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

A qq-analogue of symmetric multiple zeta value

Yoshihiro Takeyama Department of Mathematics, Institute of Pure and Applied Sciences, University of Tsukuba, Tsukuba, Ibaraki 305-8571, Japan [email protected]
Abstract.

We construct a qq-analogue of truncated version of symmetric multiple zeta values which satisfies the double shuffle relation. Using it, we define a qq-analogue of symmetric multiple zeta values and see that it satisfies many of the same relations as symmetric multiple zeta values, which are the inverse relation and a part of the double shuffle relation and the Ohno-type relation.

This work was supported by JSPS KAKENHI Grant Number 22K03243.
 This is a pre-print of an article published in The Ramanujan Journal. The final authenticated version is available online at: https://doi.org/10.1007/s11139-023-00755-9.

1. Introduction

The multiple zeta value (MZV) is the real value defined by

(1.1) ζ(𝐤)=0<m1<<mr1m1k1mrkr\displaystyle\zeta(\mathbf{k})=\sum_{0<m_{1}<\cdots<m_{r}}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}}

for a tuple of positive integers 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) with kr2k_{r}\geq 2. We set ζ()=1\zeta(\varnothing)=1 and regard it as a MZV. We denote by 𝒵\mathcal{Z} the \mathbb{Q}-linear subspace of \mathbb{R} spanned by the MZVs. It is known that 𝒵\mathcal{Z} forms a \mathbb{Q}-algebra with respect to the usual multiplication on \mathbb{R}.

Kaneko and Zagier introduced two kinds of variants of MZVs called finite multiple zeta values and symmetric multiple zeta values (see, e.g., [5]). The symmetric multiple zeta value (SMZV) is defined as an element of the quotient 𝒵/ζ(2)𝒵\mathcal{Z}/\zeta(2)\mathcal{Z}. The purpose of this paper is to construct a qq-analogue of the SMZVs which shares many of the relations among them.

The SMZV is defined as follows. Although the infinite sum (1.1) diverges if kr=1k_{r}=1, we have two kinds of its regularization, which are called the harmonic regularized MZV ζ(𝐤)\zeta^{\ast}(\mathbf{k}) and the shuffle regularized MZV ζsh(𝐤)\zeta^{\mathcyr{sh}}(\mathbf{k}) (see Section 2.2 for the details). If kr2k_{r}\geq 2, they are equal to the MZV ζ(𝐤)\zeta(\mathbf{k}). Using them we set

ζ𝒮,(k1,,kr)=i=0r(1)ki+1++krζ(k1,k2,,ki)ζ(kr,kr1,,ki+1)\displaystyle\zeta^{\mathcal{S},\bullet}(k_{1},\ldots,k_{r})=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\zeta^{\bullet}(k_{1},k_{2},\ldots,k_{i})\zeta^{\bullet}(k_{r},k_{r-1},\ldots,k_{i+1})

for {,sh}\bullet\in\{\ast,\mathcyr{sh}\}. It is known that the difference ζ(𝐤)ζsh(𝐤)\zeta^{\ast}(\mathbf{k})-\zeta^{\mathcyr{sh}}(\mathbf{k}) belongs to the ideal ζ(2)𝒵\zeta(2)\mathcal{Z} for any tuple 𝐤\mathbf{k} of positive integers. The SMZV ζ𝒮(𝐤)\zeta^{\mathcal{S}}(\mathbf{k}) is defined by

ζ𝒮(𝐤)=ζ𝒮,(𝐤)modζ(2)𝒵\displaystyle\zeta^{\mathcal{S}}(\mathbf{k})=\zeta^{\mathcal{S},\bullet}(\mathbf{k})\qquad\hbox{mod}\,\,\zeta(2)\mathcal{Z}

as an element of the quotient 𝒵/ζ(2)𝒵\mathcal{Z}/\zeta(2)\mathcal{Z}.

Now we fix a complex parameter qq satisfying 0<|q|<10<|q|<1 and define the qq-integer [n][n] for n1n\geq 1 by [n]=(1qn)/(1q)[n]=(1-q^{n})/(1-q). There are various models of a qq-analogue of the MZV (see, e.g., [11, Chapter 12]). Most of them are of the following form:

(1.2) 0<m1<<mrP1(qm1)Pr(qmr)[m1]k1[mr]kr,\displaystyle\sum_{0<m_{1}<\cdots<m_{r}}\frac{P_{1}(q^{m_{1}})\cdots P_{r}(q^{m_{r}})}{[m_{1}]^{k_{1}}\cdots[m_{r}]^{k_{r}}},

where Pj(x)P_{j}(x) is a polynomial whose degree is less than or equal to kjk_{j} for 1jr1\leq j\leq r. If Pr(0)=0P_{r}(0)=0, the infinite sum (1.2) is absolutely convergent. In this paper we call a value of the form (1.2) a qq-analogue of the MZV (qqMZV) without specifying the model.

To construct a qq-analogue of SMZVs, it would be natural to consider two kinds of regularizations of the sum (1.2) with Pr(0)0P_{r}(0)\not=0 which turn into the harmonic regularized MZV and the shuffle regularized MZV in the limit as q1q\to 1. However, there does not seem to exist any standard definition of a shuffle regularization of the sum (1.2).

To avoid the difficulty, we construct a qq-analogue of SMZVs in a different approach. In [6], Ono, Seki and Yamamoto introduced two kinds of truncations of the tt-adic symmetric multiple zeta value (tt-adic SMZV), which is an element of the formal power series ring (𝒵/ζ(2)𝒵)[[t]](\mathcal{Z}/\zeta(2)\mathcal{Z})[[t]] whose constant term is equal to the SMZV. As a corollary, we have the following expression for ζ𝒮,(𝐤)\zeta^{\mathcal{S},\bullet}(\mathbf{k}). Set

(1.3) ζM𝒮,(k1,,kr)\displaystyle\zeta_{M}^{\mathcal{S},\ast}(k_{1},\ldots,k_{r}) =i=0r(1)ki+1++kr0<m1<<mi<M0<mr<<mi+1<M1m1k1mrkr,\displaystyle=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}<M\\ 0<m_{r}<\cdots<m_{i+1}<M\end{subarray}}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}},
(1.4) ζM𝒮,sh(k1,,kr)\displaystyle\zeta_{M}^{\mathcal{S},\mathcyr{sh}}(k_{1},\ldots,k_{r}) =i=0r(1)ki+1++kr0<m1<<mi0<mr<<mi+1mi+mi+1<M1m1k1mrkr.\displaystyle=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}\\ 0<m_{r}<\cdots<m_{i+1}\\ m_{i}+m_{i+1}<M\end{subarray}}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}}.

Then we have ζ𝒮,(𝐤)=limMζM𝒮,(𝐤)\zeta^{\mathcal{S},\bullet}(\mathbf{k})=\lim_{M\to\infty}\zeta_{M}^{\mathcal{S},\bullet}(\mathbf{k}) for {,sh}\bullet\in\{\ast,\mathcyr{sh}\} [6, Corollary 2.7]. Note that we do not need any regularization here. Hence we could follow the above procedure to construct a qq-analogue of SMZVs. It is the main theme of this paper.

One important example of the relations among SMZVs is the double shuffle relation (see Theorem 2.4 below) due to Kaneko and Zagier. Jarossay proved that the tt-adic SMZVs satisfy a generalization of the double shuffle relation [4]. The point is that the truncated version of the tt-adic SMZVs also satisfies them [6, Theorem 1.6]. As a corollary we see that the truncated version of SMZVs (1.3) and (1.4) satisfies the same double shuffle relation as SMZVs.

In this paper we first construct a qq-analogue of the truncated SMZVs (1.3) and (1.4) which satisfies the double shuffle relation (Proposition 4.1 and Proposition 4.2). Second, we define a qq-analogue of SMZVs as a limit of the above truncated version. As SMZVs are defined to be the elements of the quotient 𝒵/ζ(2)𝒵\mathcal{Z}/\zeta(2)\mathcal{Z}, our qq-analogue of SMZVs is an element of the quotient of the space of the qqMZVs by a sum of two ideals 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q}. Roughly speaking, 𝒩q\mathcal{N}_{q} is the ideal generated by the qqMZVs which turn into zero in the limit as q10q\to 1-0, and 𝒫q\mathcal{P}_{q} is the ideal generated by the values m>0q2km/[m]2k\sum_{m>0}q^{2km}/[m]^{2k} with k1k\geq 1, which turns into ζ(2k)\zeta(2k) in the limit as q10q\to 1-0. Hence the ideal 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q} could be regarded as a qq-analogue of to the ideal ζ(2)𝒵\zeta(2)\mathcal{Z}.

Unlike the truncated SMZVs, the double shuffle relation of the truncated qqSMZVs does not imply that of the qqSMZVs. It is because not all of the truncated qqSMZVs converge in the limit as MM\to\infty and the space spanned by them which converge is not closed under the shuffle product (see Example 6.3 below). Hence our qq-analogue of SMZVs satisfies only a part of the double shuffle relation of SMZVs. At this stage the author does not know how to overcome this point.

The paper is organized as follows. In Section 2 we review on SMZV and its truncated version and the relations of them including the double shuffle relation. Section 3 gives some preliminaries on qqMZVs. In Section 4 we define a qq-analogue of the truncated SMZVs and prove the double shuffle relation. The proof is quite similar to that for the truncated SMZVs given in [6]. In Section 5 we construct a qq-analogue of SMZVs. In Section 6 we see that our qq-analogue shares many of the relations with SMZVs, which are the reversal relation and a part of the double shuffle relation and the Ohno-type relation due to Oyama [7]. Additionally two appendices follow. In Appendix A we discuss the asymptotic behavior of qqMZVs in the limit as q10q\to 1-0 to prove Proposition 3.4 below. Appendix B provides proofs of technical propositions.

Here we give notation used throughout. For a tuple of non-negative integers 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}), we define its weight wt(𝐤)\mathrm{wt}(\mathbf{k}) and depth dep(𝐤)\mathrm{dep}(\mathbf{k}) by wt(𝐤)=i=1rki\mathrm{wt}(\mathbf{k})=\sum_{i=1}^{r}k_{i} and dep(𝐤)=r\mathrm{dep}(\mathbf{k})=r, respectively. We call a tuple of positive integers an index. We regard the empty set \varnothing as an index whose weight and depth are zero. An index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) is said to be admissible if 𝐤=\mathbf{k}=\varnothing, or r1r\geq 1 and kr2k_{r}\geq 2. We denote the set of indices (resp. admissible indices) by II (resp. I0I_{0}). The reversal 𝐤¯\overline{\mathbf{k}} of an index 𝐤=(k1,k2,,kr)\mathbf{k}=(k_{1},k_{2},\ldots,k_{r}) is defined by 𝐤¯=(kr,,k2,k1)\overline{\mathbf{k}}=(k_{r},\ldots,k_{2},k_{1}). For the empty index we set ¯=\overline{\varnothing}=\varnothing.

In this paper we often make use of generating functions in proofs. Then we use the operations defined below without mention. Suppose that 𝔄\mathfrak{A} is a unital algebra over a commutative ring 𝒞\mathcal{C}. Then we extend the addition ++ and the multiplication \cdot on 𝔄\mathfrak{A} to the formal power series ring 𝔄[[X]]\mathfrak{A}[[X]] by f(X)+g(X)=j0(aj+bj)Xjf(X)+g(X)=\sum_{j\geq 0}(a_{j}+b_{j})X^{j} and f(X)g(X)=j,l0(ajbl)Xj+lf(X)\cdot g(X)=\sum_{j,l\geq 0}(a_{j}\cdot b_{l})X^{j+l} for f(X)=j0ajXjf(X)=\sum_{j\geq 0}a_{j}X^{j} and g(X)=l0blXlg(X)=\sum_{l\geq 0}b_{l}X^{l} with aj,bl𝔄a_{j},b_{l}\in\mathfrak{A}, respectively. Similarly, we extend a 𝒞\mathcal{C}-linear map φ:𝔄𝔅\varphi:\mathfrak{A}\rightarrow\mathfrak{B} to the 𝒞\mathcal{C}-linear map φ:𝔄[[X]]𝔅[[X]]\varphi:\mathfrak{A}[[X]]\rightarrow\mathfrak{B}[[X]] by φ(f(X))=j0φ(aj)Xj\varphi(f(X))=\sum_{j\geq 0}\varphi(a_{j})X^{j} for f(X)=j0ajXjf(X)=\sum_{j\geq 0}a_{j}X^{j}. We adopt the above convention for the formal power series ring of several variables.

2. Symmetric multiple zeta value

2.1. Multiple zeta value

For an index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) and a positive integer MM, we define the truncated multiple zeta value ζM(𝐤)\zeta_{M}(\mathbf{k}) by

ζM(𝐤)=0<m1<<mr<M1m1k1mrkr\displaystyle\zeta_{M}(\mathbf{k})=\sum_{0<m_{1}<\cdots<m_{r}<M}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}}

if 𝐤\mathbf{k} is not empty, and ζM()=1\zeta_{M}(\varnothing)=1 for the empty index. If 1Mdep(𝐤)1\leq M\leq\mathrm{dep}(\mathbf{k}), we set ζM(𝐤)=0\zeta_{M}(\mathbf{k})=0.

For an admissible index 𝐤\mathbf{k}, we define the multiple zeta value (MZV) ζ(𝐤)\zeta(\mathbf{k}) by ζ(𝐤)=limMζM(𝐤)\zeta(\mathbf{k})=\lim_{M\to\infty}\zeta_{M}(\mathbf{k}). If 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) is a non-empty admissible index, we have

ζ(𝐤)=0<m1<<mr1m1k1mrkr,\displaystyle\zeta(\mathbf{k})=\sum_{0<m_{1}<\cdots<m_{r}}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}},

which converges since kr2k_{r}\geq 2.

Let 𝔥=x,y\mathfrak{h}=\mathbb{Q}\langle x,y\rangle be the non-commutative polynomial ring of two variables xx and yy over \mathbb{Q}. We set zk=yxk1z_{k}=yx^{k-1} for k1k\geq 1, and z𝐤=zk1zkrz_{\mathbf{k}}=z_{k_{1}}\cdots z_{k_{r}} for a non-empty index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}). For the empty index we set z=1z_{\varnothing}=1.

Set 𝔥1=+y𝔥\mathfrak{h}^{1}=\mathbb{Q}+y\mathfrak{h}. It is a \mathbb{Q}-subalgebra of 𝔥\mathfrak{h}, which can be identified with the non-commutative polynomial ring over \mathbb{Q} with the set of variables {zkk1}\{z_{k}\mid k\geq 1\}. We also set 𝔥0=+y𝔥x\mathfrak{h}^{0}=\mathbb{Q}+y\mathfrak{h}x, which is a \mathbb{Q}-submodule of 𝔥1\mathfrak{h}^{1} with a basis {z𝐤𝐤I0}\{z_{\mathbf{k}}\mid\mathbf{k}\in I_{0}\}.

The harmonic product \ast on 𝔥1\mathfrak{h}^{1} is the \mathbb{Q}-bilinear map :𝔥1×𝔥1𝔥1\ast:\mathfrak{h}^{1}\times\mathfrak{h}^{1}\longrightarrow\mathfrak{h}^{1} defined by the following properties:

  1. (i)

    For any w𝔥1w\in\mathfrak{h}^{1}, it holds that w1=ww\ast 1=w and 1w=w1\ast w=w.

  2. (ii)

    For any w,w𝔥1w,w^{\prime}\in\mathfrak{h}^{1} and k,l1k,l\geq 1, it holds that (wzk)(wzl)=(wwzl)zk+(wzkw)zl+(ww)zk+l(wz_{k})\ast(w^{\prime}z_{l})=(w\ast w^{\prime}z_{l})z_{k}+(wz_{k}\ast w^{\prime})z_{l}+(w\ast w^{\prime})z_{k+l}.

The shuffle product sh:𝔥×𝔥𝔥\mathcyr{sh}:\mathfrak{h}\times\mathfrak{h}\longrightarrow\mathfrak{h} is similarly defined by the following properties and \mathbb{Q}-linearity:

  1. (i)

    For any w𝔥w\in\mathfrak{h}, it holds that wsh 1=ww\,\mathcyr{sh}\,1=w and 1shw=w1\,\mathcyr{sh}\,w=w.

  2. (ii)

    For any w,w𝔥w,w^{\prime}\in\mathfrak{h} and u,v{x,y}u,v\in\{x,y\}, it holds that (wu)sh(wv)=(wshwv)u+(wushw)v(wu)\,\mathcyr{sh}\,(w^{\prime}v)=(w\,\mathcyr{sh}\,w^{\prime}v)u+(wu\,\mathcyr{sh}\,w^{\prime})v.

