overline{t} and θ¯\overline{\theta} are the images of tt and θ\theta, respectively.

Proposition 35.

The image of Ln+1L_{n+1} in [t¯,θ¯]\mathbb{Q}[\overline{t},\overline{\theta}] is the polynomial

L¯n+1(t¯,θ¯)={i=1r(θ¯2(2i1)2t¯2)if n+1=2r is even,θ¯i=1r(θ¯2(2i)2t¯2)if n+1=2r+1 is odd.\overline{L}_{n+1}(\overline{t},\overline{\theta})=\begin{cases}\prod_{i=1}^{r}\left(\overline{\theta}^{2}-(2i-1)^{2}\overline{t}^{2}\right)&\text{if $n+1=2r$ is even},\\ \overline{\theta}\prod_{i=1}^{r}\left(\overline{\theta}^{2}-(2i)^{2}\overline{t}^{2}\right)&\text{if $n+1=2r+1$ is odd}.\end{cases} (25)
Proof.

Taking the images in [t¯,θ¯]\mathbb{Q}[\overline{t},\overline{\theta}] of the relation (24), we obtain L¯n+1=L¯n+1,n\overline{L}_{n+1}=\overline{L}_{n+1,n} satisfying

L¯0,n\displaystyle\overline{L}_{0,n} =1,L¯1,n=θ¯,\displaystyle=1,\quad\overline{L}_{1,n}=\overline{\theta},
L¯k+1,n\displaystyle\overline{L}_{k+1,n} =θ¯L¯k,nt2k(n+1k)L¯k1,n, 1kn.\displaystyle=\overline{\theta}\,\overline{L}_{k,n}-t^{2}k\left(n+1-k\right)\overline{L}_{k-1,n},\ 1\leq k\leq n.

The formula (25) is then a consequence of the following combinatorics lemma. ∎

Lemma 36.

For any mm\in\mathbb{N}, set the recurrence for λn,m(x),n{0}\lambda_{n,m}(x),n\in\mathbb{N}\cup\{0\},

λ0,m\displaystyle\lambda_{0,m} =1,λ1,m=x,\displaystyle=1,\quad\lambda_{1,m}=x,
λk+1,m\displaystyle\lambda_{k+1,m} =xλk,mk(m+1k)λk1,m,k1.\displaystyle=x\lambda_{k,m}-k\left(m+1-k\right)\lambda_{k-1,m},\quad k\geq 1.

Then we have

λm+1,m(x)={i=1r(x2(2i1)2),m+1=2r,xi=1r(x2(2i)2),m+1=2r+1.\displaystyle\lambda_{m+1,m}(x)=\begin{cases}\prod_{i=1}^{r}\left(x^{2}-(2i-1)^{2}\right),&m+1=2r,\\ x\prod_{i=1}^{r}\left(x^{2}-(2i)^{2}\right),&m+1=2r+1.\end{cases} (26)
Proof.

Notice that λi,m\lambda_{i,m} is a monic integral polynomial of degree ii for any mm. Consider the formal generating function222This generating function satisfies the differential equation y4f′′(y)(2y3my3)f(y)+(1xy+my2)f(y)=1.-y^{4}f^{\prime\prime}(y)-(2y^{3}-my^{3})f^{\prime}(y)+(1-xy+my^{2})f(y)=1.:

fm,x(y)=i=0λi,m(x)yi.f_{m,x}(y)=\sum_{i=0}^{\infty}\lambda_{i,m}(x)y^{i}.

An induction on ii immediately yields the relation fm,x1(y)+fm,x+1(y)=2fm1,x(y)f_{m,x-1}\left(y\right)+f_{m,x+1}\left(y\right)=2f_{m-1,x}\left(y\right) for any mm and xx.333Equality also holds when viewed as the solution of the corresponding differential equations. In other words, λi,m(x1)+λi,m(x+1)=2λi,m1(x)\lambda_{i,m}(x-1)+\lambda_{i,m}(x+1)=2\lambda_{i,m-1}(x) for all ii. Therefore we obtain

λm+1,m(x1)+λm+1,m(x+1)=2λm+1,m1(x)=2xλm,m1(x)\lambda_{m+1,m}(x-1)+\lambda_{m+1,m}(x+1)=2\lambda_{m+1,m-1}(x)=2x\lambda_{m,m-1}(x)

by the recurrence. Thus, since λm+1,m(x)\lambda_{m+1,m}(x) is a monic polynomial of degree m+1m+1, it is uniquely determined by the above functional equation when the polynomial λm,m1(x)\lambda_{m,m-1}(x) is given. Hence, by the induction, it suffices to show that

i=1r((x1)2(2i1)2)+i=1r((x+1)2(2i1)2)\displaystyle\prod_{i=1}^{r}\left((x-1)^{2}-(2i-1)^{2}\right)+\prod_{i=1}^{r}\left((x+1)^{2}-(2i-1)^{2}\right) =2x2i=1r1(x2(2i)2);\displaystyle=2x^{2}\prod_{i=1}^{r-1}\left(x^{2}-(2i)^{2}\right);
(x1)i=1r((x1)2(2i)2)+(x+1)i=1r((x+1)2(2i)2)\displaystyle(x-1)\prod_{i=1}^{r}\left((x-1)^{2}-(2i)^{2}\right)+(x+1)\prod_{i=1}^{r}\left((x+1)^{2}-(2i)^{2}\right) =2xi=1r(x2(2i1)2),\displaystyle=2x\prod_{i=1}^{r}\left(x^{2}-(2i-1)^{2}\right),

which are straightforward to verify. ∎

Proposition 37.

Write Ln+1L_{n+1} into the form itiPi(θ)\sum_{i}t^{i}P_{i}\left(\theta\right), where Pi(x)[x]P_{i}(x)\in\mathbb{Q}\left[x\right]. Define the integers a,ba,b by

a=max{iPi0};b=min{iPi0}.a=\max\left\{i\mid P_{i}\neq 0\right\};\quad b=\min\left\{i\mid P_{i}\neq 0\right\}.

Then we have a=2n+12a=2\left\lfloor\frac{n+1}{2}\right\rfloor and b=0b=0.

Proof.