Then 𝔥1\mathfrak{h}^{1} (resp. 𝔥\mathfrak{h}) becomes a commutative \mathbb{Q}-algebra with respect to the harmonic (resp. shuffle) product, which we denote by 𝔥1\mathfrak{h}^{1}_{\ast} (resp. 𝔥sh\mathfrak{h}_{\mathcyr{sh}}). We see that 𝔥0\mathfrak{h}^{0} is a \mathbb{Q}-subalgebra of 𝔥1\mathfrak{h}^{1}_{\ast} with respect to the harmonic product. We denote it by 𝔥0\mathfrak{h}^{0}_{\ast}. We also see that 𝔥1\mathfrak{h}^{1} and 𝔥0\mathfrak{h}^{0} are \mathbb{Q}-subalgebras of 𝔥sh\mathfrak{h}_{\mathcyr{sh}} with respect to the shuffle product. We denote them by 𝔥sh1\mathfrak{h}^{1}_{\mathcyr{sh}} and 𝔥sh0\mathfrak{h}^{0}_{\mathcyr{sh}}, respectively.

For a positive integer MM, we define the \mathbb{Q}-linear map ZM:𝔥1Z_{M}:\mathfrak{h}^{1}\to\mathbb{Q} by ZM(z𝐤)=ζM(𝐤)Z_{M}(z_{\mathbf{k}})=\zeta_{M}(\mathbf{k}) for an index 𝐤\mathbf{k}. Similarly, we define the \mathbb{Q}-linear map Z:𝔥0Z:\mathfrak{h}^{0}\to\mathbb{R} by Z(z𝐤)=ζ(𝐤)Z(z_{\mathbf{k}})=\zeta(\mathbf{k}), for an admissible index 𝐤\mathbf{k}.

Proposition 2.1.
  1. (i)

    For any w,w𝔥1w,w^{\prime}\in\mathfrak{h}^{1} and M1M\geq 1, it holds that

    ZM(ww)=ZM(w)ZM(w).\displaystyle Z_{M}(w\ast w^{\prime})=Z_{M}(w)Z_{M}(w^{\prime}).

    Hence, if ww and ww^{\prime} belong to 𝔥0\mathfrak{h}^{0}, we have

    (2.1) Z(ww)=Z(w)Z(w).\displaystyle Z(w\ast w^{\prime})=Z(w)Z(w^{\prime}).
  2. (ii)

    For any w,w𝔥0w,w^{\prime}\in\mathfrak{h}^{0}, it holds that

    (2.2) Z(wshw)=Z(w)Z(w).\displaystyle Z(w\,\mathcyr{sh}\,w^{\prime})=Z(w)Z(w^{\prime}).

The relations (2.1) and (2.2) are called the finite double shuffle relation. They imply that the map ZZ is a \mathbb{Q}-algebra homomorphism to \mathbb{R} with respect to both the harmonic and shuffle product. We denote the image of ZZ by 𝒵\mathcal{Z}, which is a \mathbb{Q}-subalgebra of \mathbb{R}.

2.2. Symmetric multiple zeta value

It is known that 𝔥1𝔥0[y]\mathfrak{h}^{1}_{\bullet}\simeq\mathfrak{h}^{0}_{\bullet}[y] for {,sh}\bullet\in\{\ast,\mathcyr{sh}\} (see [2, 8]). Hence, for {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, there uniquely exists the \mathbb{Q}-algebra homomorphism ZZ^{\bullet} from 𝔥1\mathfrak{h}^{1}_{\bullet} to the polynomial ring 𝒵[T]\mathcal{Z}[T] such that Z(y)=TZ^{\bullet}(y)=T. For an index 𝐤\mathbf{k}, we define ζ(𝐤)=Z(z𝐤)|T=0\zeta^{\ast}(\mathbf{k})=Z^{\ast}(z_{\mathbf{k}})|_{T=0} and ζsh(𝐤)=Zsh(z𝐤)|T=0\zeta^{\mathcyr{sh}}(\mathbf{k})=Z^{\mathcyr{sh}}(z_{\mathbf{k}})|_{T=0}. It is known that

(2.3) s0Xsζ(𝐤,1,,1s)=exp(n2(1)n1nζ(n)Xn)s0Xsζsh(𝐤,1,,1s)\displaystyle\sum_{s\geq 0}X^{s}\zeta^{\ast}(\mathbf{k},\underbrace{1,\ldots,1}_{s})=\exp{(\sum_{n\geq 2}\frac{(-1)^{n-1}}{n}\zeta(n)X^{n})}\sum_{s\geq 0}X^{s}\zeta^{\mathcyr{sh}}(\mathbf{k},\underbrace{1,\ldots,1}_{s})

in the formal power series ring 𝒵[[X]]\mathcal{Z}[[X]] for any index 𝐤\mathbf{k} [3, Theorem 1 and Proposition 10].

For an index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, we set

ζ𝒮,(𝐤)=i=0r(1)ki+1++krζ(k1,k2,,ki)ζ(kr,kr1,,ki+1).\displaystyle\zeta^{\mathcal{S},\bullet}(\mathbf{k})=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\zeta^{\bullet}(k_{1},k_{2},\ldots,k_{i})\zeta^{\bullet}(k_{r},k_{r-1},\ldots,k_{i+1}).

From (2.3), we see that the difference ζ𝒮,(𝐤)ζ𝒮,sh(𝐤)\zeta^{\mathcal{S},\ast}(\mathbf{k})-\zeta^{\mathcal{S},\mathcyr{sh}}(\mathbf{k}) belongs to the ideal ζ(2)𝒵=π2𝒵\zeta(2)\mathcal{Z}=\pi^{2}\mathcal{Z} of 𝒵\mathcal{Z}. The symmetric MZV (SMZV) ζ𝒮(𝐤)\zeta_{\mathcal{S}}(\mathbf{k}) is defined by

ζ𝒮(𝐤)=ζ𝒮,(𝐤)modπ2𝒵\displaystyle\zeta^{\mathcal{S}}(\mathbf{k})=\zeta^{\mathcal{S},\bullet}(\mathbf{k})\,\,\hbox{mod}\,\,\pi^{2}\mathcal{Z}

as an element of 𝒵/ζ(2)𝒵\mathcal{Z}/\zeta(2)\mathcal{Z}.

2.3. Truncated symmetric multiple zeta value

We define the \mathbb{Q}-algebra anti-automorphism ψ\psi on 𝔥1z1,z2,\mathfrak{h}^{1}\simeq\mathbb{Q}\langle z_{1},z_{2},\ldots\rangle by ψ(zk)=(1)kzk\psi(z_{k})=(-1)^{k}z_{k} for k1k\geq 1. For {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, we define the \mathbb{Q}-linear map w𝒮,:𝔥1𝔥1w^{\mathcal{S},\bullet}:\mathfrak{h}^{1}\to\mathfrak{h}^{1} by w𝒮,(1)=1w^{\mathcal{S},\bullet}(1)=1 and

w𝒮,(u1ur)=i=0ru1uiψ(ui+1ur)\displaystyle w^{\mathcal{S},\bullet}(u_{1}\cdots u_{r})=\sum_{i=0}^{r}u_{1}\cdots u_{i}\bullet\psi(u_{i+1}\cdots u_{r})

for r1r\geq 1 and u1,,ur{zk}k1u_{1},\ldots,u_{r}\in\{z_{k}\}_{k\geq 1}.

For a positive integer MM and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, we define the \mathbb{Q}-linear map ZM𝒮,:𝔥1Z^{\mathcal{S},\bullet}_{M}:\mathfrak{h}^{1}\to\mathbb{Q} by ZM𝒮,=ZMw𝒮Z^{\mathcal{S},\bullet}_{M}=Z_{M}\circ w_{\mathcal{S}}^{\bullet}. For an index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) and a positive integer MM, the value ZM𝒮,(z𝐤)Z^{\mathcal{S},\bullet}_{M}(z_{\mathbf{k}}) ({,sh})(\bullet\in\{\ast,\mathcyr{sh}\}) is expressed as follows:

ZM𝒮,(z𝐤)\displaystyle Z_{M}^{\mathcal{S},\ast}(z_{\mathbf{k}}) =i=0r(1)ki+1++kr0<m1<<mi<M0<mr<<mi+1<M1m1k1mrkr,\displaystyle=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}<M\\ 0<m_{r}<\cdots<m_{i+1}<M\end{subarray}}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}},
ZM𝒮,sh(z𝐤)\displaystyle Z_{M}^{\mathcal{S},\mathcyr{sh}}(z_{\mathbf{k}}) =i=0r(1)ki+1++kr0<m1<<mi0<mr<<mi+1mi+mi+1<M1m1k1mrkr,\displaystyle=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}\\ 0<m_{r}<\cdots<m_{i+1}\\ m_{i}+m_{i+1}<M\end{subarray}}\frac{1}{m_{1}^{k_{1}}\cdots m_{r}^{k_{r}}},

where m0m_{0} and mr+1m_{r+1} in the condition mi+mi+1<Mm_{i}+m_{i+1}<M are set equal to zero in the latter formula. We call the above values the truncated symmetric multiple zeta value.

Theorem 2.2 ([4, 6]).
  1. (i)

    For any w,w𝔥1w,w^{\prime}\in\mathfrak{h}^{1} and M1M\geq 1, it holds that

    ZM𝒮,(ww)=ZM𝒮,(w)ZM𝒮,(w),ZM𝒮,sh(wshw)=ZM𝒮,sh(wψ(w)).\displaystyle Z_{M}^{\mathcal{S},\ast}(w\ast w^{\prime})=Z_{M}^{\mathcal{S},\ast}(w)Z_{M}^{\mathcal{S},\ast}(w^{\prime}),\qquad Z_{M}^{\mathcal{S},\mathcyr{sh}}(w\,\mathcyr{sh}\,w^{\prime})=Z_{M}^{\mathcal{S},\mathcyr{sh}}(w\psi(w^{\prime})).
  2. (ii)

    It holds that w𝒮,(𝔥1)𝔥0w^{\mathcal{S},\bullet}(\mathfrak{h}^{1})\subset\mathfrak{h}^{0} for {,sh}\bullet\in\{\ast,\mathcyr{sh}\}. Hence the limit limMZM𝒮,(w)\lim_{M\to\infty}Z_{M}^{\mathcal{S},\bullet}(w) converges for any w𝔥1w\in\mathfrak{h}^{1} and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}.

  3. (iii)

    For any index 𝐤\mathbf{k} and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, it holds that

    limMZM𝒮,(z𝐤)=ζ𝒮,(𝐤).\displaystyle\lim_{M\to\infty}Z_{M}^{\mathcal{S},\bullet}(z_{\mathbf{k}})=\zeta^{\mathcal{S},\bullet}(\mathbf{k}).

2.4. Relations of SMZVs

From the definition of ζ𝒮,\zeta^{\mathcal{S},\bullet}, we obtain the following equality, which is called the reversal relation:

Theorem 2.3.

For any index 𝐤\mathbf{k}, it holds that

ζ𝒮(𝐤¯)=(1)wt(𝐤)ζ𝒮(𝐤).\displaystyle\zeta^{\mathcal{S}}(\overline{\mathbf{k}})=(-1)^{\mathrm{wt}(\mathbf{k})}\zeta^{\mathcal{S}}(\mathbf{k}).

For indices 𝐤,𝐥\mathbf{k},\mathbf{l} and 𝐦\mathbf{m}, we define the non-negative integer d𝐤,𝐥,𝐦({,sh})d_{\mathbf{k},\mathbf{l}}^{\,\bullet,\mathbf{m}}\,(\bullet\in\{\ast,\mathcyr{sh}\}) by

(2.4) z𝐤z𝐥=𝐦d𝐤,𝐥,𝐦z𝐦.\displaystyle z_{\mathbf{k}}\bullet z_{\mathbf{l}}=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\bullet,\mathbf{m}}z_{\mathbf{m}}.

The following relation is called the double shuffle relation of SMZVs, which can be obtained from Theorem 2.2.

Theorem 2.4.

For any index 𝐤\mathbf{k} and 𝐥\mathbf{l}, it holds that

ζ𝒮(𝐤)ζ𝒮(𝐥)=𝐦d𝐤,𝐥,𝐦ζ𝒮(𝐦),\displaystyle\zeta^{\mathcal{S}}(\mathbf{k})\zeta^{\mathcal{S}}(\mathbf{l})=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\ast,\mathbf{m}}\zeta^{\mathcal{S}}(\mathbf{m}),
(1)wt(𝐥)ζ𝒮(𝐤,𝐥¯)=𝐦d𝐤,𝐥sh,𝐦ζ𝒮(𝐦).\displaystyle(-1)^{\mathrm{wt}(\mathbf{l})}\zeta^{\mathcal{S}}(\mathbf{k},\overline{\mathbf{l}})=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\mathcyr{sh},\mathbf{m}}\zeta^{\mathcal{S}}(\mathbf{m}).

In [7], Oyama proved that Theorem 2.4 and the identity ζ𝒮(k,,k)=0\zeta^{\mathcal{S}}(k,\ldots,k)=0 for any k1k\geq 1 imply the Ohno-type relation. To write down it, we define the Hoffman dual 𝐤\mathbf{k}^{\vee} of an index 𝐤\mathbf{k}. We define the \mathbb{Q}-algebra automorphism τ\tau on 𝔥\mathfrak{h} by τ(x)=y\tau(x)=y and τ(y)=x\tau(y)=x. For a non-empty index 𝐤\mathbf{k}, the monomial z𝐤z_{\mathbf{k}} is written in the form z𝐤=ywz_{\mathbf{k}}=yw with a monomial w𝔥w\in\mathfrak{h}. Then the Hoffman dual index of 𝐤\mathbf{k} is defined to be the index 𝐤\mathbf{k}^{\vee} satisfying z𝐤=yτ(w)z_{\mathbf{k}^{\vee}}=y\tau(w). For example, if 𝐤=(3,1,2,1)\mathbf{k}=(3,1,2,1), we have z𝐤=yx2y2xyz_{\mathbf{k}}=yx^{2}y^{2}xy and z𝐤=yτ(x2y2xy)=y3x2yx=z12z3z2z_{\mathbf{k}^{\vee}}=y\tau(x^{2}y^{2}xy)=y^{3}x^{2}yx=z_{1}^{2}z_{3}z_{2}, hence 𝐤=(1,1,3,2)\mathbf{k}^{\vee}=(1,1,3,2). For the empty index we set =\varnothing^{\vee}=\varnothing.

Theorem 2.5 (Ohno-type relation [7]).

For any index 𝐤\mathbf{k} and any non-negative integer mm, it holds that

𝐞(0)rwt(𝐞)=mζ𝒮(𝐤+𝐞)=𝐞(0)swt(𝐞)=mζ𝒮((𝐤+𝐞)),\displaystyle\sum_{\begin{subarray}{c}\mathbf{e}\in(\mathbb{Z}_{\geq 0})^{r}\\ \mathrm{wt}(\mathbf{e})=m\end{subarray}}\zeta^{\mathcal{S}}(\mathbf{k}+\mathbf{e})=\sum_{\begin{subarray}{c}\mathbf{e}\in(\mathbb{Z}_{\geq 0})^{s}\\ \mathrm{wt}(\mathbf{e})=m\end{subarray}}\zeta^{\mathcal{S}}((\mathbf{k}^{\vee}+\mathbf{e})^{\vee}),

where r=dep(𝐤)r=\mathrm{dep}(\mathbf{k}) and s=dep(𝐤)s=\mathrm{dep}(\mathbf{k}^{\vee}).

3. A qq-analogue of multiple zeta value

3.1. Algebraic formulation

Set 𝒞=[]\mathcal{C}=\mathbb{Q}[\hbar], where \hbar is a formal variable. Let =𝒞a,b\mathfrak{H}=\mathcal{C}\langle a,b\rangle be the non-commutative polynomial ring of two variables aa and bb over 𝒞\mathcal{C}. For k1k\geq 1 we set

gk=bak,ek=b(a+)ak1.\displaystyle g_{k}=ba^{k},\qquad e_{k}=b(a+\hbar)a^{k-1}.

They are related to each other as

gk=()k1g1+j=2k()kjej\displaystyle g_{k}=(-\hbar)^{k-1}g_{1}+\sum_{j=2}^{k}(-\hbar)^{k-j}e_{j}

for k1k\geq 1, and

(3.1) ek=gk+gk1\displaystyle e_{k}=g_{k}+\hbar g_{k-1}

for k2k\geq 2. Note that e1g1=be_{1}-g_{1}=\hbar b.

We set

A={b}{bakk1}={e1g1}{gkk1},\displaystyle A=\{\hbar b\}\cup\{ba^{k}\mid k\geq 1\}=\{e_{1}-g_{1}\}\cup\{g_{k}\mid k\geq 1\},

which is an algebraically independent set, and denote by 𝒞A\mathcal{C}\langle A\rangle the 𝒞\mathcal{C}-subalgebra of \mathfrak{H} generated by 11 and AA. The depth of a monomial u1ur(u1,,urA)u_{1}\cdots u_{r}\,(u_{1},\ldots,u_{r}\in A) is defined to be rr.

We define the 𝒞\mathcal{C}-submodule 0^\widehat{\mathfrak{H}^{0}} of 𝒞A\mathcal{C}\langle A\rangle by

0^=𝒞+k1𝒞Agk.\displaystyle\widehat{\mathfrak{H}^{0}}=\mathcal{C}+\sum_{k\geq 1}\mathcal{C}\langle A\rangle g_{k}.

Then the set consisting of the elements

(3.2) (e1g1)α1gβ1+1(e1g1)αrgβr+1\displaystyle(e_{1}-g_{1})^{\alpha_{1}}g_{\beta_{1}+1}\cdots(e_{1}-g_{1})^{\alpha_{r}}g_{\beta_{r}+1}

with r0r\geq 0 and α1,,αr,β1,,βr0\alpha_{1},\ldots,\alpha_{r},\beta_{1},\ldots,\beta_{r}\geq 0 forms a free basis of 0^\widehat{\mathfrak{H}^{0}}.

Let qq be a complex parameter satisfying 0<|q|<10<|q|<1. We endow the complex number field \mathbb{C} with 𝒞\mathcal{C}-module structure such that \hbar acts as multiplication by 1q1-q.

We denote by 𝔷\mathfrak{z} the 𝒞\mathcal{C}-submodule of 𝒞A\mathcal{C}\langle A\rangle spanned by the set AA. For a positive integer mm, we define the 𝒞\mathcal{C}-linear map Fq(m;):𝔷F_{q}(m;\cdot):\mathfrak{z}\longrightarrow\mathbb{C} by

(3.3) Fq(m;e1g1)=1q,Fq(m;gk)=qkm[m]k\displaystyle F_{q}(m;e_{1}-g_{1})=1-q,\qquad F_{q}(m;g_{k})=\frac{q^{km}}{[m]^{k}}

for k1k\geq 1, where [m][m] is the qq-integer

[m]=1qm1q=1+q++qm1.\displaystyle[m]=\frac{1-q^{m}}{1-q}=1+q+\cdots+q^{m-1}.

Then, from e1=(e1g1)+g1e_{1}=(e_{1}-g_{1})+g_{1} and (3.1) for k2k\geq 2, we have

(3.4) Fq(m;ek)=q(k1)m[m]k\displaystyle F_{q}(m;e_{k})=\frac{q^{(k-1)m}}{[m]^{k}}

for any k1k\geq 1 because (1q)[m]+qm=1(1-q)[m]+q^{m}=1.

For a positive integer MM, we define the 𝒞\mathcal{C}-linear map Zq,M:𝒞AZ_{q,M}:\mathcal{C}\langle A\rangle\longrightarrow\mathbb{C} by Zq,M(1)=1Z_{q,M}(1)=1 and

Zq,M(u1ur)=0<m1<<mr<Mi=1rFq(mi;ui)\displaystyle Z_{q,M}(u_{1}\cdots u_{r})=\sum_{0<m_{1}<\cdots<m_{r}<M}\prod_{i=1}^{r}F_{q}(m_{i};u_{i})

for u1,,urAu_{1},\ldots,u_{r}\in A. Then we see that, if w0^w\in\widehat{\mathfrak{H}^{0}}, Zq,M(w)Z_{q,M}(w) converges in the limit as MM\to\infty. Hence we can define the 𝒞\mathcal{C}-linear map Zq:0^Z_{q}:\widehat{\mathfrak{H}^{0}}\longrightarrow\mathbb{C} by

Zq(w)=limMZq,M(w)\displaystyle Z_{q}(w)=\lim_{M\to\infty}Z_{q,M}(w)

for w0^w\in\widehat{\mathfrak{H}^{0}}. In this paper we call the value Zq(w)Z_{q}(w) with w0^w\in\widehat{\mathfrak{H}^{0}} a qq-analogue of MZV (qqMZV).

Remark 3.1.

There are various models of qq-analogue of MZV (see [11, Chapter 12]). We can represent them in the form of Zq(w)Z_{q}(w) with some w0^w\in\widehat{\mathfrak{H}^{0}}. For example, from (3.3) and (3.4), we see that

Zq(gk1gkr)=0<m1<<mrqk1m1++krmr[m1]k1[mr]kr,\displaystyle Z_{q}(g_{k_{1}}\cdots g_{k_{r}})=\sum_{0<m_{1}<\cdots<m_{r}}\frac{q^{k_{1}m_{1}+\cdots+k_{r}m_{r}}}{[m_{1}]^{k_{1}}\cdots[m_{r}]^{k_{r}}},
Zq(ek1ekr)=0<m1<<mrq(k11)m1++(kr1)mr[m1]k1[mr]kr,\displaystyle Z_{q}(e_{k_{1}}\cdots e_{k_{r}})=\sum_{0<m_{1}<\cdots<m_{r}}\frac{q^{(k_{1}-1)m_{1}+\cdots+(k_{r}-1)m_{r}}}{[m_{1}]^{k_{1}}\cdots[m_{r}]^{k_{r}}},

which are called the Schlesinger-Zudilin model and the Bradley-Zhao model, respectively.

3.2. Double shuffle relation of qqMZVs

The harmonic product and the shuffle product associated with the qqMZV are defined as follows.

First, we define the symmetric 𝒞\mathcal{C}-bilinear map :𝔷×𝔷𝔷\circ_{\hbar}:\mathfrak{z}\times\mathfrak{z}\to\mathfrak{z} by

(e1g1)(e1g1)=(e1g1),(e1g1)gk=gk,gkgl=gk+l\displaystyle(e_{1}-g_{1})\circ_{\hbar}(e_{1}-g_{1})=\hbar(e_{1}-g_{1}),\quad(e_{1}-g_{1})\circ_{\hbar}g_{k}=\hbar g_{k},\quad g_{k}\circ_{\hbar}g_{l}=g_{k+l}

for k,l1k,l\geq 1. Then we see that

(3.5) Fq(m;uv)=Fq(m;u)Fq(m;v)\displaystyle F_{q}(m;u\circ_{\hbar}v)=F_{q}(m;u)F_{q}(m;v)

for any u,v𝔷u,v\in\mathfrak{z} and m1m\geq 1. The harmonic product \ast_{\hbar} is the 𝒞\mathcal{C}-bilinear binary operation on 𝒞A\mathcal{C}\langle A\rangle uniquely defined by the following properties:

  1. (i)

    For any w𝒞Aw\in\mathcal{C}\langle A\rangle, it holds that w1=ww\ast_{\hbar}1=w and 1w=w1\ast_{\hbar}w=w.

  2. (ii)

    For any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and u,vAu,v\in A, it holds that (wu)(wv)=(wwv)u+(wuw)v+(ww)(uv)(wu)\ast_{\hbar}(w^{\prime}v)=(w\ast_{\hbar}w^{\prime}v)u+(wu\ast_{\hbar}w^{\prime})v+(w\ast_{\hbar}w^{\prime})(u\circ_{\hbar}v).

The harmonic product \ast_{\hbar} on 𝒞A\mathcal{C}\langle A\rangle is commutative and associative.

Proposition 3.2.

The 𝒞\mathcal{C}-submodule 0^\widehat{\mathfrak{H}^{0}} of 𝒞A\mathcal{C}\langle A\rangle is closed under the harmonic product \ast_{\hbar}, and it holds that Zq,M(ww)=Zq,M(w)Zq,M(w)Z_{q,M}(w\ast_{\hbar}w^{\prime})=Z_{q,M}(w)Z_{q,M}(w^{\prime}) for any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and M1M\geq 1. Therefore, we have Zq(ww)=Zq(w)Zq(w)Z_{q}(w\ast_{\hbar}w^{\prime})=Z_{q}(w)Z_{q}(w^{\prime}) for any w,w0^w,w^{\prime}\in\widehat{\mathfrak{H}^{0}}.

Next, we define the shuffle product. Consider the \mathbb{Q}-linear right action of \mathfrak{H} on a \mathbb{C}-valued function f(t)f(t) defined by

(f)(t)=(1q)f(t),(fa)(t)=(1q)j=1f(qjt),(fb)(t)=t1tf(t).\displaystyle(f\hbar)(t)=(1-q)f(t),\qquad(fa)(t)=(1-q)\sum_{j=1}^{\infty}f(q^{j}t),\qquad(fb)(t)=\frac{t}{1-t}f(t).

Then, for any function ff and gg, it holds that

(3.6) faga=(fag+fga+(1q)fg)a,fbg=fgb=(fg)b\displaystyle fa\cdot ga=(fa\cdot g+f\cdot ga+(1-q)f\cdot g)a,\quad fb\cdot g=f\cdot gb=(f\cdot g)b

if all terms are well-defined, where \cdot denotes the usual multiplication of functions. Motivated by the above relations we define the shuffle product on \mathfrak{H} as the 𝒞\mathcal{C}-bilinear binary operation uniquely defined by the following properties:

  1. (i)

    For any ww\in\mathfrak{H}, it holds that wsh 1=ww\,\mathcyr{sh}_{\hbar}\,1=w and 1shw=w1\,\mathcyr{sh}_{\hbar}\,w=w.

  2. (ii)

    For any w,ww,w^{\prime}\in\mathfrak{H}, it holds that

    washwa=(washw+wshwa+wshw)a,\displaystyle wa\,\mathcyr{sh}_{\hbar}\,w^{\prime}a=(wa\,\mathcyr{sh}_{\hbar}\,w^{\prime}+w\,\mathcyr{sh}_{\hbar}\,w^{\prime}a+\hbar w\,\mathcyr{sh}_{\hbar}\,w^{\prime})a,
    wbshw=wshwb=(wshw)b.\displaystyle wb\,\mathcyr{sh}_{\hbar}\,w^{\prime}=w\,\mathcyr{sh}_{\hbar}\,w^{\prime}b=(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})b.

The shuffle product is commutative and associative. Note that

(3.7) (w(e1g1))shw=wsh(w(e1g1))=(wshw)(e1g1)\displaystyle(w(e_{1}-g_{1}))\,\mathcyr{sh}_{\hbar}\,w^{\prime}=w\,\mathcyr{sh}_{\hbar}\,(w^{\prime}(e_{1}-g_{1}))=(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})(e_{1}-g_{1})

for any w,ww,w^{\prime}\in\mathfrak{H} since e1g1=be_{1}-g_{1}=\hbar b.

Proposition 3.3.

The 𝒞\mathcal{C}-submodules 𝒞A\mathcal{C}\langle A\rangle and 0^\widehat{\mathfrak{H}^{0}} of \mathfrak{H} are closed under the shuffle product sh\mathcyr{sh}_{\hbar}. For any u1,,ur,v1,,vs𝔷u_{1},\ldots,u_{r},v_{1},\ldots,v_{s}\in\mathfrak{z} and M1M\geq 1, it holds that

(3.8) Zq,M(u1urshv1vs)=0<m1<<mr0<n1<<nsmr+ns<Mi=1rFq(mi;ui)j=1sFq(nj;vj).\displaystyle Z_{q,M}(u_{1}\cdots u_{r}\,\mathcyr{sh}_{\hbar}\,v_{1}\cdots v_{s})=\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{r}\\ 0<n_{1}<\cdots<n_{s}\\ m_{r}+n_{s}<M\end{subarray}}\prod_{i=1}^{r}F_{q}(m_{i};u_{i})\prod_{j=1}^{s}F_{q}(n_{j};v_{j}).

Therefore, we have Zq(wshw)=Zq(w)Zq(w)Z_{q}(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})=Z_{q}(w)Z_{q}(w^{\prime}) for any w,w0^w,w^{\prime}\in\widehat{\mathfrak{H}^{0}}.

Proof.

Note that gk+1=gkag_{k+1}=g_{k}a for k1k\geq 1. For w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and k,l1k,l\geq 1, it holds that

(3.9) wg1shwg1\displaystyle wg_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{1} =(wshwg1+wg1shw+(wshw)(e1g1))g1,\displaystyle=(w\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{1}+wg_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}+(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})(e_{1}-g_{1}))g_{1},
(3.10) wg1shwgk+1\displaystyle wg_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{k+1} =(wshwgk+1)g1+(wg1shwgk)a+(wshwgk)g1,\displaystyle=(w\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{k+1})g_{1}+(wg_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{k})a+\hbar(w\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{k})g_{1},
(3.11) wgk+1shwgl+1\displaystyle wg_{k+1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{l+1} =(wgkshwgl+1+wgk+1shwgl+wgkshwgl)a.\displaystyle=(wg_{k}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{l+1}+wg_{k+1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{l}+\hbar wg_{k}\,\mathcyr{sh}_{\hbar}\,w^{\prime}g_{l})a.

These formulas imply that

(3.12) 𝒞Agksh𝒞Aglmin(k,l)jk+l1𝒞Agj\displaystyle\mathcal{C}\langle A\rangle g_{k}\,\mathcyr{sh}_{\hbar}\,\mathcal{C}\langle A\rangle g_{l}\subset\sum_{\min{(k,l)}\leq j\leq k+l-1}\mathcal{C}\langle A\rangle g_{j}

for k,l1k,l\geq 1. From (3.7) and (3.12), we see that 𝒞A\mathcal{C}\langle A\rangle and 0^\widehat{\mathfrak{H}^{0}} are closed under the shuffle product.

Denote by \mathcal{F} the \mathbb{C}-vector space of holomorphic functions on the unit disc. We endow \mathcal{F} with 𝒞\mathcal{C}-module structure such that \hbar acts as multiplication by 1q1-q. Let 𝟏(t)=1\mathbf{1}(t)=1 be the constant function. We define the 𝒞\mathcal{C}-linear map Lq():𝒞AL_{q}(\cdot):\mathcal{C}\langle A\rangle\longrightarrow\mathcal{F} by

Lq(u1ur)(t)=((𝟏)u1ur)(t),\displaystyle L_{q}(u_{1}\cdots u_{r})(t)=((\mathbf{1})u_{1}\cdots u_{r})(t),

where u1,,urAu_{1},\ldots,u_{r}\in A. We see that

Lq(u1ur)(t)=0<m1<<mrtmri=1rFq(mi;ui)\displaystyle L_{q}(u_{1}\cdots u_{r})(t)=\sum_{0<m_{1}<\cdots<m_{r}}t^{m_{r}}\prod_{i=1}^{r}F_{q}(m_{i};u_{i})

for u1,,ur𝔷u_{1},\ldots,u_{r}\in\mathfrak{z}. Hence, Zq,M(w)Z_{q,M}(w) is equal to the sum of the coefficients of tmt^{m} in Lq(w)(t)L_{q}(w)(t) over 0m<M0\leq m<M for w𝒞Aw\in\mathcal{C}\langle A\rangle. By (3.6), we have Lq(wshw)=Lq(w)Lq(w)L_{q}(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})=L_{q}(w)L_{q}(w^{\prime}) for any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle. Thus we obtain (3.8). ∎

As stated in Proposition 3.2 and Proposition 3.3, we have

(3.13) Zq(ww)=Zq(w)Zq(w),Zq(wshw)=Zq(w)Zq(w)\displaystyle Z_{q}(w\ast w^{\prime})=Z_{q}(w)Z_{q}(w^{\prime}),\qquad Z_{q}(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})=Z_{q}(w)Z_{q}(w^{\prime})

for any w,w0^w,w^{\prime}\in\widehat{\mathfrak{H}^{0}}. We call them the double shuffle relation for the qqMZVs. We denote by 𝒵q^\widehat{\mathcal{Z}_{q}} the image of the map Zq:0^Z_{q}:\widehat{\mathfrak{H}^{0}}\rightarrow\mathbb{C}. Then 𝒵q^\widehat{\mathcal{Z}_{q}} is a 𝒞\mathcal{C}-subalgebra of \mathbb{C}.

3.3. Limit of qqMZVs as q10q\to 1-0

We denote by 𝔫0\mathfrak{n}_{0} the 𝒞\mathcal{C}-submodule of 0^\widehat{\mathfrak{H}^{0}} spanned by the elements of the form (3.2) with r1r\geq 1 and αs,βt1\alpha_{s},\beta_{t}\geq 1 for some 1str1\leq s\leq t\leq r. We define the 𝒞\mathcal{C}-modules 𝔫\mathfrak{n} and 0\mathfrak{H}^{0} by

𝔫=𝔫0+0^,0=𝒞+k2𝒞Agk+𝔫.\displaystyle\mathfrak{n}=\mathfrak{n}_{0}+\hbar\,\widehat{\mathfrak{H}^{0}},\qquad\mathfrak{H}^{0}=\mathcal{C}+\sum_{k\geq 2}\mathcal{C}\langle A\rangle g_{k}+\mathfrak{n}.

For a non-empty index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}), we set

g𝐤=gk1gkr.\displaystyle g_{\mathbf{k}}=g_{k_{1}}\cdots g_{k_{r}}.

For the empty index, we set g=1g_{\varnothing}=1. Note that the quotient 0/𝔫\mathfrak{H}^{0}/\mathfrak{n} is a free \mathbb{Q}-module which has a basis {g𝐤𝐤I0}\{g_{\mathbf{k}}\mid\mathbf{k}\in I_{0}\}.

Proposition 3.4.

Here we consider a limit as q10q\to 1-0 with qq being real.

  1. (i)

    If 𝐤\mathbf{k} is an admissible index, it holds that limq10Zq(g𝐤)=ζ(𝐤)\lim_{q\to 1-0}Z_{q}(g_{\mathbf{k}})=\zeta(\mathbf{k}).