By the recurrence (24), if we set degt=1\deg t=1 and degθ=0\deg\theta=0, we easily see that LjL_{j} has degree 2j22\left\lfloor\frac{j}{2}\right\rfloor by the interchanging relation θt=t+tθ\theta t=t+t\theta. Thus, we have a=2n+12a=2\left\lfloor\frac{n+1}{2}\right\rfloor. On the other hand, if we set degt=0\deg t=0 and degθ=1\deg\theta=1, we see that the leading term of Lk+1L_{k+1} is given by θk+1\theta^{k+1}. Therefore, we conclude that b=0b=0 by Proposition 35. ∎

A.2 Two-scale Bessel moments

From now on, we take for granted the properties of modified Bessel functions I0(t),K0(t)I_{0}(t),K_{0}(t) in the treatise [13].

Recall the Bessel moments IKMk(a,b){\mathrm{IKM}}_{k}(a,b) given in section 5.1. For r1r\in\mathbb{Z}_{\geq 1}, define the two r×rr\times r matrices

Mr=(IKM2r1(i1,2j2))1i,jr,Nr=(IKM2r(i1,2j2))1i,jr.M_{r}=\begin{pmatrix}{\mathrm{IKM}}_{2r-1}(i-1,2j-2)\end{pmatrix}_{1\leq i,j\leq r},\quad N_{r}=\begin{pmatrix}{\mathrm{IKM}}_{2r}(i-1,2j-2)\end{pmatrix}_{1\leq i,j\leq r}.

We aim to determine the two scalars detMr\det M_{r}, detNr\det N_{r} adapting the inductive methods explored by Zhou [9].

For the initial values, we have [13, §13.21, Eq.(8)]

M1=0K0(t)dt=π2,M_{1}=\int_{0}^{\infty}K_{0}(t)\,{\mathrm{d}}t=\frac{\pi}{2}, (27)

and, by [13, §13.72], one has

N1=0K02(t)dt\displaystyle N_{1}=\int_{0}^{\infty}K_{0}^{2}(t)\,{\mathrm{d}}t =120e2tcoshxcoshydxdydt\displaystyle=\frac{1}{2}\int_{0}^{\infty}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}e^{-2t\cosh x\cosh y}\,{\mathrm{d}}x{\mathrm{d}}y{\mathrm{d}}t
=14dxdycoshxcoshy\displaystyle=\frac{1}{4}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}\frac{{\mathrm{d}}x{\mathrm{d}}y}{\cosh x\cosh y}
=π24.\displaystyle=\frac{\pi^{2}}{4}.

For r0r\in\mathbb{Z}_{\geq 0}, let ω2r+1(x)\omega_{2r+1}(x) be the Wronskian of the (2r+1)(2r+1) functions fi(x)f_{i}(x)

fi(x)={0I0(xt)Ii10(t)K02ri+1(t)dt,1ir,0K0(xt)Iir10(t)K03ri+1(t)dt,r<i2r+1.f_{i}(x)=\begin{cases}\int_{0}^{\infty}I_{0}(xt)I^{i-1}_{0}(t)K_{0}^{2r-i+1}(t)\,{\mathrm{d}}t,&1\leq i\leq r,\\ \int_{0}^{\infty}K_{0}(xt)I^{i-r-1}_{0}(t)K_{0}^{3r-i+1}(t)\,{\mathrm{d}}t,&r<i\leq 2r+1.\end{cases} (28)

The functions fif_{i} are well-defined and analytic on the interval (0,2)(0,2) and hence so is ω2r+1\omega_{2r+1}. In particular, ω1(x)=0K0(xt)dt=π2x\omega_{1}(x)=\int_{0}^{\infty}K_{0}(xt)\,{\mathrm{d}}t=\frac{\pi}{2x} by (27).

For r1r\in\mathbb{Z}_{\geq 1}, let ω2r(x)\omega_{2r}(x) be the Wronskian of the 2r2r functions gi(x)g_{i}(x) where

gi(x)={0I0(xt)Ii10(t)K02ri(t)dt,1ir,0K0(xt)Iir10(t)K03ri(t)dt,r<i2r.g_{i}(x)=\begin{cases}\int_{0}^{\infty}I_{0}(xt)I^{i-1}_{0}(t)K_{0}^{2r-i}(t)\,{\mathrm{d}}t,&1\leq i\leq r,\\ \int_{0}^{\infty}K_{0}(xt)I^{i-r-1}_{0}(t)K_{0}^{3r-i}(t)\,{\mathrm{d}}t,&r<i\leq 2r.\end{cases}

All entries in the Wronskian matrix are well-defined analytic functions on the interval (0,1)(0,1) and so is ω2r(x)\omega_{2r}(x).

Proposition 38.

The determinant ωk(x)\omega_{k}(x) and its evaluation at x=1x=1 are given by the following formulae:

  1. 1.

    For r1r\in\mathbb{Z}_{\geq 1},

    ω2r+1(x)\displaystyle\omega_{2r+1}(x) =(1)r(r+1)22[1x2i=1r(2i)2(2i)2x2]2r+12Γ(r+12)2(detNr)2,\displaystyle=\frac{(-1)^{\frac{r(r+1)}{2}}}{2}\left[\frac{1}{x^{2}}\prod_{i=1}^{r}\frac{(2i)^{2}}{(2i)^{2}-x^{2}}\right]^{\frac{2r+1}{2}}\Gamma\Big{(}\frac{r+1}{2}\Big{)}^{2}\big{(}\det N_{r}\big{)}^{2},
    ω2r+1(1)\displaystyle\omega_{2r+1}(1) =(1)r(r+1)/2detMrdetMr+1.\displaystyle=(-1)^{r(r+1)/2}\det M_{r}\cdot\det M_{r+1}.
  2. 2.

    For r2r\in\mathbb{Z}_{\geq 2},

    ω2r(x)\displaystyle\omega_{2r}(x) =(1)r(r+1)2[1xi=1r(2i1)2(2i1)2x2]r(detMr)2,\displaystyle=(-1)^{\frac{r(r+1)}{2}}\left[\frac{1}{x}\prod_{i=1}^{r}\frac{(2i-1)^{2}}{(2i-1)^{2}-x^{2}}\right]^{r}\big{(}\det M_{r}\big{)}^{2},
    limx12r(1x)rω2r(x)\displaystyle\lim_{x\to 1^{-}}2^{r}(1-x)^{r}\omega_{2r}(x) =(1)r(r+1)/2(r1)!detNr1detNr.\displaystyle=(-1)^{r(r+1)/2}(r-1)!\det N_{r-1}\cdot\det N_{r}.