  2. (ii)

    For any w𝔫w\in\mathfrak{n}, it holds that limq10Zq(w)=0\lim_{q\to 1-0}Z_{q}(w)=0.

See Appendix A for the proof.

We define the unital \mathbb{Q}-algebra homomorphism ι:𝒞A𝔥1\iota:\mathcal{C}\langle A\rangle\rightarrow\mathfrak{h}^{1} by

(3.14) ι()=0,ι(e1g1)=0,ι(gk)=zk\displaystyle\iota(\hbar)=0,\quad\iota(e_{1}-g_{1})=0,\quad\iota(g_{k})=z_{k}

for k1k\geq 1. Then we see that the restriction of ι\iota to 0\mathfrak{H}^{0} is a surjection onto 𝔥0\mathfrak{h}^{0} and its kernel is equal to 𝔫\mathfrak{n}. Therefore, from Proposition 3.4, we obtain the following corollary.

Corollary 3.5.

For any w0w\in\mathfrak{H}^{0}, it holds that

(3.15) limq10Zq(w)=Z(ι(w)).\displaystyle\lim_{q\to 1-0}Z_{q}(w)=Z(\iota(w)).
Remark 3.6.

The equality (3.15) does not necessarily hold for w0^0w\in\widehat{\mathfrak{H}^{0}}\setminus\mathfrak{H}^{0} such that the limit of Zq(w)Z_{q}(w) as q10q\to 1-0 converges. For example, we see that

Zq((e1g1)g1)\displaystyle Z_{q}((e_{1}-g_{1})g_{1}) =(1q)20<m<nqn1qn=(1q)20<n,l(n1)qnl\displaystyle=(1-q)^{2}\sum_{0<m<n}\frac{q^{n}}{1-q^{n}}=(1-q)^{2}\sum_{0<n,l}(n-1)q^{nl}
=(1q)20<lq2l(1ql)2=Zq(g2),\displaystyle=(1-q)^{2}\sum_{0<l}\frac{q^{2l}}{(1-q^{l})^{2}}=Z_{q}(g_{2}),

which is a special case of the resummation identity

Zq((e1g1)α1gβ1+1(e1g1)αrgβr+1)=Zq((e1g1)βrgαr+1(e1g1)β1gα1+1)\displaystyle Z_{q}((e_{1}-g_{1})^{\alpha_{1}}g_{\beta_{1}+1}\cdots(e_{1}-g_{1})^{\alpha_{r}}g_{\beta_{r}+1})=Z_{q}((e_{1}-g_{1})^{\beta_{r}}g_{\alpha_{r}+1}\cdots(e_{1}-g_{1})^{\beta_{1}}g_{\alpha_{1}+1})

for α1,,αr,β1,,βr0\alpha_{1},\ldots,\alpha_{r},\beta_{1},\ldots,\beta_{r}\geq 0 [9, Theorem 4]. Hence, the limit of Zq((e1g1)g1)Z_{q}((e_{1}-g_{1})g_{1}) as q10q\to 1-0 is equal to ζ(2)\zeta(2). However, we have ι((e1g1)g1)=0\iota((e_{1}-g_{1})g_{1})=0.

3.4. Restoration of finite double shuffle relation of MZVs

Proposition 3.7.

The 𝒞\mathcal{C}-module 𝔫\mathfrak{n} is an ideal of 0^\widehat{\mathfrak{H}^{0}} with respect to both the harmonic product and the shuffle product.

Proof.

It suffices to show that 𝔫00^𝔫\mathfrak{n}_{0}\ast_{\hbar}\widehat{\mathfrak{H}^{0}}\subset\mathfrak{n} and 𝔫0sh0^𝔫\mathfrak{n}_{0}\,\mathcyr{sh}_{\hbar}\,\widehat{\mathfrak{H}^{0}}\subset\mathfrak{n}. The former follows from the definition of the harmonic product. The latter follows from (3.7) and (3.12). ∎

Corollary 3.8.

The 𝒞\mathcal{C}-module 0\mathfrak{H}^{0} is closed under both the harmonic product and the shuffle product.

Proof.

From gkgl=gk+lg_{k}\circ_{\hbar}g_{l}=g_{k+l}, we see that k2𝒞Agk\sum_{k\geq 2}\mathcal{C}\langle A\rangle g_{k} is closed under the harmonic product. It is also closed under the shuffle product because of (3.12). Hence Proposition 3.7 implies that 0\mathfrak{H}^{0} is closed. ∎

We set

𝒵q=Zq(0).\displaystyle\mathcal{Z}_{q}=Z_{q}(\mathfrak{H}^{0}).

Corollary 3.8 implies that 𝒵q\mathcal{Z}_{q} is a 𝒞\mathcal{C}-subalgebra of 𝒵q^\widehat{\mathcal{Z}_{q}}.

Proposition 3.9.

For any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, it holds that

(3.16) ι(ww)=ι(w)ι(w).\displaystyle\iota(w\bullet_{\hbar}w^{\prime})=\iota(w)\bullet\iota(w^{\prime}).
Proof.

We may assume that ww and ww^{\prime} are monomials in AA. If w=1w=1 or w=1w^{\prime}=1, the desired equality (3.16) is trivial. Now we proceed the proof by induction on the sum of the depth of ww and ww^{\prime}.

Since ι((e1g1)u)=0\iota((e_{1}-g_{1})\circ_{\hbar}u)=0 for any u𝔷u\in\mathfrak{z} and ι(gkgl)=zk+l\iota(g_{k}\circ_{\hbar}g_{l})=z_{k+l} for k,l1k,l\geq 1, we see that (3.16) holds for =\bullet=\ast.

We consider the case of =sh\bullet=\mathcyr{sh}. If ww or ww^{\prime} is the form of u(e1g1)u(e_{1}-g_{1}) with a monomial uu of AA, we see that (3.16) from ι(e1g1)=0\iota(e_{1}-g_{1})=0 and (3.7). Hence it suffices to prove the case where w=ugkw=ug_{k} and w=vglw^{\prime}=vg_{l} with monomials u,vu,v in AA and k,l1k,l\geq 1. For that purpose we use the following formula. Note that, in the formal power series ring [[X]]\mathfrak{H}[[X]], we have

k=1gkXk1=g111aX.\displaystyle\sum_{k=1}^{\infty}g_{k}X^{k-1}=g_{1}\frac{1}{1-aX}.
Lemma 3.10.

For any u,vu,v\in\mathfrak{H}, we have

ug111aXshvg111aY\displaystyle ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,vg_{1}\frac{1}{1-aY}
={(1+X)(ug111aXshv)+(1+Y)(ushvg111aY)+(ushv)(e1g1)}\displaystyle=\left\{(1+\hbar X)\left(ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,v\right)+(1+\hbar Y)\left(u\,\mathcyr{sh}_{\hbar}\,vg_{1}\frac{1}{1-aY}\right)+(u\,\mathcyr{sh}_{\hbar}\,v)(e_{1}-g_{1})\right\}
×g111a(X+Y+XY).\displaystyle\times g_{1}\frac{1}{1-a(X+Y+\hbar XY)}.

See Appendix B.1 for the proof. From the above equality, we see that

ι(ug111aXshvg111aY)=ι(ug111aXshv+ushvg111aY)z111x(X+Y)\displaystyle\iota(ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,vg_{1}\frac{1}{1-aY})=\iota(ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,v+u\,\mathcyr{sh}_{\hbar}\,vg_{1}\frac{1}{1-aY})z_{1}\frac{1}{1-x(X+Y)}

in the formal power series ring 𝔥[[X,Y]]\mathfrak{h}[[X,Y]]. The induction hypothesis implies that the right hand side is equal to

{ι(u)z111xXshι(v)+ι(u)shι(v)z111xY}z111(X+Y)x.\displaystyle\left\{\iota(u)z_{1}\frac{1}{1-xX}\,\mathcyr{sh}\,\iota(v)+\iota(u)\,\mathcyr{sh}\,\iota(v)z_{1}\frac{1}{1-xY}\right\}z_{1}\frac{1}{1-(X+Y)x}.

We see that it is equal to

ι(u)z111xXshι(v)z111xY=ι(ug111aX)shι(vg111aY)\displaystyle\iota(u)z_{1}\frac{1}{1-xX}\,\mathcyr{sh}\,\iota(v)z_{1}\frac{1}{1-xY}=\iota(ug_{1}\frac{1}{1-aX})\,\mathcyr{sh}\,\iota(vg_{1}\frac{1}{1-aY})

in the same way as the proof of Lemma 3.10. This completes the proof of Proposition 3.9 for =sh\bullet=\mathcyr{sh}. ∎

Using Corollary 3.8 and Proposition 3.9, we restore the finite double shuffle relation of MZVs from (3.13) as follows. Let 𝐤\mathbf{k} and 𝐥\mathbf{l} be admissible indices. Then, for {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, we see that

Z(z𝐤)Z(z𝐥)\displaystyle Z(z_{\mathbf{k}})Z(z_{\mathbf{l}}) =limq10Zq(g𝐤)Zq(g𝐥)=limq10Zq(g𝐤g𝐥)\displaystyle=\lim_{q\to 1-0}Z_{q}(g_{\mathbf{k}})Z_{q}(g_{\mathbf{l}})=\lim_{q\to 1-0}Z_{q}(g_{\mathbf{k}}\bullet g_{\mathbf{l}})
=Z(ι(g𝐤g𝐥))=Z(ι(g𝐤)ι(g𝐥))=Z(z𝐤z𝐥).\displaystyle=Z(\iota(g_{\mathbf{k}}\bullet g_{\mathbf{l}}))=Z(\iota(g_{\mathbf{k}})\bullet\iota(g_{\mathbf{l}}))=Z(z_{\mathbf{k}}\bullet z_{\mathbf{l}}).

4. A qq-analogue of truncated SMZV

Let ψ({,sh})\psi^{\bullet}\,(\bullet\in\{\ast,\mathcyr{sh}\}) be the 𝒞\mathcal{C}-algebra anti-involution on 𝒞A=𝒞b,ba,ba2,\mathcal{C}\langle A\rangle=\mathcal{C}\langle\hbar b,ba,ba^{2},\ldots\rangle defined by

ψ(b)=b,ψ(bak)=b(a)k,\displaystyle\psi^{\ast}(\hbar b)=\hbar b,\quad\psi^{\ast}(ba^{k})=b(-a)^{k},
ψsh(b)=b,ψsh(bak)=b(a)k\displaystyle\psi^{\mathcyr{sh}}(\hbar b)=\hbar b,\quad\psi^{\mathcyr{sh}}(ba^{k})=b(-a-\hbar)^{k}

for k1k\geq 1. For {,sh}\bullet\in\{\ast,\mathcyr{sh}\}, we define the 𝒞\mathcal{C}-linear map w𝒮,:𝒞A𝒞Aw^{\mathcal{S},\bullet}_{\hbar}:\mathcal{C}\langle A\rangle\to\mathcal{C}\langle A\rangle by w𝒮,(1)=1w^{\mathcal{S},\bullet}_{\hbar}(1)=1 and

w𝒮,(u1ur)=i=0ru1uiψ(ui+1ur)\displaystyle w^{\mathcal{S},\bullet}_{\hbar}(u_{1}\cdots u_{r})=\sum_{i=0}^{r}u_{1}\cdots u_{i}\bullet_{\hbar}\psi^{\bullet}(u_{i+1}\cdots u_{r})

for r1r\geq 1 and u1,,urAu_{1},\ldots,u_{r}\in A.

Now we define the 𝒞\mathcal{C}-linear map Zq,M𝒮,:𝒞AZ^{\mathcal{S},\bullet}_{q,M}:\mathcal{C}\langle A\rangle\to\mathbb{C} for {,sh}\bullet\in\{\ast,\mathcyr{sh}\} and M1M\geq 1 by Zq,M𝒮,=Zq,Mw𝒮,Z^{\mathcal{S},\bullet}_{q,M}=Z_{q,M}\circ w^{\mathcal{S},\bullet}_{\hbar}. We call the value Zq,M𝒮,(w)Z^{\mathcal{S},\bullet}_{q,M}(w) with w𝒞Aw\in\mathcal{C}\langle A\rangle a qq-analogue of truncated symmetric multiple zeta value.

From Proposition 3.2 and Proposition 3.3, we see that

(4.1) Zq,M𝒮,(u1ur)\displaystyle Z_{q,M}^{\mathcal{S},\ast}(u_{1}\cdots u_{r}) =i=0r0<m1<<mi<M0<mr<<mi+1<Mj=1iFq(mj;uj)j=i+1rFq(mj;ψ(uj)),\displaystyle=\sum_{i=0}^{r}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}<M\\ 0<m_{r}<\cdots<m_{i+1}<M\end{subarray}}\prod_{j=1}^{i}F_{q}(m_{j};u_{j})\prod_{j=i+1}^{r}F_{q}(m_{j};\psi^{\ast}(u_{j})),
(4.2) Zq,M𝒮,sh(u1ur)\displaystyle Z_{q,M}^{\mathcal{S},\mathcyr{sh}}(u_{1}\cdots u_{r}) =i=0r0<m1<<mi0<mr<<mi+1mi+mi+1<Mj=1iFq(mj;uj)j=i+1rFq(mj;ψsh(uj))\displaystyle=\sum_{i=0}^{r}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}\\ 0<m_{r}<\cdots<m_{i+1}\\ m_{i}+m_{i+1}<M\end{subarray}}\prod_{j=1}^{i}F_{q}(m_{j};u_{j})\prod_{j=i+1}^{r}F_{q}(m_{j};\psi^{\mathcyr{sh}}(u_{j}))

for u1,,urAu_{1},\ldots,u_{r}\in A.

We prove the double shuffle relation of the qq-analogue of truncated SMZVs. The proof is similar to that of the truncated SMZVs in [6].

Proposition 4.1.

For any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and M1M\geq 1, it holds that

(4.3) Zq,M𝒮,(ww)=Zq,M𝒮,(w)Zq,M𝒮,(w).\displaystyle Z^{\mathcal{S},\ast}_{q,M}(w\ast_{\hbar}w^{\prime})=Z^{\mathcal{S},\ast}_{q,M}(w)Z^{\mathcal{S},\ast}_{q,M}(w^{\prime}).
Proof.

Note that ψ\psi^{\ast} is an anti-automorphism. We extend the 𝒞\mathcal{C}-linear map Fq(m;):𝔷F_{q}(m;\cdot):\mathfrak{z}\to\mathbb{C} for m<0m<0 by Fq(m;u)=Fq(m;ψ(u))F_{q}(m;u)=F_{q}(-m;\psi^{\ast}(u)) for u𝔷u\in\mathfrak{z}. Then, from (4.1), we see that

(4.4) Zq,M𝒮,(u1ur)=m1mr0<|m1|,,|mr|<Mi=1rFq(mi;ui)\displaystyle Z^{\mathcal{S},\ast}_{q,M}(u_{1}\cdots u_{r})=\sum_{\begin{subarray}{c}m_{1}\prec\cdots\prec m_{r}\\ 0<|m_{1}|,\ldots,|m_{r}|<M\end{subarray}}\prod_{i=1}^{r}F_{q}(m_{i};u_{i})

for u1,,urAu_{1},\ldots,u_{r}\in A, where \prec is Kontsevich’s order on the set ({0}){=}(\mathbb{Z}\setminus\{0\})\sqcup\{\infty=-\infty\} defined by 12=211\prec 2\prec\cdots\prec\infty=-\infty\prec\cdots\prec-2\prec-1.

From the definition of \circ_{\hbar} and ψ\psi^{\ast}, we see that

ψ(uv)=ψ(u)ψ(v)\displaystyle\psi^{\ast}(u\circ_{\hbar}v)=\psi^{\ast}(u)\circ_{\hbar}\psi^{\ast}(v)

for any u,vAu,v\in A. Hence the relation (3.5) holds for any non-zero integer mm and u,vAu,v\in A, and we obtain (4.3) by using the expression (4.4) in the same way as the proof for the harmonic product relation of the truncated MZVs ζM(𝐤)\zeta_{M}(\mathbf{k}). ∎

Proposition 4.2.

For any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and M1M\geq 1, it holds that

(4.5) Zq,M𝒮,sh(wshw)=Zq,M𝒮,sh(wψsh(w)).\displaystyle Z^{\mathcal{S},\mathcyr{sh}}_{q,M}(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})=Z^{\mathcal{S},\mathcyr{sh}}_{q,M}(w\,\psi^{\mathcyr{sh}}(w^{\prime})).
Proof.

We extend the 𝒞\mathcal{C}-linear map Fq(m;):𝔷F_{q}(m;\cdot):\mathfrak{z}\to\mathbb{C} for m<0m<0 by Fq(m;u)=Fq(m;ψsh(u))F_{q}(m;u)=F_{q}(-m;\psi^{\mathcyr{sh}}(u)) for uAu\in A. Then we see that (3.3) holds for any nonzero integer mm and k1k\geq 1.