The above proposition leads to the recursive formulae

detMrdetMr+1\displaystyle\det M_{r}\cdot\det M_{r+1} =12[2rr!2r+1(2r+1)!!]2r+1Γ(r+12)2(detNr)2\displaystyle=\frac{1}{2}\left[\frac{2^{r}r!\sqrt{2r+1}}{(2r+1)!!}\right]^{2r+1}\Gamma\Big{(}\frac{r+1}{2}\Big{)}^{2}\big{(}\det N_{r}\big{)}^{2} (r1),\displaystyle(r\geq 1),
detNr1detNr\displaystyle\det N_{r-1}\cdot\det N_{r} =2r(r1)![(2r1)!!2r2rr!]2r(detMr)2\displaystyle=\frac{2^{r}}{(r-1)!}\left[\frac{(2r-1)!!\sqrt{2r}}{2^{r}r!}\right]^{2r}\big{(}\det M_{r}\big{)}^{2} (r2).\displaystyle(r\geq 2). (29)

With the initial data M1=π2,N1=π24M_{1}=\frac{\pi}{2},N_{1}=\frac{\pi^{2}}{4} and the relation

Γ(r2)Γ(r+12)=(r1)!2r1π,\Gamma\left(\frac{r}{2}\right)\Gamma\left(\frac{r+1}{2}\right)=\frac{\left(r-1\right)!}{2^{r-1}}\sqrt{\pi},

one immediately obtains the following results by induction.

Corollary 39.

For positive integers rr, we have

detMr\displaystyle\det M_{r} =πr(r+1)2r(r3)a=1r1ara2a+12a+1,\displaystyle=\sqrt{\pi}^{r(r+1)}\sqrt{2}^{r(r-3)}\prod_{a=1}^{r-1}\frac{a^{r-a}}{\sqrt{2a+1}^{2a+1}},
detNr\displaystyle\det N_{r} =1Γ(r+12)π(r+1)22r(r+3)a=1r1(2a+1)ra(a+1)a+1.\displaystyle=\frac{1}{\Gamma\left(\frac{r+1}{2}\right)}\frac{\sqrt{\pi}^{(r+1)^{2}}}{\sqrt{2}^{r(r+3)}}\prod_{a=1}^{r-1}\frac{(2a+1)^{r-a}}{(a+1)^{a+1}}.

In particular, the two scalars (2r1)!!πmrdetMr\sqrt{(2r-1)!!}\,\pi^{-m_{r}}\det M_{r} and πnrdetNr\pi^{-n_{r}}\det N_{r} are positive rational numbers, where mr=r(r+1)2m_{r}=\frac{r(r+1)}{2} and nr=(r+1)22n_{r}=\left\lfloor\frac{(r+1)^{2}}{2}\right\rfloor.

A.3 The Vanhove operators

The adjoint Ln+1L_{n+1}^{*} of Ln+1L_{n+1} is derived under the convolution (t,t)(t,t)(t,\partial_{t})\mapsto(t,-\partial_{t}) (so θ(θ+1)\theta\mapsto-(\theta+1)) and hence the leading term of the signed adjoint Λn+1=(1)n+1Ln+1\Lambda_{n+1}=(-1)^{n+1}L_{n+1}^{*} equals L¯n+1(θ¯,t¯)\overline{L}_{n+1}(\overline{\theta},\overline{t}) by Proposition 35. For F(xt)=I0(xt),K0(xt)F(xt)=I_{0}(xt),K_{0}(xt) and G(t)=I0a(t)K0na(t)G(t)=I_{0}^{a}(t)K_{0}^{n-a}(t), we have, by integration by parts,

0(Λn+1F(xt))G(t)dt=(1)n+10F(xt)(Ln+1G(t))dt=0.\int_{0}^{\infty}(\Lambda_{n+1}F(xt))G(t)\,{\mathrm{d}}t=(-1)^{n+1}\int_{0}^{\infty}F(xt)(L_{n+1}G(t))\,{\mathrm{d}}t=0.

The Vanhove operator Vn+1x,x±1V_{n+1}\in\mathbb{Q}\left\langle\partial_{x},x^{\pm 1}\right\rangle is of order (n+1)(n+1) such that Vn+1F(xt)=Λn+1F(xt)V_{n+1}F(xt)=\Lambda_{n+1}F(xt). So one has Vn+1fi=0V_{n+1}f_{i}=0 for fi(x)f_{i}(x) in (28) and consequently ωn+1(x)\omega_{n+1}(x) satisfies a first order linear differential equation (See (30) below).

Lemma 40.

Let λn+1(x)=L¯n+1(1,x1)[x1]\lambda_{n+1}(x)=\overline{L}_{n+1}(1,x^{-1})\in\mathbb{Q}[x^{-1}] of order 2n+122\left\lfloor\frac{n+1}{2}\right\rfloor with respect to x1x^{-1}. Let θx=xx\theta_{x}=x\partial_{x}. One has

Vn+1\displaystyle V_{n+1} =λn+1(x)θxn+1+(n+1)[λn+1(x)+xλn+1(x)2]θxn+δ1\displaystyle=\lambda_{n+1}(x)\theta_{x}^{n+1}+(n+1)\left[\lambda_{n+1}(x)+\frac{x\lambda_{n+1}^{\prime}(x)}{2}\right]\theta_{x}^{n}+\delta_{1}
=xn+1λn+1(x)xn+1+n+12xn[(n+2)λn+1(x)+xλn+1(x)]xn+δ2\displaystyle=x^{n+1}\lambda_{n+1}(x)\partial_{x}^{n+1}+\frac{n+1}{2}x^{n}\left[(n+2)\lambda_{n+1}(x)+x\lambda_{n+1}^{\prime}(x)\right]\partial_{x}^{n}+\delta_{2}

where δ1,δ2\delta_{1},\delta_{2} are of order at most (n1)(n-1) with respect to x\partial_{x} in x,x±1\mathbb{Q}\langle\partial_{x},x^{\pm 1}\rangle.

Proof.

By Vanhove [18], there exists L~n1x,x±1\widetilde{L}_{n-1}\in\mathbb{Q}\langle\partial_{x},x^{\pm 1}\rangle of order (n1)(n-1) such that

tL~n1F(xt)=Λn+1F(xt)t.t\widetilde{L}_{n-1}F(xt)=\Lambda_{n+1}\frac{F(xt)}{t}.

The operator L~n1\widetilde{L}_{n-1} is of the form ([9, Eq. (4.29)])

L~n1=x2λ(x)θxn1+x2[2(n1)λ(x)+n12xλ(x)]θxn2+δ~\widetilde{L}_{n-1}=x^{2}\lambda(x)\theta_{x}^{n-1}+x^{2}\left[2(n-1)\lambda(x)+\frac{n-1}{2}x\lambda^{\prime}(x)\right]\theta_{x}^{n-2}+\widetilde{\delta}

where δ~\widetilde{\delta} is of order at most (n3)(n-3) with respect to x\partial_{x} in x,x±1\mathbb{Q}\langle\partial_{x},x^{\pm 1}\rangle444Comparing L~n1(θx)\widetilde{L}_{n-1}(\theta_{x}) with Zhou’s Vanhove operator L~n1(θu)\widetilde{L}_{n-1}(\theta_{u}), we set his variable u=x2u=x^{2} and multiply L~n1(θu)\widetilde{L}_{n-1}(\theta_{u}) by 2n12^{n-1}..

Set

Δn(θt)=Λn+1(θt)Λn+1(θt1).\Delta_{n}(\theta_{t})=\Lambda_{n+1}(\theta_{t})-\Lambda_{n+1}(\theta_{t}-1).

Since θt1t=1t(θt1)\theta_{t}\frac{1}{t}=\frac{1}{t}(\theta_{t}-1) in t,t\mathbb{Q}\left\langle\partial_{t},t\right\rangle, we have

Λn+1(θt)F(xt)\displaystyle\Lambda_{n+1}(\theta_{t})F(xt) =tΛn+1(θt)F(xt)t+Δn(θt)F(xt)\displaystyle=t\Lambda_{n+1}(\theta_{t})\frac{F(xt)}{t}+\Delta_{n}(\theta_{t})F(xt)
=[t2L~n1(θx)+Δn(θt)]F(xt).\displaystyle=\Big{[}t^{2}\widetilde{L}_{n-1}(\theta_{x})+\Delta_{n}(\theta_{t})\Big{]}F(xt).

Since t2F(xt)=1x2θx2F(xt)t^{2}F(xt)=\frac{1}{x^{2}}\theta_{x}^{2}F(xt), we have

t2L~n1(θx)F(xt)\displaystyle t^{2}\widetilde{L}_{n-1}(\theta_{x})F(xt) =L~n1(θx)1x2θx2F(xt)\displaystyle=\widetilde{L}_{n-1}(\theta_{x})\frac{1}{x^{2}}\theta_{x}^{2}F(xt)
=1x2L~n1(θx2)θx2F(xt)\displaystyle=\frac{1}{x^{2}}\widetilde{L}_{n-1}(\theta_{x}-2)\theta_{x}^{2}F(xt)

and the differential operator reads

λ(x)θxn+1+n12xλ(x)θxn+δ3\lambda(x)\theta_{x}^{n+1}+\frac{n-1}{2}x\lambda^{\prime}(x)\theta_{x}^{n}+\delta_{3}

where δ3\delta_{3} is of order at most (n1)(n-1) with respect to x\partial_{x} in x,x±1\mathbb{Q}\langle\partial_{x},x^{\pm 1}\rangle.

On the other hand, since θtF(xt)=θxF(xt)\theta_{t}F(xt)=\theta_{x}F(xt) and by Proposition 35, we have

Δn(θt)F(xt)=[Λn+1(θ)Λn+1(θ1)]F(xt)=[((n+1)λ(x)+xλ(x))θxn+δ4]F(xt)\Delta_{n}(\theta_{t})F(xt)=\left[\Lambda_{n+1}\left(\theta\right)-\Lambda_{n+1}\left(\theta-1\right)\right]F\left(xt\right)=\left[\left((n+1)\lambda(x)+x\lambda^{\prime}(x)\right)\theta_{x}^{n}+\delta_{4}\right]F(xt)

where δ4\delta_{4} is of order at most (n1)(n-1) with respect to x\partial_{x} in x,x1\mathbb{Q}\langle\partial_{x},x^{-1}\rangle. Therefore the leading two terms of Vn+1V_{n+1} are determined. ∎

Rationality of ωn+1(x)\omega_{n+1}(x)

Lemma 40 yields

ωn+1(x)=n+12x[(n+2)+xλn+1(x)λn+1(x)]ωn+1(x).\omega_{n+1}^{\prime}(x)=-\frac{n+1}{2x}\left[(n+2)+\frac{x\lambda_{n+1}^{\prime}(x)}{\lambda_{n+1}(x)}\right]\omega_{n+1}(x). (30)

Since ωn+1(x)\omega_{n+1}(x) takes real values on (0,1)(0,1), one obtains

ωn+1(x)=Cn+1[(1)n+12xn+2λn+1(x)]n+12\omega_{n+1}(x)=C_{n+1}\left[(-1)^{\left\lfloor\frac{n+1}{2}\right\rfloor}x^{n+2}\lambda_{n+1}(x)\right]^{-\frac{n+1}{2}}

for some real constant Cn+1C_{n+1} for each n0n\in\mathbb{Z}_{\geq 0}. We shall determine Cn+1C_{n+1} by investigating the limiting behavior of ωn+1(x)\omega_{n+1}(x) as x0+x\to 0^{+}.

A.4 Singularities of ωn+1(x)\omega_{n+1}(x)

For F(xt)=I0(xt)F(xt)=I_{0}(xt) or K0(xt)K_{0}(xt), we have

xF(xt)=tF(xt),x2F(xt)=txF(xt)+t2F(xt).\partial_{x}F(xt)=tF^{\prime}(xt),\quad\partial_{x}^{2}F(xt)=-\frac{t}{x}F^{\prime}(xt)+t^{2}F(xt).