Let Tq,M:𝒞A×𝒞A×𝒞AT_{q,M}:\mathcal{C}\langle A\rangle\times\mathcal{C}\langle A\rangle\times\mathcal{C}\langle A\rangle\to\mathbb{C} be the 𝒞\mathcal{C}-trilinear map defined by Tq,M(1,1,1)=1T_{q,M}(1,1,1)=1 and

Tq,M(u1ur1,v1vr2,w1wr31)\displaystyle T_{q,M}(u_{1}\cdots u_{r_{1}},v_{1}\cdots v_{r_{2}},w_{1}\cdots w_{r_{3}-1})
=t=13j=1rt(li(s))Dj(t)i=1r1Fq(l1(1)++li(1);ui)i=1r2Fq(l1(2)++li(2);vi)\displaystyle=\sum_{t=1}^{3}\sum_{j=1}^{r_{t}}\sum_{(l_{i}^{(s)})\in D_{j}^{(t)}}\prod_{i=1}^{r_{1}}F_{q}(l_{1}^{(1)}+\cdots+l_{i}^{(1)};u_{i})\prod_{i=1}^{r_{2}}F_{q}(l_{1}^{(2)}+\cdots+l_{i}^{(2)};v_{i})
×i=1r31Fq(|l(1)|+|l(2)|+l1(3)++li(3);wi),\displaystyle\qquad\qquad\qquad\qquad{}\times\prod_{i=1}^{r_{3}-1}F_{q}(|l^{(1)}|+|l^{(2)}|+l_{1}^{(3)}+\cdots+l_{i}^{(3)};w_{i}),

where r1,r20,r31r_{1},r_{2}\geq 0,r_{3}\geq 1, u1,,ur1,v1,,vr2,w1,,wr31Au_{1},\ldots,u_{r_{1}},v_{1},\ldots,v_{r_{2}},w_{1},\ldots,w_{r_{3}-1}\in A and |l(s)|=j=1rslj(s)(s=1,2)|l^{(s)}|=\sum_{j=1}^{r_{s}}l_{j}^{(s)}\,(s=1,2). The summation region Dj(t)D_{j}^{(t)} is the subset of r1+r2+r3\mathbb{Z}^{r_{1}+r_{2}+r_{3}} consisting of tuples (li(s))1s31irs(l_{i}^{(s)})_{\begin{subarray}{c}1\leq s\leq 3\\ 1\leq i\leq r_{s}\end{subarray}} satisfying the following conditions:

s=13i=1rsli(s)=0,M<lj(t)<0,li(s)>0for any(s,i)(t,j).\displaystyle\sum_{s=1}^{3}\sum_{i=1}^{r_{s}}l_{i}^{(s)}=0,\quad-M<l_{j}^{(t)}<0,\quad l_{i}^{(s)}>0\,\,\hbox{for any}\,(s,i)\not=(t,j).

Then we see that Tq,M(u,v,1)=Zq,M𝒮,sh(uψsh(v))T_{q,M}(u,v,1)=Z_{q,M}^{\mathcal{S},\mathcyr{sh}}(u\,\psi^{\mathcyr{sh}}(v)) for u,v𝒞Au,v\in\mathcal{C}\langle A\rangle. Hence it suffices to show that

(4.6) Tq,M(u,v,w)=Zq,M𝒮,sh((ushv)w)\displaystyle T_{q,M}(u,v,w)=Z_{q,M}^{\mathcal{S},\mathcyr{sh}}((u\,\mathcyr{sh}_{\hbar}\,v)w)

for any u,v,w𝒞Au,v,w\in\mathcal{C}\langle A\rangle.

From (4.2) and the definition of Tq,MT_{q,M}, we see that (4.6) holds if u=1u=1 or v=1v=1. Set u=u1ur1u=u_{1}\cdots u_{r_{1}} and v=v1vr2v=v_{1}\cdots v_{r_{2}}. We prove (4.6) by induction on r1+r2r_{1}+r_{2}. Since Tq,M(u,v,w)T_{q,M}(u,v,w) is symmetric with respect to uu and vv, it suffices to consider the two cases: (i) ur1=e1g1u_{r_{1}}=e_{1}-g_{1},   (ii) ur1=gk,vr2=glu_{r_{1}}=g_{k},\,v_{r_{2}}=g_{l} with k,l1k,l\geq 1. The case (i) follows from (3.7) and Fq(m;e1g1)=Fq(m+n;e1g1)=1qF_{q}(m;e_{1}-g_{1})=F_{q}(m+n;e_{1}-g_{1})=1-q for integers mm and nn satisfying m0m\not=0 and m+n0m+n\not=0. To show the case (ii), we set

Hq(m;X)=k1Xk1Fq(m;gk)=qm[m]qmX.\displaystyle H_{q}(m;X)=\sum_{k\geq 1}X^{k-1}F_{q}(m;g_{k})=\frac{q^{m}}{[m]-q^{m}X}.

It holds that

Hq(m;X)Hq(n;Y)\displaystyle H_{q}(m;X)H_{q}(n;Y)
={(1+(1q)X)Hq(m;X)+(1+(1q)Y)Hq(n;Y)+1q}\displaystyle=\left\{(1+(1-q)X)H_{q}(m;X)+(1+(1-q)Y)H_{q}(n;Y)+1-q\right\}
×Hq(m+n;X+Y+(1q)XY)\displaystyle\times H_{q}(m+n;X+Y+(1-q)XY)

for non-zero integers mm and nn satisfying m+n0m+n\not=0. From the above relation and Lemma 3.10, we see that (4.6) holds in the case (ii) under the induction hypothesis. ∎

5. A qq-analogue of SMZV

In this section, we define a qq-analogue of the SMZV. To this aim, we use the map Zq𝒮,({,sh})Z_{q}^{\mathcal{S},\bullet}\,(\bullet\in\{\ast,\mathcyr{sh}\}), which corresponds to ζ𝒮,\zeta^{\mathcal{S},\bullet}, defined as follows.

First we define the map Zq𝒮,Z_{q}^{\mathcal{S},\ast}.

Proposition 5.1.

For any index 𝐤\mathbf{k}, the element w𝒮,(g𝐤)w_{\hbar}^{\mathcal{S},\ast}(g_{\mathbf{k}}) belongs to the \mathbb{Z}-submodule 𝐥I0g𝐥\oplus_{\mathbf{l}\in I_{0}}\mathbb{Z}\,g_{\mathbf{l}} of 0\mathfrak{H}^{0}.

Proof.

From the definition of the harmonic product \ast_{\hbar}, we see that g𝐤g𝐥=𝐦d𝐤,𝐥,𝐦g𝐦g_{\mathbf{k}}\ast_{\hbar}g_{\mathbf{l}}=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\ast,\mathbf{m}}g_{\mathbf{m}}, where d𝐤,𝐥,𝐦d_{\mathbf{k},\mathbf{l}}^{\ast,\mathbf{m}} is given by (2.4). Hence the proof of Proposition 5.4 in [6] with =\bullet=\ast and n=0n=0 works also for our map w𝒮,w_{\hbar}^{\mathcal{S},\ast}. ∎

Definition 5.2.

Set

𝔤=𝐤I𝒞g𝐤,\displaystyle\mathfrak{g}=\bigoplus_{\mathbf{k}\in I}\mathcal{C}g_{\mathbf{k}},

which is a 𝒞\mathcal{C}-submodule of 0^\widehat{\mathfrak{H}^{0}}. We define the 𝒞\mathcal{C}-linear map Zq𝒮,:𝔤𝒵qZ_{q}^{\mathcal{S},\ast}:\mathfrak{g}\rightarrow\mathcal{Z}_{q} by

Zq𝒮,(g𝐤)=limMZq,M𝒮,(g𝐤)=Zq(w𝒮,(g𝐤))\displaystyle Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})=\lim_{M\to\infty}Z_{q,M}^{\mathcal{S},\ast}(g_{\mathbf{k}})=Z_{q}(w_{\hbar}^{\mathcal{S},\ast}(g_{\mathbf{k}}))

for an index 𝐤\mathbf{k}. More explicitly, we have

Zq𝒮,(g𝐤)=i=0r(1)ki+1++kr0<m1<<mi0<mr<<mi+1qk1m1++krmr[m1]k1[mr]kr\displaystyle Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}\\ 0<m_{r}<\cdots<m_{i+1}\end{subarray}}\frac{q^{k_{1}m_{1}+\cdots+k_{r}m_{r}}}{[m_{1}]^{k_{1}}\cdots[m_{r}]^{k_{r}}}

for a non-empty index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}).

We want to define the map Zq𝒮,shZ_{q}^{\mathcal{S},\mathcyr{sh}} similarly by taking the limit of Zq,M𝒮,shZ_{q,M}^{\mathcal{S},\mathcyr{sh}} as MM\to\infty. For that purpose, however, we should determine the domain carefully because it is not closed under the concatenation product unlike the 𝒞\mathcal{C}-module 𝔤\mathfrak{g} in Definition 5.2 as seen by the following example.

Example 5.3.

We have w𝒮,sh(e1)=w𝒮,sh(g1)=e1g1w_{\hbar}^{\mathcal{S},\mathcyr{sh}}(e_{1})=-w_{\hbar}^{\mathcal{S},\mathcyr{sh}}(g_{1})=e_{1}-g_{1} and Zq,M𝒮,sh(e1)=Zq,M𝒮,sh(g1)=(1q)(M1)Z_{q,M}^{\mathcal{S},\mathcyr{sh}}(e_{1})=-Z_{q,M}^{\mathcal{S},\mathcyr{sh}}(g_{1})=(1-q)(M-1). Hence a 𝒞\mathcal{C}-linear combination of e1e_{1} and g1g_{1} whose image by Zq,M𝒮,shZ_{q,M}^{\mathcal{S},\mathcyr{sh}} converges in the limit as MM\to\infty should be proportional to e1+g1e_{1}+g_{1}. However, we see that

Zq,M𝒮,sh((e1+g1)2)=ZM((e1g1)2)=(1q)2(M12)(M).\displaystyle Z_{q,M}^{\mathcal{S},\mathcyr{sh}}((e_{1}+g_{1})^{2})=Z_{M}((e_{1}-g_{1})^{2})=(1-q)^{2}\binom{M-1}{2}\to\infty\qquad(M\to\infty).

Therefore, the concatenation product (e1+g1)2(e_{1}+g_{1})^{2} of e1+g1e_{1}+g_{1} does not belong to the domain of the map limMZq,M𝒮,sh\lim_{M\to\infty}Z_{q,M}^{\mathcal{S},\mathcyr{sh}}. Here we note that, because w𝒮,sh(e1g1+g1e1)=(e1g1)2w_{\hbar}^{\mathcal{S},\mathcyr{sh}}(e_{1}g_{1}+g_{1}e_{1})=-(e_{1}-g_{1})^{2}, it holds that Zq,M𝒮,sh(e12+2e1g1+2g1e1+g12)=0Z_{q,M}^{\mathcal{S},\mathcyr{sh}}(e_{1}^{2}+2e_{1}g_{1}+2g_{1}e_{1}+g_{1}^{2})=0, which clearly converges as MM\to\infty.

To describe the domain of the map w𝒮,shw_{\hbar}^{\mathcal{S},\mathcyr{sh}}, we introduce the element E1m(m0)E_{1^{m}}\,(m\geq 0) of 𝒞A\mathcal{C}\langle A\rangle defined by

E1m=1(m+1)!j=0mg1shjshe1sh(mj),\displaystyle E_{1^{m}}=\frac{1}{(m+1)!}\sum_{j=0}^{m}g_{1}^{\mathcyr{sh}_{\hbar}j}\mathcyr{sh}_{\hbar}e_{1}^{\mathcyr{sh}_{\hbar}(m-j)},

where ush0=1u^{\mathcyr{sh}_{\hbar}0}=1 and ushj=ushush(j1)u^{\mathcyr{sh}_{\hbar}j}=u\mathcyr{sh}_{\hbar}u^{\mathcyr{sh}_{\hbar}(j-1)} for uu\in\mathfrak{H} and j1j\geq 1. For example, we have

E10=1,E11=12(e1+g1),E12=16(e12+2e1g1+2g1e1+g12).\displaystyle E_{1^{0}}=1,\qquad E_{1^{1}}=\frac{1}{2}(e_{1}+g_{1}),\qquad E_{1^{2}}=\frac{1}{6}(e_{1}^{2}+2e_{1}g_{1}+2g_{1}e_{1}+g_{1}^{2}).

Let 𝐤\mathbf{k} be a non-empty index. It can be written uniquely in the form

(5.1) 𝐤=(1,,1s0,t1+2,1,,1s1,,tr+2,1,,1sr)\displaystyle\mathbf{k}=(\underbrace{1,\ldots,1}_{s_{0}},t_{1}+2,\underbrace{1,\ldots,1}_{s_{1}},\ldots,t_{r}+2,\underbrace{1,\ldots,1}_{s_{r}})

with r0r\geq 0 and s0,,sr,t1,,tr0s_{0},\ldots,s_{r},t_{1},\ldots,t_{r}\geq 0, where the right hand side reads (1,,1s0)(\underbrace{1,\ldots,1}_{s_{0}}) if r=0r=0. Then we set

E𝐤=E1s0et1+2E1s1etr+2E1sr.\displaystyle E_{\mathbf{k}}=E_{1^{s_{0}}}e_{t_{1}+2}E_{1^{s_{1}}}\cdots e_{t_{r}+2}E_{1^{s_{r}}}.

For the empty index, we set E=1E_{\varnothing}=1.

Proposition 5.4.

For any index 𝐤\mathbf{k}, the element w𝒮,sh(E𝐤)w_{\hbar}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}) belongs to 0\mathfrak{H}^{0}.

See Appendix B.2 and Appendix B.3 for the proof.

Definition 5.5.

Set

𝔢=𝐤I𝒞E𝐤.\displaystyle\mathfrak{e}=\bigoplus_{\mathbf{k}\in I}\mathcal{C}E_{\mathbf{k}}.

We define the 𝒞\mathcal{C}-linear map Zq𝒮,sh:𝔢𝒵qZ_{q}^{\mathcal{S},\mathcyr{sh}}:\mathfrak{e}\rightarrow\mathcal{Z}_{q} by

Zq𝒮,sh(E𝐤)=limMZq,M𝒮,sh(E𝐤)=Zq(w𝒮,sh(E𝐤)).\displaystyle Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}})=\lim_{M\to\infty}Z_{q,M}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}})=Z_{q}(w_{\hbar}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}})).

for an index 𝐤\mathbf{k}.

Example 5.6.

If k2k\geq 2, it holds that Fq(m;ψsh(ek))=Fq(m;ek)=(1)kqm/[m]kF_{q}(m;\psi^{\mathcyr{sh}}(e_{k}))=F_{q}(-m;e_{k})=(-1)^{k}q^{m}/[m]^{k} for m1m\geq 1. Hence, for a non-empty index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) whose all components are larger than one, we have

Zq𝒮,sh(E𝐤)\displaystyle Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}) =Zq𝒮,sh(ek1ekr)\displaystyle=Z_{q}^{\mathcal{S},\mathcyr{sh}}(e_{k_{1}}\cdots e_{k_{r}})
=i=0r(1)ki+1++kr0<m1<<mi0<mr<<mi+1q(k11)m1++(ki1)mi[m1]k1[mi]kiqmi+1++mr[mi+1]ki+1[mr]kr.\displaystyle=\sum_{i=0}^{r}(-1)^{k_{i+1}+\cdots+k_{r}}\sum_{\begin{subarray}{c}0<m_{1}<\cdots<m_{i}\\ 0<m_{r}<\cdots<m_{i+1}\end{subarray}}\frac{q^{(k_{1}-1)m_{1}+\cdots+(k_{i}-1)m_{i}}}{[m_{1}]^{k_{1}}\cdots[m_{i}]^{k_{i}}}\frac{q^{m_{i+1}+\cdots+m_{r}}}{[m_{i+1}]^{k_{i+1}}\cdots[m_{r}]^{k_{r}}}.
Example 5.7.

From Proposition B.4 and Lemma B.7, we see that w𝒮,sh(E1m)=0w_{\hbar}^{\mathcal{S},\mathcyr{sh}}(E_{1^{m}})=0 for m1m\geq 1. Hence we have Zq𝒮,sh(E1,,1)=Zq𝒮,sh(E1m)=0Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{1,\ldots,1})=Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{1^{m}})=0.

Using the two maps Zq𝒮,Z_{q}^{\mathcal{S},\ast} and Zq𝒮,shZ_{q}^{\mathcal{S},\mathcyr{sh}}, we define a qq-analogue of the SMZV. Recall that we need the relation (2.3) to define the SMZV. We show the corresponding relation in the qq-analogue case.

Set

𝒩q=Zq(𝔫).\displaystyle\mathcal{N}_{q}=Z_{q}(\mathfrak{n}).

From Proposition 3.7, we see that 𝒩q\mathcal{N}_{q} is an ideal of 𝒵q^\widehat{\mathcal{Z}_{q}}. Note that (1q)𝒵q𝒩q(1-q)\mathcal{Z}_{q}\subset\mathcal{N}_{q} since 0𝔫\hbar\,\mathfrak{H}^{0}\subset\mathfrak{n}.