So ωn+1(x)\omega_{n+1}(x) coincides with the determinant of the matrix Ωn+1(x)\Omega_{n+1}(x) of size (n+1)(n+1) whose (i,j)(i,j)-entry is

{0I0(xt)I0j1(t)K0nj+1(t)ti1dt,1jn+12,i=1,3,,2n2+1,0tI0(xt)I0j1(t)K0nj+1(t)ti2dt,1jn+12,i=2,4,,2n+12,0K0(xt)I0jr1(t)K0nj+r+1(t)ti1dt,n+12<jn+1,i=1,3,,2n2+1,0tK0(xt)I0jr1(t)K0nj+r+1(t)ti2dt,n+12<jn+1,i=2,4,,2n+12.\begin{cases}\int_{0}^{\infty}I_{0}(xt)I_{0}^{j-1}(t)K_{0}^{n-j+1}(t)t^{i-1}\,{\mathrm{d}}t,&1\leq j\leq\left\lfloor\frac{n+1}{2}\right\rfloor,i=1,3,\cdots,2\left\lfloor\frac{n}{2}\right\rfloor+1,\\ \int_{0}^{\infty}tI_{0}^{\prime}(xt)I_{0}^{j-1}(t)K_{0}^{n-j+1}(t)t^{i-2}\,{\mathrm{d}}t,&1\leq j\leq\left\lfloor\frac{n+1}{2}\right\rfloor,i=2,4,\cdots,2\left\lfloor\frac{n+1}{2}\right\rfloor,\\ \int_{0}^{\infty}K_{0}(xt)I_{0}^{j-r-1}(t)K_{0}^{n-j+r+1}(t)t^{i-1}\,{\mathrm{d}}t,&\left\lfloor\frac{n+1}{2}\right\rfloor<j\leq n+1,i=1,3,\cdots,2\left\lfloor\frac{n}{2}\right\rfloor+1,\\ \int_{0}^{\infty}tK_{0}^{\prime}(xt)I_{0}^{j-r-1}(t)K_{0}^{n-j+r+1}(t)t^{i-2}\,{\mathrm{d}}t,&\left\lfloor\frac{n+1}{2}\right\rfloor<j\leq n+1,i=2,4,\cdots,2\left\lfloor\frac{n+1}{2}\right\rfloor.\end{cases}

Properties of I0(t)I_{0}(t) and K0(t)K_{0}(t)

We collect some properties of the modified Bessel functions I0(t)I_{0}(t) and K0(t)K_{0}(t) in order to obtain information of ωn+1(x)\omega_{n+1}(x) as x0+,1x\to 0^{+},1^{-}.

The function I0(t)I_{0}(t) is entire and even; it is real and increasing on the half line [1,)[1,\infty). The function K0(t)K_{0}(t) has a logarithmic pole at x=0x=0; it is real and decreasing on (0,)(0,\infty). On the half plane (t)>0\Re(t)>0, we have the asymptotic approximations

I0(t)=et2πt(1+O(1t)),K0(t)=π2tet(1+O(1t))I_{0}(t)=\frac{e^{t}}{\sqrt{2\pi t}}\left(1+O\Big{(}\frac{1}{t}\Big{)}\right),\quad K_{0}(t)=\sqrt{\frac{\pi}{2t}}e^{-t}\left(1+O\Big{(}\frac{1}{t}\Big{)}\right)

as tt\to\infty. In particular, for a positive integer aa,

[I0(t)K0(t)]a1(2t)a=O(1ta+1)[I_{0}(t)K_{0}(t)]^{a}-\frac{1}{(2t)^{a}}=O\Big{(}\frac{1}{t^{a+1}}\Big{)}

as tt\to\infty along the real line. One has the boundedness

supt>0|tK0(t)+1|t(1+|logt|)<.\sup_{t>0}\frac{|tK^{\prime}_{0}(t)+1|}{t(1+|\log t|)}<\infty.

For c0c\in\mathbb{Z}_{\geq 0}, one has the evaluation [13, §13.21, Eq.(8)]

0K0(t)tcdt=2c1Γ(c+12)2.\int_{0}^{\infty}K_{0}(t)t^{c}\,{\mathrm{d}}t=2^{c-1}\Gamma\Big{(}\frac{c+1}{2}\Big{)}^{2}.

Integrations

With the data collected above, we list some consequences for the integrals that appear in the entries of the matrix Ωn+1(x)\Omega_{n+1}(x).

For 0a<b0\leq a<b and c0c\geq 0, one obtains

0K0(xt)I0a(t)K0b(t)tcdt\displaystyle\int_{0}^{\infty}K_{0}(xt)I_{0}^{a}(t)K_{0}^{b}(t)t^{c}\,{\mathrm{d}}t =O(logx),\displaystyle=O(\log x), (31)
0I0(xt)I0a(t)K0b(t)tcdt\displaystyle\int_{0}^{\infty}I_{0}^{\prime}(xt)I_{0}^{a}(t)K_{0}^{b}(t)t^{c}\,{\mathrm{d}}t =O(x)\displaystyle=O(x) (32)

and

0tK0(xt)I0a(t)K0b(t)tcdt\displaystyle\int_{0}^{\infty}tK^{\prime}_{0}(xt)I_{0}^{a}(t)K_{0}^{b}(t)t^{c}\,{\mathrm{d}}t =1x[0I0a(t)K0b(t)tcdt0(xtK0(xt)+1)I0a(t)K0b(t)tcdt]\displaystyle=\frac{-1}{x}\left[\int_{0}^{\infty}I_{0}^{a}(t)K_{0}^{b}(t)t^{c}\,{\mathrm{d}}t-\int_{0}^{\infty}\big{(}xtK_{0}^{\prime}(xt)+1\big{)}I_{0}^{a}(t)K_{0}^{b}(t)t^{c}\,{\mathrm{d}}t\right]
=1x0I0a(t)K0b(t)tcdt+O(logx)\displaystyle=\frac{-1}{x}\int_{0}^{\infty}I_{0}^{a}(t)K_{0}^{b}(t)t^{c}\,{\mathrm{d}}t+O(\log x) (33)

as x0+x\to 0^{+}. For 0ca0\leq c\leq a and as x0+x\to 0^{+}, we thus have

0K0(xt)I0a(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}K_{0}(xt)I_{0}^{a}(t)K_{0}^{a}(t)t^{c}\,{\mathrm{d}}t =O(0K0(xt)dt)=O(1x),\displaystyle=O\Big{(}\int_{0}^{\infty}K_{0}(xt)\,{\mathrm{d}}t\Big{)}=O\Big{(}\frac{1}{x}\Big{)}, (34)
0tK0(xt)I0a(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}tK_{0}^{\prime}(xt)I_{0}^{a}(t)K_{0}^{a}(t)t^{c}\,{\mathrm{d}}t =O(0tK0(xt)dt)=O(1x2).\displaystyle=O\Big{(}\int_{0}^{\infty}tK_{0}^{\prime}(xt)\,{\mathrm{d}}t\Big{)}=O\Big{(}\frac{1}{x^{2}}\Big{)}. (35)

If 0a<c0\leq a<c and as x0+x\to 0^{+}, then

0K0(xt)I0a(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}K_{0}(xt)I_{0}^{a}(t)K_{0}^{a}(t)t^{c}{\mathrm{d}}t =0K0(xt)tca2adt+0K0(xt)[I0a(t)K0a(t)1(2t)a]tcdt\displaystyle=\int_{0}^{\infty}\frac{K_{0}(xt)t^{c-a}}{2^{a}}{\mathrm{d}}t+\int_{0}^{\infty}K_{0}(xt)\Big{[}I_{0}^{a}(t)K_{0}^{a}(t)-\frac{1}{(2t)^{a}}\Big{]}t^{c}{\mathrm{d}}t
=2c2a1xca+1Γ(ca+12)2+O(1xca),\displaystyle=\frac{2^{c-2a-1}}{x^{c-a+1}}\Gamma\Big{(}\frac{c-a+1}{2}\Big{)}^{2}+O\Big{(}\frac{1}{x^{c-a}}\Big{)}, (36)
0tK0(xt)I0a(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}tK_{0}^{\prime}(xt)I_{0}^{a}(t)K_{0}^{a}(t)t^{c}{\mathrm{d}}t =0tK0(xt)tca2adt+0tK0(xt)[I0a(t)K0a(t)1(2t)a]tcdt\displaystyle=\int_{0}^{\infty}\frac{tK_{0}^{\prime}(xt)t^{c-a}}{2^{a}}{\mathrm{d}}t+\int_{0}^{\infty}tK_{0}^{\prime}(xt)\Big{[}I_{0}^{a}(t)K_{0}^{a}(t)-\frac{1}{(2t)^{a}}\Big{]}t^{c}{\mathrm{d}}t
=ca+12ax0K0(xt)tcadt+O(1x2)\displaystyle=\frac{c-a+1}{2^{a}x}\int_{0}^{\infty}K_{0}(xt)t^{c-a}\,{\mathrm{d}}t+O\Big{(}\frac{1}{x^{2}}\Big{)}
=O(1xca+2).\displaystyle=O\Big{(}\frac{1}{x^{c-a+2}}\Big{)}. (37)

On the real line, we have [9, Lemma 4.5]

limx10I0(xt)K0(t)dtlog(1x)=12\lim_{x\to 1^{-}}\frac{\int_{0}^{\infty}I_{0}(xt)K_{0}(t)\,{\mathrm{d}}t}{-\log(1-x)}=\frac{1}{2}

and for c0c\in\mathbb{Z}_{\geq 0},

limx1(1x)c+10I0(xt)K0(t)tc+1dt=c!2=limx1(1x)c+10tI0(xt)K0(t)tcdt.\lim_{x\to 1^{-}}(1-x)^{c+1}\int_{0}^{\infty}I_{0}(xt)K_{0}(t)t^{c+1}\,{\mathrm{d}}t=\frac{c!}{2}=\lim_{x\to 1^{-}}(1-x)^{c+1}\int_{0}^{\infty}tI_{0}^{\prime}(xt)K_{0}(t)t^{c}\,{\mathrm{d}}t.

Therefore for a1,a>ca\geq 1,a>c and x1x\to 1^{-}, one has

0I0(xt)I0a1(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}I_{0}(xt)I_{0}^{a-1}(t)K_{0}^{a}(t)t^{c}\,{\mathrm{d}}t =0I0(xt)K0(t)[I0a1(t)K0a1(t)tc]dt\displaystyle=\int_{0}^{\infty}I_{0}(xt)K_{0}(t)\big{[}I_{0}^{a-1}(t)K_{0}^{a-1}(t)t^{c}\big{]}\,{\mathrm{d}}t
=O(0I0(xt)K0(t)dt)\displaystyle=O\Big{(}\int_{0}^{\infty}I_{0}(xt)K_{0}(t)\,{\mathrm{d}}t\Big{)}
=O(log(1x)),\displaystyle=O\big{(}\log(1-x)\big{)}, (38)
0tI0(xt)I0a1(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}tI_{0}^{\prime}(xt)I_{0}^{a-1}(t)K_{0}^{a}(t)t^{c}\,{\mathrm{d}}t =0tI0(xt)K0(t)[I0a1(t)K0a1(t)tc]dt\displaystyle=\int_{0}^{\infty}tI_{0}^{\prime}(xt)K_{0}(t)\big{[}I_{0}^{a-1}(t)K_{0}^{a-1}(t)t^{c}\big{]}\,{\mathrm{d}}t
=O(0tI0(xt)K0(t)dt)\displaystyle=O\Big{(}\int_{0}^{\infty}tI^{\prime}_{0}(xt)K_{0}(t)\,{\mathrm{d}}t\Big{)}
=O(11x).\displaystyle=O\Big{(}\frac{1}{1-x}\Big{)}. (39)