For {,sh}\bullet\in\{\ast,\mathcyr{sh}\} and f(X)𝒞A[[X]]f(X)\in\mathcal{C}\langle A\rangle[[X]] satisfying f(0)=0f(0)=0, we define the exponential exp(f(X))\exp_{\bullet_{\hbar}}(f(X)) with respect to the product \bullet_{\hbar} by

exp(f(X))=1+n11n!f(X)f(X)n times.\displaystyle\exp_{\bullet_{\hbar}}{\left(f(X)\right)}=1+\sum_{n\geq 1}\frac{1}{n!}\underbrace{f(X)\bullet_{\hbar}\cdots\bullet_{\hbar}f(X)}_{\hbox{\tiny$n$ times}}.
Theorem 5.8.

For any admissible index 𝐤\mathbf{k}, it holds that

(5.2) Zq(g𝐤11g1X)exp(n2(1)n1nZq(gn)Xn)Zq(E𝐤expsh(g1X))\displaystyle Z_{q}\left(g_{\mathbf{k}}\,\frac{1}{1-g_{1}X}\right)\equiv\exp{\left(\sum_{n\geq 2}\frac{(-1)^{n-1}}{n}Z_{q}(g_{n})X^{n}\right)}Z_{q}\left(E_{\mathbf{k}}\exp_{\mathcyr{sh}_{\hbar}}(g_{1}X)\right)

in 𝒵q^[[X]]\widehat{\mathcal{Z}_{q}}[[X]] modulo 𝒩q[[X]]\mathcal{N}_{q}[[X]].

Proof.

We start from the following equality. See Appendix B.4 for the proof.

Lemma 5.9.

Set

(5.3) R(X)=ebX1b=n=1Xnn!(e1g1)n1.\displaystyle R(X)=\frac{e^{\hbar bX}-1}{\hbar b}=\sum_{n=1}^{\infty}\frac{X^{n}}{n!}(e_{1}-g_{1})^{n-1}.

For any admissible index 𝐤\mathbf{k}, it holds that

(5.4) 11R(X)g1shg𝐤11g1X11g1XshE𝐤11R(X)g1\displaystyle\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,g_{\mathbf{k}}\frac{1}{1-g_{1}X}\equiv\frac{1}{1-g_{1}X}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{k}}\frac{1}{1-R(X)g_{1}}

modulo the 𝒞\mathcal{C}-submodule 𝔫[[X]]\mathfrak{n}[[X]] of [[X]]\mathfrak{H}[[X]].

We calculate the image by the map ZqZ_{q} of the both hand sides of (5.4) using Proposition 3.3 and (B.2). Then we obtain

exp(Zq(g1)X)Zq(g𝐤11g1X)Zq(11g1X)Zq(E𝐤expsh(g1X))\displaystyle\exp{(Z_{q}(g_{1})X)}Z_{q}\left(g_{\mathbf{k}}\,\frac{1}{1-g_{1}X}\right)\equiv Z_{q}\left(\frac{1}{1-g_{1}X}\right)Z_{q}\left(E_{\mathbf{k}}\exp_{\mathcyr{sh}_{\hbar}}(g_{1}X)\right)

modulo 𝒩q[[X]]\mathcal{N}_{q}[[X]]. It holds that

(5.5) 11gkX=exp(n1(1)n1ngnkXn)\displaystyle\frac{1}{1-g_{k}X}=\exp_{\ast_{\hbar}}{\left(\sum_{n\geq 1}\frac{(-1)^{n-1}}{n}g_{nk}X^{n}\right)}

for any k1k\geq 1 (see [3, Corollary 1]). Hence, from Proposition 3.2, we have

Zq(11g1X)=exp(n1(1)n1nZq(gn)Xn).\displaystyle Z_{q}\left(\frac{1}{1-g_{1}X}\right)=\exp{\left(\sum_{n\geq 1}\frac{(-1)^{n-1}}{n}Z_{q}(g_{n})X^{n}\right)}.

It implies (5.2) since 𝒩q\mathcal{N}_{q} is an ideal of 𝒵q^\widehat{\mathcal{Z}_{q}}. ∎

We denote by 𝒫q\mathcal{P}_{q} the ideal of 𝒵q\mathcal{Z}_{q} generated by the set {Zq(g2k)k1}\{Z_{q}(g_{2k})\mid k\geq 1\}.

Corollary 5.10.

For any index 𝐤\mathbf{k}, the difference Zq𝒮,(g𝐤)Zq𝒮,sh(E𝐤)Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})-Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}) belongs to the ideal 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q}.

Proof.

From the definition of w𝒮,({,sh})w_{\hbar}^{\mathcal{S},\bullet}\,(\bullet\in\{\ast,\mathcyr{sh}\}), Proposition B.4, Corollary B.6 and Lemma B.7, we see that the difference Zq𝒮,(g𝐤)Zq𝒮,sh(E𝐤)Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})-Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}) is a signed sum of

(5.6) s=0m(1)ms(Zq(g𝐦g1s)Zq(g𝐦g1ms)Zq(E𝐦g1shss!)Zq(E𝐦g1sh(ms)(ms)!))\displaystyle\sum_{s=0}^{m}(-1)^{m-s}\left(Z_{q}(g_{\mathbf{m}}g_{1}^{s})Z_{q}(g_{\mathbf{m}^{\prime}}g_{1}^{m-s})-Z_{q}(E_{\mathbf{m}}\frac{g_{1}^{\mathcyr{sh}_{\hbar}s}}{s!})Z_{q}(E_{\mathbf{m}^{\prime}}\frac{g_{1}^{\mathcyr{sh}_{\hbar}(m-s)}}{(m-s)!})\right)

with admissible indices 𝐦,𝐦\mathbf{m},\mathbf{m}^{\prime} and m0m\geq 0 modulo Zq(0^)Z_{q}(\hbar\widehat{\mathfrak{H}^{0}}). We denote (5.6) by JmJ_{m} and calculate the generating function J(X)=m0JmXmJ(X)=\sum_{m\geq 0}J_{m}X^{m}. From Theorem 5.8, we see that

J(X)\displaystyle J(X) =Zq(g𝐦11g1X)Zq(g𝐦11+g1X)Zq(E𝐦expsh(g1X))Zq(E𝐦expsh(g1X))\displaystyle=Z_{q}(g_{\mathbf{m}}\frac{1}{1-g_{1}X})Z_{q}(g_{\mathbf{m}^{\prime}}\frac{1}{1+g_{1}X})-Z_{q}(E_{\mathbf{m}}\exp_{\mathcyr{sh}_{\hbar}}{(g_{1}X)})Z_{q}(E_{\mathbf{m}^{\prime}}\exp_{\mathcyr{sh}_{\hbar}}{(-g_{1}X)})
{exp(k1Zq(g2k)kX2k)1}Zq(g𝐦11g1X)Zq(g𝐦11+g1X)\displaystyle\equiv\left\{\exp{\left(\sum_{k\geq 1}\frac{Z_{q}(g_{2k})}{k}X^{2k}\right)}-1\right\}Z_{q}(g_{\mathbf{m}}\frac{1}{1-g_{1}X})Z_{q}(g_{\mathbf{m}^{\prime}}\frac{1}{1+g_{1}X})

modulo 𝒩q[[X]]\mathcal{N}_{q}[[X]]. The right hand side belongs to 𝒫q[[X]]\mathcal{P}_{q}[[X]]. ∎

Now we are in a position to define a qq-analogue of the SMZV.

Definition 5.11.

For an index 𝐤\mathbf{k}, we define a qq-analogue of SMZV (qqSMZV) ζq𝒮(𝐤)\zeta_{q}^{\mathcal{S}}(\mathbf{k}) as an element of the quotient 𝒵q/(𝒩q+𝒫q)\mathcal{Z}_{q}/(\mathcal{N}_{q}+\mathcal{P}_{q}) by

ζq𝒮(𝐤)=Zq𝒮,(g𝐤)=Zq𝒮,sh(E𝐤)\displaystyle\zeta_{q}^{\mathcal{S}}(\mathbf{k})=Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})=Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}})

modulo 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q}.

Example 5.12.

Let kk be a positive integer. From the definition of w𝒮,w_{\hbar}^{\mathcal{S},\ast} and (5.5), we see that

s0Xsw𝒮,(gks)=11gkX11(1)kgkX=exp(n1(1)n1n(1+(1)kn)gknXn).\displaystyle\sum_{s\geq 0}X^{s}w_{\hbar}^{\mathcal{S},\ast}(g_{k}^{s})=\frac{1}{1-g_{k}X}\ast_{\hbar}\frac{1}{1-(-1)^{k}g_{k}X}=\exp_{\ast_{\hbar}}{\left(\sum_{n\geq 1}\frac{(-1)^{n-1}}{n}(1+(-1)^{kn})g_{kn}X^{n}\right)}.

Hence Proposition 3.2 implies that

s0XsZq𝒮,(gks)=exp(n1(1)n1n(1+(1)kn)Zq(gkn)Xn).\displaystyle\sum_{s\geq 0}X^{s}Z_{q}^{\mathcal{S},\ast}(g_{k}^{s})=\exp{\left(\sum_{n\geq 1}\frac{(-1)^{n-1}}{n}(1+(-1)^{kn})Z_{q}(g_{kn})X^{n}\right)}.

If kk is odd, then 1+(1)kn=01+(-1)^{kn}=0 unless nn is even. Therefore, the right hand side belongs to 1+𝒫q[[X]]1+\mathcal{P}_{q}[[X]] and we have

(5.7) ζq𝒮(k,,kr)=0\displaystyle\zeta_{q}^{\mathcal{S}}(\underbrace{k,\ldots,k}_{r})=0

for any k,r1k,r\geq 1. In particular, the qqSMZV of depth one is always equal to zero.

Example 5.13.

We consider the qqSMZV of depth two. Set 𝐤=(k1,k2)\mathbf{k}=(k_{1},k_{2}). If k1k_{1} and k2k_{2} are even, then w𝒮,(g𝐤)=2gk1gk2gk1+k2w_{\hbar}^{\mathcal{S},\ast}(g_{\mathbf{k}})=2g_{k_{1}}\ast_{\hbar}g_{k_{2}}-g_{k_{1}+k_{2}}. If k1k_{1} and k2k_{2} are odd, we have w𝒮,(g𝐤)=gk1+k2w_{\hbar}^{\mathcal{S},\ast}(g_{\mathbf{k}})=-g_{k_{1}+k_{2}}. In both cases we see that Zq(w𝒮,(g𝐤))Z_{q}(w_{\hbar}^{\mathcal{S},\ast}(g_{\mathbf{k}})) belongs to 𝒫q\mathcal{P}_{q} from Proposition 3.2. Therefore ζq𝒮(k1,k2)=0\zeta_{q}^{\mathcal{S}}(k_{1},k_{2})=0 if k1+k2k_{1}+k_{2} is even.

We consider the case where k1+k2k_{1}+k_{2} is odd. To this aim we calculate the qqMZV Zq(gk1gk2)Z_{q}(g_{k_{1}}g_{k_{2}}) whose weight is odd modulo 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q}. The calculation is similar to that for MZV in [1, 10].

Suppose that kk is odd and k3k\geq 3. For 1m<k1\leq m<k, we have

gmgkm=gmgkm+gkmgm+gk.\displaystyle g_{m}\ast_{\hbar}g_{k-m}=g_{m}g_{k-m}+g_{k-m}g_{m}+g_{k}.

From Lemma 3.10, we also see that

m,l1Xm1Yl1gmshgl\displaystyle\sum_{m,l\geq 1}X^{m-1}Y^{l-1}g_{m}\,\mathcyr{sh}_{\hbar}\,g_{l} =(1+X)m,l1Xk1(X+Y+XY)l1gmgl\displaystyle=(1+\hbar X)\sum_{m,l\geq 1}X^{k-1}(X+Y+\hbar XY)^{l-1}g_{m}g_{l}
+(1+Y)m,l1Yk1(X+Y+XY)l1gmgl\displaystyle+(1+\hbar Y)\sum_{m,l\geq 1}Y^{k-1}(X+Y+\hbar XY)^{l-1}g_{m}g_{l}
+l1(X+Y+XY)l1(e1g1)gl.\displaystyle+\sum_{l\geq 1}(X+Y+\hbar XY)^{l-1}(e_{1}-g_{1})g_{l}.

Since k3k\geq 3 and (e1g1)gl(e_{1}-g_{1})g_{l} belongs to 𝔫0\mathfrak{n}_{0} if l2l\geq 2, we have

gmshgkmj1((kj1km1)+(kj1m1))gjgkj\displaystyle g_{m}\,\mathcyr{sh}_{\hbar}\,g_{k-m}\equiv\sum_{j\geq 1}\left(\binom{k-j-1}{k-m-1}+\binom{k-j-1}{m-1}\right)g_{j}g_{k-j}

modulo 𝔫\mathfrak{n}. Note that mm or kmk-m is even. Hence, from Proposition 3.2 and Proposition 3.3, we see that

(5.8) Zq(gmgkm+gkmgm+gk)0,\displaystyle Z_{q}(g_{m}g_{k-m}+g_{k-m}g_{m}+g_{k})\equiv 0,
(5.9) j1((kj1km1)+(kj1m1))Zq(gjgkj)0\displaystyle\sum_{j\geq 1}\left(\binom{k-j-1}{k-m-1}+\binom{k-j-1}{m-1}\right)Z_{q}(g_{j}g_{k-j})\equiv 0

modulo 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q}. Set

𝒟(X,Y)=m=1k1Xm1Ykm1Zq(gmgkm),𝒵(X,Y)=Xk1Yk1XYZq(gk).\displaystyle\mathcal{D}(X,Y)=\sum_{m=1}^{k-1}X^{m-1}Y^{k-m-1}Z_{q}(g_{m}g_{k-m}),\qquad\mathcal{Z}(X,Y)=\frac{X^{k-1}-Y^{k-1}}{X-Y}Z_{q}(g_{k}).

Then (5.8) and (5.9) imply that

𝒟(X,Y)+𝒟(Y,X)+𝒵(X,Y)0,𝒟(X,X+Y)+𝒟(Y,X+Y)0\displaystyle\mathcal{D}(X,Y)+\mathcal{D}(Y,X)+\mathcal{Z}(X,Y)\equiv 0,\qquad\mathcal{D}(X,X+Y)+\mathcal{D}(Y,X+Y)\equiv 0

modulo (𝒩q+𝒫q)[X,Y](\mathcal{N}_{q}+\mathcal{P}_{q})[X,Y], respectively. Using these relations and 𝒟(X,Y)=𝒟(X,Y)\mathcal{D}(-X,-Y)=-\mathcal{D}(X,Y), we obtain

𝒟(X,Y)12(𝒵(X,Y)+𝒵(XY,X)+𝒵(Y,XY)).\displaystyle\mathcal{D}(X,Y)\equiv-\frac{1}{2}\left(\mathcal{Z}(X,Y)+\mathcal{Z}(X-Y,X)+\mathcal{Z}(-Y,X-Y)\right).

Hence we find that

(5.10) Zq(gmgkm)12(1+(1)m(km))Zq(gk)\displaystyle Z_{q}(g_{m}g_{k-m})\equiv-\frac{1}{2}\left(1+(-1)^{m}\binom{k}{m}\right)Z_{q}(g_{k})

modulo 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q} if 1m<k1\leq m<k and kk is odd.

Now suppose that k1,k21k_{1},k_{2}\geq 1 and k1+k2k_{1}+k_{2} is odd. Since k1k_{1} or k2k_{2} is even, we see that

Zq𝒮,(gk1gk2)\displaystyle Z_{q}^{\mathcal{S},\ast}(g_{k_{1}}g_{k_{2}}) =Zq(gk1gk2+(1)k2gk1gk2gk2gk1)\displaystyle=Z_{q}(g_{k_{1}}g_{k_{2}}+(-1)^{k_{2}}g_{k_{1}}\ast_{\hbar}g_{k_{2}}-g_{k_{2}}g_{k_{1}})
Zq(gk1gk2gk2gk1)\displaystyle\equiv Z_{q}(g_{k_{1}}g_{k_{2}}-g_{k_{2}}g_{k_{1}})
(1)k2(k1+k2k1)Zq(gk1+k2)\displaystyle\equiv(-1)^{k_{2}}\binom{k_{1}+k_{2}}{k_{1}}Z_{q}(g_{k_{1}+k_{2}})

modulo 𝒩q+𝒫q\mathcal{N}_{q}+\mathcal{P}_{q} by using Proposition 3.2 and (5.10).

From the above arguments we see that

ζq𝒮(k1,k2)=(1)k2(k1+k2k1)Zq(gk1+k2)\displaystyle\zeta_{q}^{\mathcal{S}}(k_{1},k_{2})=(-1)^{k_{2}}\binom{k_{1}+k_{2}}{k_{1}}Z_{q}(g_{k_{1}+k_{2}})

in the quotient 𝒵q/(𝒩q+𝒫q)\mathcal{Z}_{q}/(\mathcal{N}_{q}+\mathcal{P}_{q}) for any k1,k21k_{1},k_{2}\geq 1. It is a qq-analogue of the formula for the SMZV of depth two

ζ𝒮(k1,k2)=(1)k2(k1+k2k1)ζ(k1+k2)\displaystyle\zeta^{\mathcal{S}}(k_{1},k_{2})=(-1)^{k_{2}}\binom{k_{1}+k_{2}}{k_{1}}\zeta(k_{1}+k_{2})

modulo ζ(2)𝒵\zeta(2)\mathcal{Z} (see, e.g., [5]).