If ca1c\geq a\geq 1 and x1x\to 1^{-}, then

0I0(xt)I0a1(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}I_{0}(xt)I_{0}^{a-1}(t)K_{0}^{a}(t)t^{c}\,{\mathrm{d}}t =0I0(xt)K0(t)tca+12a1dt+0I0(xt)K0(t)[I0a1(t)K0a1(t)1(2t)a1]tcdt\displaystyle\begin{aligned} =&\int_{0}^{\infty}\frac{I_{0}(xt)K_{0}(t)t^{c-a+1}}{2^{a-1}}\,{\mathrm{d}}t\\ &\hskip 20.0pt+\int_{0}^{\infty}I_{0}(xt)K_{0}(t)\Big{[}I_{0}^{a-1}(t)K_{0}^{a-1}(t)-\frac{1}{(2t)^{a-1}}\Big{]}t^{c}\,{\mathrm{d}}t\end{aligned}
=(ca)!2a(1x)ca+1+o(1(1x)ca+1),\displaystyle=\frac{(c-a)!}{2^{a}(1-x)^{c-a+1}}+o\Big{(}\frac{1}{(1-x)^{c-a+1}}\Big{)}, (40)
0tI0(xt)I0a1(t)K0a(t)tcdt\displaystyle\int_{0}^{\infty}tI^{\prime}_{0}(xt)I_{0}^{a-1}(t)K_{0}^{a}(t)t^{c}\,{\mathrm{d}}t =0tI0(xt)K0(t)tca+12a1dt+0tI0(xt)K0(t)[I0a1(t)K0a1(t)1(2t)a1]tcdt\displaystyle\begin{aligned} =&\int_{0}^{\infty}\frac{tI^{\prime}_{0}(xt)K_{0}(t)t^{c-a+1}}{2^{a-1}}\,{\mathrm{d}}t\\ &\hskip 20.0pt+\int_{0}^{\infty}tI^{\prime}_{0}(xt)K_{0}(t)\Big{[}I_{0}^{a-1}(t)K_{0}^{a-1}(t)-\frac{1}{(2t)^{a-1}}\Big{]}t^{c}\,{\mathrm{d}}t\end{aligned}
=(ca+1)!2a(1x)ca+2+o(1(1x)ca+2).\displaystyle=\frac{(c-a+1)!}{2^{a}(1-x)^{c-a+2}}+o\Big{(}\frac{1}{(1-x)^{c-a+2}}\Big{)}. (41)

Notice that the error terms in the above two formulas are of class small oo; it is needed in the investigation of the limit of ω2r\omega_{2r} as x1x\to 1^{-} below.

Evaluation of ω2r+1(x)\omega_{2r+1}(x) at x=1x=1

All entries of Ω2r+1(x)\Omega_{2r+1}(x) can be evaluated at x=1x=1. We move the 2i2i-th row to row ii in Ω2r+1(1)\Omega_{2r+1}(1) for 1ir1\leq i\leq r and then subtract the (r+j+1)(r+j+1)-st column from the jj-th column of the resulting matrix for 1jr1\leq j\leq r. By (3) on the upper-left block, we obtain

ω2r+1(1)=(1)r(r+1)2det(Mr0Mr+1).\omega_{2r+1}(1)=(-1)^{\frac{r(r+1)}{2}}\det\begin{pmatrix}M_{r}&*\\ 0&M_{r+1}\end{pmatrix}.

Behavior of ω2r+1(x)\omega_{2r+1}(x) as x0+x\to 0^{+}

Fix r1r\geq 1. We move row (2i1)(2i-1) of Ω2r+1(x)\Omega_{2r+1}(x) to row ii for 1ir1\leq i\leq r, which creates a sign (1)r(r1)/2(-1)^{r(r-1)/2} to the determinant ω2r+1(x)\omega_{2r+1}(x). As x0+x\to 0^{+}, the resulting matrix decomposes into (r,r,1)×(r,r,1)(r,r,1)\times(r,r,1) blocks of the form

(Nr+o(1)O(logx)O(1xr)O(x)1xNr+O(logx)O(1xr)O(1)O(logx)12xr+1Γ(r+12)2+O(1xr))\begin{pmatrix}N_{r}+o(1)&O(\log x)&O\big{(}\frac{1}{x^{r}}\big{)}\\ O(x)&\frac{-1}{x}N_{r}+O(\log x)&O\big{(}\frac{1}{x^{r}}\big{)}\\ O(1)&O(\log x)&\frac{1}{2x^{r+1}}\Gamma\big{(}\frac{r+1}{2}\big{)}^{2}+O\big{(}\frac{1}{x^{r}}\big{)}\end{pmatrix}

by direct evaluation and (32) in the left three blocks, (31) and (33) in the middle, and (34), (35), (36) and (37) in the last column. The leading term of ω2r+1(x)\omega_{2r+1}(x), which is of order x(2r+1)x^{-(2r+1)}, comes from the diagonal blocks and one gets

limx0+x2r+1ω2r+1(x)=(1)r(r+1)212Γ(r+12)2det2Nr.\lim_{x\to 0^{+}}x^{2r+1}\omega_{2r+1}(x)=(-1)^{\frac{r(r+1)}{2}}\frac{1}{2}\Gamma\Big{(}\frac{r+1}{2}\Big{)}^{2}{\det}^{2}N_{r}.

Behavior of ω2r(x)\omega_{2r}(x) as x1x\to 1^{-}

Fix r2r\geq 2. We move 2i2i-th row of Ω2r(x)\Omega_{2r}(x) to row ii for i=1,2,,(r1)i=1,2,\cdots,(r-1) and rr-th column to the last, which adds a sign (1)r(r+1)/2(-1)^{r(r+1)/2} to the determinant ω2r(x)\omega_{2r}(x). We subtract jj-th column by (r+j)(r+j)-th for j=1,2,,(r1)j=1,2,\cdots,(r-1). As x1x\to 1^{-}, the resulting matrix decomposes into (r1,r,1)×(r1,r,1)(r-1,r,1)\times(r-1,r,1) blocks of the form

(Nr1+o(1)O(1)O(1(1x)r1)0Nr+o(1)O(1(1x)r1)O(1)O(1)(r1)!2r(1x)r+o(1(1x)r))\begin{pmatrix}N_{r-1}+o(1)&O(1)&O\big{(}\frac{1}{(1-x)^{r-1}}\big{)}\\ 0&N_{r}+o(1)&O\big{(}\frac{1}{(1-x)^{r-1}}\big{)}\\ O(1)&O(1)&\frac{(r-1)!}{2^{r}(1-x)^{r}}+o\big{(}\frac{1}{(1-x)^{r}}\big{)}\end{pmatrix}

by (3) and direct evaluation in the left three blocks, direct evaluation in the middle, and (38), (39), (40) and (41) in the last column. The leading term of ω2r(x)\omega_{2r}(x), which is of order (1x)r(1-x)^{-r}, comes from the diagonal blocks. It yields

limx1(1x)rω2r(x)=(1)r(r+1)2(r1)!2rdetNr1detNr.\lim_{x\to 1^{-}}(1-x)^{r}\omega_{2r}(x)=(-1)^{\frac{r(r+1)}{2}}\frac{(r-1)!}{2^{r}}\det N_{r-1}\det N_{r}.

Behavior of ω2r(x)\omega_{2r}(x) as x0+x\to 0^{+}

Fix r2r\geq 2. We move row (2i1)(2i-1) of Ω2r(x)\Omega_{2r}(x) to row ii for 1ir1\leq i\leq r, which adds a sign (1)r(r1)/2(-1)^{r(r-1)/2} to the determinant ω2r(x)\omega_{2r}(x). As x0+x\to 0^{+}, the resulting matrix decomposes into four blocks of equal size of the form

(Mr+o(1)O(logx)O(x)1xMr+O(logx))\begin{pmatrix}M_{r}+o(1)&O(\log x)\\ O(x)&\frac{-1}{x}M_{r}+O(\log x)\end{pmatrix}

by direct evaluation and (32) in the left two blocks and (31) and (33) in the right. This leads to

limx0+xrω2r(x)=(1)r(r+1)2det2Mr.\lim_{x\to 0^{+}}x^{r}\omega_{2r}(x)=(-1)^{\frac{r(r+1)}{2}}{\det}^{2}M_{r}.
Remark 41.

Proposition 38 indeed holds for ω2(x)\omega_{2}(x) by the same analysis if we set detN0=1\det N_{0}=1; it is consistent with the relation (29) for r=1r=1.

References

  • \bibcommenthead
  • Fresán et al. [2022] Fresán, J., Sabbah, C., Yu, J.-D.: Hodge theory of Kloosterman connections. Duke Math. J. 171(8), 1649–1747 (2022) https://doi.org/10.1215/00127094-2021-0036
  • Fresán et al. [2023] Fresán, J., Sabbah, C., Yu, J.-D.: Quadratic relations between Bessel moments. Algebra Number Theory 17(3), 541–602 (2023) https://doi.org/10.2140/ant.2023.17.541
  • Broadhurst [2016] Broadhurst, D.: Feynman integrals, L-series and Kloosterman moments. Commun. Number Theory Phys. 10(3), 527–569 (2016) https://doi.org/10.4310/CNTP.2016.v10.n3.a3
  • Bailey et al. [2008] Bailey, D.H., Borwein, J.M., Broadhurst, D., Glasser, M.L.: Elliptic integral evaluations of Bessel moments and applications. J. Phys. A 41(20), 205203–46 (2008) https://doi.org/10.1088/1751-8113/41/20/205203
  • Broadhurst and Mellit [2016] Broadhurst, D., Mellit, A.: Perturbative quantum field theory informs algebraic geometry. PoS LL2016, 079 (2016) https://doi.org/10.22323/1.260.0079
  • Zhou [2019] Zhou, Y.: Hilbert transforms and sum rules of Bessel moments. Ramanujan J. 48(1), 159–172 (2019) https://doi.org/10.1007/s11139-017-9945-y
  • Zhou [2022] Zhou, Y.: \mathbb{Q}-linear dependence of certain Bessel moments. Ramanujan J. 58(3), 723–746 (2022) https://doi.org/10.1007/s11139-021-00416-9
  • Fresán et al. [2023] Fresán, J., Sabbah, C., Yu, J.-D.: Quadratic relations between periods of connections. Tohoku Math. J. (2) 75(2), 175–213 (2023) https://doi.org/10.2748/tmj.20211209
  • Zhou [2018] Zhou, Y.: Wrońskian factorizations and Broadhurst-Mellit determinant formulae. Commun. Number Theory Phys. 12(2), 355–407 (2018) https://doi.org/10.4310/cntp.2018.v12.n2.a5
  • Bloch and Esnault [2004] Bloch, S., Esnault, H.: Homology for irregular connections. J. Théor. Nombres Bordeaux 16(2), 357–371 (2004)
  • Katz [1990] Katz, N.M.: Exponential Sums and Differential Equations. Annals of Mathematics Studies, vol. 124, p. 430. Princeton University Press, Princeton, NJ (1990). https://doi.org/10.1515/9781400882434
  • [12] NIST Digital Library of Mathematical Functions. https://dlmf.nist.gov/, Release 1.2.1 of 2024-06-15. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds. https://dlmf.nist.gov/
  • Watson [1995] Watson, G.N.: A Treatise on the Theory of Bessel Functions. Cambridge Mathematical Library, p. 804. Cambridge University Press, Cambridge (1995). Reprint of the second (1944) edition
  • Kim [2005] Kim, T.: A note on the alternating sums of powers of consecutive integers. arXiv:math/0508233 (2005)
  • Han [2020] Han, G.-N.: Hankel continued fractions and Hankel determinants of the Euler numbers. Trans. Amer. Math. Soc. 373(6), 4255–4283 (2020) https://doi.org/10.1090/tran/8031
  • Borwein and Salvy [2008] Borwein, J.M., Salvy, B.: A proof of a recurrence for Bessel moments. Experiment. Math. 17(2), 223–230 (2008)
  • Bronstein et al. [1997] Bronstein, M., Mulders, T., Weil, J.-A.: On symmetric powers of differential operators. In: Proceedings of the 1997 International Symposium on Symbolic and Algebraic Computation (Kihei, HI), pp. 156–163. ACM, New York (1997). https://doi.org/10.1145/258726.258771
  • Vanhove [2014] Vanhove, P.: The physics and the mixed Hodge structure of Feynman integrals. In: String-Math 2013. Proc. Sympos. Pure Math., vol. 88, pp. 161–194. Amer. Math. Soc., Providence, RI (2014). https://doi.org/10.1090/pspum/088/01455