We check that our qqSMZV is really a qq-analogue of the SMZV.

Theorem 5.14.

For any index 𝐤\mathbf{k}, it holds that

limq10Zq𝒮,(g𝐤)=ζ𝒮,(𝐤),limq10Zq𝒮,sh(E𝐤)=ζ𝒮,sh(𝐤).\displaystyle\lim_{q\to 1-0}Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})=\zeta^{\mathcal{S},\ast}(\mathbf{k}),\qquad\lim_{q\to 1-0}Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}})=\zeta^{\mathcal{S},\mathcyr{sh}}(\mathbf{k}).
Proof.

We see that ι(ψ(w))=ψ(ι(w))\iota(\psi^{\bullet}(w))=\psi(\iota(w)) for any w𝒞Aw\in\mathcal{C}\langle A\rangle from the definition of ι,ψ\iota,\psi^{\bullet} and ψ\psi. Hence, Proposition 3.9 implies that ι(w𝒮,(w))=w𝒮,(ι(w))\iota(w_{\hbar}^{\mathcal{S},\bullet}(w))=w^{\mathcal{S},\bullet}(\iota(w)) for any w𝒞Aw\in\mathcal{C}\langle A\rangle and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}. From the definition of ι\iota, we have ι(g𝐤)=z𝐤\iota(g_{\mathbf{k}})=z_{\mathbf{k}} for any index 𝐤\mathbf{k}. Moreover, we have ι(ek)=zk\iota(e_{k})=z_{k} for k1k\geq 1 and

ι(E1m)=1(m+1)!j=0mι(g1)shjshι(e1)sh(mj)=z1shmm!=z1m\displaystyle\iota(E_{1^{m}})=\frac{1}{(m+1)!}\sum_{j=0}^{m}\iota(g_{1})^{\mathcyr{sh}j}\,\mathcyr{sh}\,\iota(e_{1})^{\mathcyr{sh}(m-j)}=\frac{z_{1}^{\mathcyr{sh}m}}{m!}=z_{1}^{m}

for m0m\geq 0. Hence ι(E𝐤)=z𝐤\iota(E_{\mathbf{k}})=z_{\mathbf{k}} for any index 𝐤\mathbf{k}. Now the desired formula follows from Corollary 3.5. ∎

The limit as q10q\to 1-0 of any element of 𝒩q\mathcal{N}_{q} is zero and that of 𝒫q\mathcal{P}_{q} is contained in ζ(2)𝒵\zeta(2)\mathcal{Z} because limq10Zq(g2k)=ζ(2k)ζ(2)k\lim_{q\to 1-0}Z_{q}(g_{2k})=\zeta(2k)\in\mathbb{Q}\,\zeta(2)^{k} for k1k\geq 1. Therefore, we have the well-defined map 𝒵q/(𝒩q+𝒫q)𝒵/ζ(2)𝒵\mathcal{Z}_{q}/(\mathcal{N}_{q}+\mathcal{P}_{q})\to\mathcal{Z}/\zeta(2)\mathcal{Z} which sends the equivalent class of f(q)𝒵qf(q)\in\mathcal{Z}_{q} to that of limq10f(q)\lim_{q\to 1-0}f(q). Theorem 5.14 implies that the map sends the qqSMZV ζq𝒮(𝐤)\zeta_{q}^{\mathcal{S}}(\mathbf{k}) to the SMZV ζ𝒮(𝐤)\zeta^{\mathcal{S}}(\mathbf{k}). In this sense we may regard ζq𝒮(𝐤)\zeta_{q}^{\mathcal{S}}(\mathbf{k}) as a qq-analogue of ζ𝒮(𝐤)\zeta^{\mathcal{S}}(\mathbf{k}).

6. Relations of the qq-analogue of symmetric multiple zeta value

6.1. Reversal relation

Theorem 6.1.

For any index 𝐤\mathbf{k}, we have

Zq𝒮,(g𝐤¯)=(1)wt(𝐤)Zq𝒮,(g𝐤),\displaystyle Z_{q}^{\mathcal{S},\ast}(g_{\overline{\mathbf{k}}})=(-1)^{\mathrm{wt}(\mathbf{k})}Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}}),

and

Zq𝒮,sh(E𝐤¯)(1)wt(𝐤)Zq𝒮,sh(E𝐤)\displaystyle Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\overline{\mathbf{k}}})\equiv(-1)^{\mathrm{wt}(\mathbf{k})}Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}})

modulo (1q)𝒵q(1-q)\mathcal{Z}_{q}. Therefore, for any index 𝐤\mathbf{k}, it holds that

ζq𝒮(𝐤¯)=(1)wt(𝐤)ζq𝒮(𝐤).\displaystyle\zeta_{q}^{\mathcal{S}}(\overline{\mathbf{k}})=(-1)^{\mathrm{wt}(\mathbf{k})}\zeta_{q}^{\mathcal{S}}(\mathbf{k}).
Proof.

Since ψ\psi^{\bullet} is an anti-involution, we see that w𝒮,(ψ(w))=w𝒮,(w)w_{\hbar}^{\mathcal{S},\bullet}(\psi^{\bullet}(w))=w_{\hbar}^{\mathcal{S},\bullet}(w) for any w𝒞Aw\in\mathcal{C}\langle A\rangle and {,sh}\bullet\in\{\ast,\mathcyr{sh}\}. From the definition of ψ\psi^{\ast}, we have ψ(g𝐤)=(1)wt(𝐤)g𝐤¯\psi^{\ast}(g_{\mathbf{k}})=(-1)^{\mathrm{wt}(\mathbf{k})}g_{\overline{\mathbf{k}}}. We also have ψsh(E𝐤)(1)wt(𝐤)E𝐤¯\psi^{\mathcyr{sh}}(E_{\mathbf{k}})\equiv(-1)^{\mathrm{wt}(\mathbf{k})}E_{\overline{\mathbf{k}}} modulo 𝔢\hbar\,\mathfrak{e} (see Corollary B.6). Thus we obtain the desired equalities. ∎

6.2. Double shuffle relation

From Proposition 4.1, we obtain the following relation of qqSMZVs.

Proposition 6.2.

For any index 𝐤\mathbf{k} and 𝐥\mathbf{l}, it holds that

Zq𝒮,(g𝐤g𝐥)=Zq𝒮,(g𝐤)Zq𝒮,(g𝐥).\displaystyle Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}}\ast_{\hbar}g_{\mathbf{l}})=Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{k}})Z_{q}^{\mathcal{S},\ast}(g_{\mathbf{l}}).

Next we consider the shuffle relation. Note that the product E𝐤shE𝐥E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}} does not necessarily belong to 𝔢\mathfrak{e} as follows.

Example 6.3.

We have

E1shE1=14(g1+e1)sh(g1+e1)=14(g12+3g1e1+3e1g1+e12),\displaystyle E_{1}\,\mathcyr{sh}_{\hbar}\,E_{1}=\frac{1}{4}(g_{1}+e_{1})\,\mathcyr{sh}_{\hbar}\,(g_{1}+e_{1})=\frac{1}{4}(g_{1}^{2}+3g_{1}e_{1}+3e_{1}g_{1}+e_{1}^{2}),

which is not a 𝒞\mathcal{C}-linear combination of E11=E12=(e12+2e1g1+2g1e1+g12)/6,E2=e2E_{11}=E_{1^{2}}=(e_{1}^{2}+2e_{1}g_{1}+2g_{1}e_{1}+g_{1}^{2})/6,E_{2}=e_{2} and E1=(e1+g1)/2E_{1}=(e_{1}+g_{1})/2.

However, we have the following proposition. We set

𝔢0=𝐤I0𝒞E𝐤.\displaystyle\mathfrak{e}^{0}=\sum_{\mathbf{k}\in I_{0}}\mathcal{C}E_{\mathbf{k}}.
Proposition 6.4.

Let 𝐤\mathbf{k} and 𝐥\mathbf{l} be an index. If at least one of 𝐤\mathbf{k} and 𝐥\mathbf{l} is admissible, then E𝐤shE𝐥E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}} belongs to 𝔢\mathfrak{e}. If both 𝐤\mathbf{k} and 𝐥\mathbf{l} are admissible, then E𝐤shE𝐥E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}} belongs to 𝔢0\mathfrak{e}^{0}.

See Appendix B.5 for the proof.

Proposition 6.5.

Let 𝐤\mathbf{k} and 𝐥\mathbf{l} be an index and suppose that at least one of them is admissible. Then it holds that

Zq𝒮,sh(E𝐤shE𝐥)(1)wt(𝐥)Zq𝒮,sh(E(𝐤,𝐥¯)),\displaystyle Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}\mathcyr{sh}_{\hbar}E_{\mathbf{l}})\equiv(-1)^{\mathrm{wt}(\mathbf{l})}Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{(\mathbf{k},\overline{\mathbf{l}})}),

modulo (1q)𝒵q(1-q)\mathcal{Z}_{q}, where (𝐤,𝐥¯)(\mathbf{k},\overline{\mathbf{l}}) is the concatenation of 𝐤\mathbf{k} and 𝐥¯\overline{\mathbf{l}}.

Proof.

Proposition 4.2 implies that Zq𝒮,sh(E𝐤shE𝐥)=Zq𝒮,sh(E𝐤ψsh(E𝐥))Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}})=Z_{q}^{\mathcal{S},\mathcyr{sh}}(E_{\mathbf{k}}\psi^{\mathcyr{sh}}(E_{\mathbf{l}})). From Corollary B.6, we see that E𝐤ψsh(E𝐥)(1)wt(𝐥)E(𝐤,𝐥¯)E_{\mathbf{k}}\psi^{\mathcyr{sh}}(E_{\mathbf{l}})\equiv(-1)^{\mathrm{wt}(\mathbf{l})}E_{(\mathbf{k},\overline{\mathbf{l}})} modulo 𝔢\hbar\,\mathfrak{e} if 𝐤\mathbf{k} or 𝐥\mathbf{l} is admissible. ∎

From Proposition 6.2 and Proposition 6.5, we obtain the double shuffle relation of qqSMZVs as follows.

Theorem 6.6.

Let d𝐤,𝐥,𝐦d_{\mathbf{k},\mathbf{l}}^{\,\bullet,\mathbf{m}} be the non-negative integer defined by (2.4).

  1. (i)

    For any index 𝐤\mathbf{k} and 𝐥\mathbf{l}, it holds that

    ζq𝒮(𝐤)ζq𝒮(𝐥)=𝐦d𝐤,𝐥,𝐦ζq𝒮(𝐦).\displaystyle\zeta_{q}^{\mathcal{S}}(\mathbf{k})\zeta_{q}^{\mathcal{S}}(\mathbf{l})=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\ast,\mathbf{m}}\zeta_{q}^{\mathcal{S}}(\mathbf{m}).
  2. (ii)

    Moreover, if at least one of 𝐤\mathbf{k} and 𝐥\mathbf{l} is admissible, it holds that

    (1)wt(𝐥)ζq𝒮(𝐤,𝐥¯)=𝐦d𝐤,𝐥sh,𝐦ζq𝒮(𝐦).\displaystyle(-1)^{\mathrm{wt}(\mathbf{l})}\zeta_{q}^{\mathcal{S}}(\mathbf{k},\overline{\mathbf{l}})=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\mathcyr{sh},\mathbf{m}}\zeta_{q}^{\mathcal{S}}(\mathbf{m}).
Proof.

From the definition of the harmonic product \ast_{\hbar}, we see that g𝐤g𝐥=𝐦d𝐤,𝐥,𝐦g𝐦g_{\mathbf{k}}\ast_{\hbar}g_{\mathbf{l}}=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\ast,\mathbf{m}}g_{\mathbf{m}}. Hence (i) is true because of Proposition 6.2.

We prove (ii) by using the map ι\iota defined by (3.14). Proposition 6.4 implies that there exists c𝐤,𝐥𝐦()[]c_{\mathbf{k},\mathbf{l}}^{\mathbf{m}}(\hbar)\in\mathbb{Q}[\hbar] such that E𝐤shE𝐥=𝐦c𝐤,𝐥𝐦()E𝐦E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}}=\sum_{\mathbf{m}}c_{\mathbf{k},\mathbf{l}}^{\mathbf{m}}(\hbar)E_{\mathbf{m}}. As shown in the proof of Theorem 5.14, we have ι(E𝐤)=z𝐤\iota(E_{\mathbf{k}})=z_{\mathbf{k}} for any index 𝐤\mathbf{k}. Hence we see that

ι(E𝐤shE𝐥)=𝐦c𝐤,𝐥𝐦(0)z𝐦.\displaystyle\iota(E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}})=\sum_{\mathbf{m}}c_{\mathbf{k},\mathbf{l}}^{\mathbf{m}}(0)z_{\mathbf{m}}.

On the other hand, Proposition 3.9 implies that

ι(E𝐤shE𝐥)=z𝐤shz𝐥=𝐦d𝐤,𝐥sh,𝐦z𝐦.\displaystyle\iota(E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}})=z_{\mathbf{k}}\,\mathcyr{sh}\,z_{\mathbf{l}}=\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\mathcyr{sh},\mathbf{m}}z_{\mathbf{m}}.

Hence c𝐤,𝐥𝐦(0)=d𝐤,𝐥sh,𝐦c_{\mathbf{k},\mathbf{l}}^{\mathbf{m}}(0)=d_{\mathbf{k},\mathbf{l}}^{\,\mathcyr{sh},\mathbf{m}} and

E𝐤shE𝐥𝐦d𝐤,𝐥sh,𝐦E𝐦𝔢.\displaystyle E_{\mathbf{k}}\,\mathcyr{sh}_{\hbar}\,E_{\mathbf{l}}-\sum_{\mathbf{m}}d_{\mathbf{k},\mathbf{l}}^{\,\mathcyr{sh},\mathbf{m}}E_{\mathbf{m}}\in\hbar\,\mathfrak{e}.

Therefore we see that (ii) is true from Proposition 6.5 and (1q)𝒵q𝒩q(1-q)\mathcal{Z}_{q}\subset\mathcal{N}_{q}. ∎

6.3. Ohno-type relation

From Theorem 6.6, we see that the qqSMZVs satisfy a part of the Ohno-type relation.

Theorem 6.7.

Let 𝐤\mathbf{k} be a non-empty admissible index. Then it holds that

(6.1) 𝐞(0)rwt(𝐞)=mζq𝒮(𝐤+𝐞)=𝐞(0)swt(𝐞)=mζq𝒮((𝐤+𝐞))\displaystyle\sum_{\begin{subarray}{c}\mathbf{e}\in(\mathbb{Z}_{\geq 0})^{r}\\ \mathrm{wt}(\mathbf{e})=m\end{subarray}}\zeta_{q}^{\mathcal{S}}(\mathbf{k}+\mathbf{e})=\sum_{\begin{subarray}{c}\mathbf{e}\in(\mathbb{Z}_{\geq 0})^{s}\\ \mathrm{wt}(\mathbf{e})=m\end{subarray}}\zeta_{q}^{\mathcal{S}}((\mathbf{k}^{\vee}+\mathbf{e})^{\vee})

for any m1m\geq 1, where r=dep(𝐤)r=\mathrm{dep}(\mathbf{k}) and s=dep(𝐤)s=\mathrm{dep}(\mathbf{k}^{\vee}).

Proof.

We define the \mathbb{Q}-linear map η:𝔥1𝒵q/(𝒩q+𝒫q)\eta:\mathfrak{h}^{1}\rightarrow\mathcal{Z}_{q}/(\mathcal{N}_{q}+\mathcal{P}_{q}) by η(z𝐤)=ζq𝒮(𝐤)\eta(z_{\mathbf{k}})=\zeta^{\mathcal{S}}_{q}(\mathbf{k}) for an index 𝐤\mathbf{k}. Theorem 6.6 and (5.7) imply the following properties:

  1. (i)

    For any index 𝐤\mathbf{k} and n1n\geq 1, it holds that η(z𝐤z1n)=0\eta(z_{\mathbf{k}}\ast z_{1}^{n})=0.

  2. (ii)

    For any admissible index 𝐤\mathbf{k} and n0n\geq 0, it holds that η(z𝐤shz1n)=(1)nη(z𝐤z1n)\eta(z_{\mathbf{k}}\,\mathcyr{sh}\,z_{1}^{n})=(-1)^{n}\eta(z_{\mathbf{k}}z_{1}^{n}).

For a non-empty index 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) and s0s\geq 0, we define the element as(𝐤)a_{s}(\mathbf{k}) of 𝔥1\mathfrak{h}^{1} by

as(𝐤)=1ir(y1lki(yjl(i)x)),\displaystyle a_{s}(\mathbf{k})=\sum\prod_{1\leq i\leq r}^{\curvearrowright}\left(y\prod_{1\leq l\leq k_{i}}^{\curvearrowright}(y^{j_{l}^{(i)}}x)\right),

where the sum is over the set

{(jl(i))1ir1lki(0)wt(𝐤)i=1rl=1kijl(i)=s}.\displaystyle\left\{(j_{l}^{(i)})_{\begin{subarray}{c}1\leq i\leq r\\ 1\leq l\leq k_{i}\end{subarray}}\in(\mathbb{Z}_{\geq 0})^{\mathrm{wt}(\mathbf{k})}\mid\sum_{i=1}^{r}\sum_{l=1}^{k_{i}}j_{l}^{(i)}=s\right\}.

We also set

A𝐤,s,p=λ1,,λr{0,1}λ1++λr=pas(k1+λ11,,kr+λr1)\displaystyle A_{\mathbf{k},s,p}=\sum_{\begin{subarray}{c}\lambda_{1},\ldots,\lambda_{r}\in\{0,1\}\\ \lambda_{1}+\cdots+\lambda_{r}=p\end{subarray}}a_{s}(k_{1}+\lambda_{1}-1,\ldots,k_{r}+\lambda_{r}-1)

for s,p0s,p\geq 0. Then we have

p=0min(n,r)m+s=npm,s0(1)sA𝐤,s,pshz1m=z𝐤z1n\displaystyle\sum_{p=0}^{\min(n,r)}\sum_{\begin{subarray}{c}m+s=n-p\\ m,s\geq 0\end{subarray}}(-1)^{s}A_{\mathbf{k},s,p}\,\mathcyr{sh}\,z_{1}^{m}=z_{\mathbf{k}}\ast z_{1}^{n}

for any n1n\geq 1 (see equation (3) in [7]).

Now suppose that 𝐤\mathbf{k} is admissible. Then A𝐤,s,pA_{\mathbf{k},s,p} belongs to 𝐥I0z𝐥\sum_{\mathbf{l}\in I_{0}}\mathbb{Q}z_{\mathbf{l}} for any s,p0s,p\geq 0. Therefore we can use the properties (i) and (ii), and see that

p=0min(n,r)(1)pm+s=npm,s0η(A𝐤,s,pz1m)=0\displaystyle\sum_{p=0}^{\min(n,r)}(-1)^{p}\sum_{\begin{subarray}{c}m+s=n-p\\ m,s\geq 0\end{subarray}}\eta(A_{\mathbf{k},s,p}z_{1}^{m})=0

for any n1n\geq 1. From the above identity, we obtain (6.1) in the same way as the proof of Theorem 2.5 in [7]. ∎

Appendix A Proof of Proposition 3.4

Lemma A.1.

Suppose that α0\alpha\geq 0 and max(1,α)β2α+1\max{(1,\alpha)}\leq\beta\leq 2\alpha+1. Then the function f(x)=xα(1x)/(1xβ)f(x)=x^{\alpha}(1-x)/(1-x^{\beta}) is non-decreasing on the interval [0,1)[0,1).

Proof.

If β=1\beta=1, the statement is trivial. We assume that β>1\beta>1. From the assumption we see that α(β1)/2>0\alpha\geq(\beta-1)/2>0. Set

g(x)=x1α(1xβ)2f(x)=α(α+1)x+(βα)xβ+(αβ+1)xβ+1.\displaystyle g(x)=x^{1-\alpha}(1-x^{\beta})^{2}f^{\prime}(x)=\alpha-(\alpha+1)x+(\beta-\alpha)x^{\beta}+(\alpha-\beta+1)x^{\beta+1}.

Since g(0)=α>0g(0)=\alpha>0 and g(1)=0g(1)=0, it suffices to show that g(x)<0g^{\prime}(x)<0 on the interval (0,1)(0,1). Set

h(x)=xβg(x)=(α+1)xβ+β(βα)x1+(αβ+1)(β+1).\displaystyle h(x)=x^{-\beta}g^{\prime}(x)=-(\alpha+1)x^{-\beta}+\beta(\beta-\alpha)x^{-1}+(\alpha-\beta+1)(\beta+1).

Then we see that h(1)=0h(1)=0 and, from the assumption,

h(x)=βxβ1(α+1(βα)xβ1)>βxβ1(2αβ+1)0\displaystyle h^{\prime}(x)=\beta x^{-\beta-1}(\alpha+1-(\beta-\alpha)x^{\beta-1})>\beta x^{-\beta-1}(2\alpha-\beta+1)\geq 0

for 0<x<10<x<1. Therefore we see that h(x)<0h(x)<0, and hence g(x)<0g^{\prime}(x)<0, on the interval (0,1)(0,1). ∎

Corollary A.2.

Under the assumption of Lemma A.1, it holds that

xα1x1xβ1β\displaystyle x^{\alpha}\frac{1-x}{1-x^{\beta}}\leq\frac{1}{\beta}

for 0<x<10<x<1.

Proof.

It is because f(x)1/βf(x)\to 1/\beta as x10x\to 1-0. ∎

Proposition A.3.

Suppose that 0<q<10<q<1. Set

Bq(α1,,αr;β1,,βr)=l1,,lr1j=1r(ljαj)(1q)αjqlj/2(l1++lj)βj+1.\displaystyle B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r})=\sum_{l_{1},\ldots,l_{r}\geq 1}\prod_{j=1}^{r}\binom{l_{j}}{\alpha_{j}}\frac{(1-q)^{\alpha_{j}}q^{l_{j}/2}}{(l_{1}+\cdots+l_{j})^{\beta_{j}+1}}.

Then, for any non-negative integer α1,,αr,β1,,βr\alpha_{1},\ldots,\alpha_{r},\beta_{1},\ldots,\beta_{r}, it holds that

0Zq((e1g1)α1gβ1+1(e1g1)αrgβr+1)Bq(α1,,αr;β1,,βr).\displaystyle 0\leq Z_{q}((e_{1}-g_{1})^{\alpha_{1}}g_{\beta_{1}+1}\cdots(e_{1}-g_{1})^{\alpha_{r}}g_{\beta_{r}+1})\leq B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r}).
Proof.

From Corollary A.2, we see that

0qn[n]=q(n1)/21q1qnq(n+1)/2q(n+1)/2n\displaystyle 0\leq\frac{q^{n}}{[n]}=q^{(n-1)/2}\frac{1-q}{1-q^{n}}q^{(n+1)/2}\leq\frac{q^{(n+1)/2}}{n}

for any n1n\geq 1. Therefore, it holds that

0\displaystyle 0 Zq((e1g1)α1gβ1+1(e1g1)αrgβr+1)\displaystyle\leq Z_{q}((e_{1}-g_{1})^{\alpha_{1}}g_{\beta_{1}+1}\cdots(e_{1}-g_{1})^{\alpha_{r}}g_{\beta_{r}+1})
=0=n0<n1<<nrj=1r(1q)αj(njnj11αj)(qnj[nj])βj+1\displaystyle=\sum_{0=n_{0}<n_{1}<\cdots<n_{r}}\prod_{j=1}^{r}(1-q)^{\alpha_{j}}\binom{n_{j}-n_{j-1}-1}{\alpha_{j}}\left(\frac{q^{n_{j}}}{[n_{j}]}\right)^{\beta_{j}+1}
0=n0<n1<<nrj=1r(1q)αj(njnj1αj)(q(nj+1)/2nj)βj+1.\displaystyle\leq\sum_{0=n_{0}<n_{1}<\cdots<n_{r}}\prod_{j=1}^{r}(1-q)^{\alpha_{j}}\binom{n_{j}-n_{j-1}}{\alpha_{j}}\left(\frac{q^{(n_{j}+1)/2}}{n_{j}}\right)^{\beta_{j}+1}.

Set lj=njnj1l_{j}=n_{j}-n_{j-1} for 1jr1\leq j\leq r. Then the right-hand side is dominated by Bq(α1,,αr;β1,,βr)B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r}) since 0<q<10<q<1. ∎

Proposition A.4.

Suppose that 0<q<10<q<1 and α1,,αr,β1,,βr\alpha_{1},\ldots,\alpha_{r},\beta_{1},\ldots,\beta_{r} are non-negative integers. Then it holds that

(A.1) Bq(α1,,αr;β1,,βr)=O((log(1q))r)(q10).\displaystyle B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r})=O((-\log{(1-q)})^{r})\qquad(q\to 1-0).

Moreover, if αs1\alpha_{s}\geq 1 and βt1\beta_{t}\geq 1 for some 1str1\leq s\leq t\leq r, then we have

(A.2) Bq(α1,,αr;β1,,βr)=(1q)O((log(1q))r)(q10).\displaystyle B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r})=(1-q)\,O((-\log{(1-q)})^{r})\qquad(q\to 1-0).
Proof.

For a non-negative integer α\alpha, we define

φα(x)=(1x)αl1(lα)xl/2l,φ~α(x)=(1x)αl1(lα)xl/2l2.\displaystyle\varphi_{\alpha}(x)=(1-x)^{\alpha}\sum_{l\geq 1}\binom{l}{\alpha}\frac{x^{l/2}}{l},\qquad\tilde{\varphi}_{\alpha}(x)=(1-x)^{\alpha}\sum_{l\geq 1}\binom{l}{\alpha}\frac{x^{l/2}}{l^{2}}.

We have

φ0(x)=log(1x1/2),φα(x)=xα/2α(1+x1/2)α(α1).\displaystyle\varphi_{0}(x)=-\log{(1-x^{1/2})},\qquad\varphi_{\alpha}(x)=\frac{x^{\alpha/2}}{\alpha}(1+x^{1/2})^{\alpha}\quad(\alpha\geq 1).

Hence φα(x)=O(log(1x))\varphi_{\alpha}(x)=O(-\log{(1-x)}) as x10x\to 1-0 for any α0\alpha\geq 0. If α1\alpha\geq 1, we see that

0φ~α(x)=(1x)αl11α(l1α1)xl/2l1xαφα1(x)\displaystyle 0\leq\tilde{\varphi}_{\alpha}(x)=(1-x)^{\alpha}\sum_{l\geq 1}\frac{1}{\alpha}\binom{l-1}{\alpha-1}\frac{x^{l/2}}{l}\leq\frac{1-x}{\alpha}\varphi_{\alpha-1}(x)

for 0x10\leq x\leq 1. Hence φ~α(x)=(1x)O(log(1x))\tilde{\varphi}_{\alpha}(x)=(1-x)\,O(-\log{(1-x)}) as x10x\to 1-0 for any α1\alpha\geq 1.

Since βj0\beta_{j}\geq 0 for 1jr1\leq j\leq r, we see that

0Bq(α1,,αr;β1,,βr)j=1rφαj(q).\displaystyle 0\leq B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r})\leq\prod_{j=1}^{r}\varphi_{\alpha_{j}}(q).

Hence we have (A.1). Now assume further that αs1\alpha_{s}\geq 1 and βt1\beta_{t}\geq 1 for some 1str1\leq s\leq t\leq r. If s=ts=t, it holds that

0Bq(α1,,αr;β1,,βr)φ~αs(q)1jrjsφαj(q).\displaystyle 0\leq B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r})\leq\tilde{\varphi}_{\alpha_{s}}(q)\prod_{\begin{subarray}{c}1\leq j\leq r\\ j\not=s\end{subarray}}\varphi_{\alpha_{j}}(q).

Hence we obtain (A.2) if s=ts=t. If s<ts<t, we see that

0\displaystyle 0 Bq(α1,,αr;β1,,βr)\displaystyle\leq B_{q}(\alpha_{1},\ldots,\alpha_{r};\beta_{1},\ldots,\beta_{r})
l1,,lr11jrjs,t{(1q)αj(ljαj)qlj/2lj}(1q)αs+αt(lsαs)(ltαt)qls/2(l1++ls)βs+βt+1qlt/2l1++lt\displaystyle\leq\sum_{l_{1},\ldots,l_{r}\geq 1}\prod_{\begin{subarray}{c}1\leq j\leq r\\ j\not=s,t\end{subarray}}\left\{(1-q)^{\alpha_{j}}\binom{l_{j}}{\alpha_{j}}\frac{q^{l_{j}/2}}{l_{j}}\right\}(1-q)^{\alpha_{s}+\alpha_{t}}\binom{l_{s}}{\alpha_{s}}\binom{l_{t}}{\alpha_{t}}\frac{q^{l_{s}/2}}{(l_{1}+\cdots+l_{s})^{\beta_{s}+\beta_{t}+1}}\frac{q^{l_{t}/2}}{l_{1}+\cdots+l_{t}}
l1,,lr11jrjs{(1q)αj(ljαj)qlj/2lj}(1q)αs(lsαs)qls/2ls2\displaystyle\leq\sum_{l_{1},\ldots,l_{r}\geq 1}\prod_{\begin{subarray}{c}1\leq j\leq r\\ j\not=s\end{subarray}}\left\{(1-q)^{\alpha_{j}}\binom{l_{j}}{\alpha_{j}}\frac{q^{l_{j}/2}}{l_{j}}\right\}(1-q)^{\alpha_{s}}\binom{l_{s}}{\alpha_{s}}\frac{q^{l_{s}/2}}{l_{s}^{2}}
=φ~αs(q)1jrjsφαj(q)\displaystyle=\tilde{\varphi}_{\alpha_{s}}(q)\prod_{\begin{subarray}{c}1\leq j\leq r\\ j\not=s\end{subarray}}\varphi_{\alpha_{j}}(q)

since βt1\beta_{t}\geq 1. Thus we get (A.2) in the case where s<ts<t either. ∎

Proof of Proposition 3.4.

From Lemma A.1 and the monotone convergence theorem, we see that Zq(g𝐤)ζ(𝐤)Z_{q}(g_{\mathbf{k}})\to\zeta(\mathbf{k}) as q10q\to 1-0 for any admissible index 𝐤\mathbf{k}. Suppose that ww is an element of 0^\widehat{\mathfrak{H}^{0}} of the form (3.2). Proposition A.3 and (A.1) imply that Zq(w)=O((log(1q))r)Z_{q}(w)=O((-\log{(1-q)})^{r}) as q10q\to 1-0. Hence Zq(w)0Z_{q}(\hbar w)\to 0 in the limit as q10q\to 1-0. If w𝔫0w\in\mathfrak{n}_{0}, then there exist ss and tt such that αs1,βt1\alpha_{s}\geq 1,\beta_{t}\geq 1 and 1str1\leq s\leq t\leq r. Then Proposition A.3 and (A.2) imply that Zq(w)0Z_{q}(w)\to 0 in the limit as q10q\to 1-0. Thus we see that limq10Zq(w)=0\lim_{q\to 1-0}Z_{q}(w)=0 for any w𝔫w\in\mathfrak{n}. ∎

Appendix B Proofs

B.1. Formulas of shuffle product

Proposition B.1.

Suppose that U(X)U(X) and V(X)V(X) belong to the ideal X𝒞[b][[X]]X\mathcal{C}[b][[X]] of the formal power series ring 𝒞[b][[X]]\mathcal{C}[b][[X]] generated by XX. Then it holds that

(w11U(X)ashw11V(Y)a){1(U(X)+V(Y)+U(X)V(Y))a}\displaystyle\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}\,w^{\prime}\frac{1}{1-V(Y)a}\right)\left\{1-(U(X)+V(Y)+\hbar\,U(X)V(Y))a\right\}
=wshw+(w11U(X)ashw)(1U(X)a)+(wshw11V(Y)a)(1V(Y)a)\displaystyle=-w\,\mathcyr{sh}_{\hbar}\,w^{\prime}+\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}\,w^{\prime}\right)(1-U(X)a)+\left(w\,\mathcyr{sh}_{\hbar}\,w^{\prime}\frac{1}{1-V(Y)a}\right)(1-V(Y)a)

for any w,ww,w^{\prime}\in\mathfrak{H}.

Proof.

Set

I(X,Y)=w11U(X)ashw11V(Y)a.\displaystyle I(X,Y)=w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}\,w^{\prime}\frac{1}{1-V(Y)a}.

Using (1U(X)a)1=1+(1U(X)a)1U(X)a(1-U(X)a)^{-1}=1+(1-U(X)a)^{-1}U(X)a, we see that

I(X,Y)\displaystyle I(X,Y) =wshw+wshw11V(Y)a+w11U(X)ashw\displaystyle=-w\,\mathcyr{sh}_{\hbar}\,w^{\prime}+w\,\mathcyr{sh}_{\hbar}w^{\prime}\frac{1}{1-V(Y)a}+w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}w^{\prime}
+w11U(X)aU(X)ashw11V(Y)aV(Y)a.\displaystyle+w\frac{1}{1-U(X)a}U(X)a\,\mathcyr{sh}_{\hbar}w^{\prime}\frac{1}{1-V(Y)a}V(Y)a.

Since all the coefficients of U(X)U(X) and V(Y)V(Y) are polynomials in bb, the fourth term of the right hand side is equal to

(w11U(X)ashw11V(Y)aV(Y)a)U(X)a\displaystyle\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}w^{\prime}\frac{1}{1-V(Y)a}V(Y)a\right)U(X)a
+(w11U(X)aU(X)ashw11V(Y)a)V(Y)a\displaystyle+\left(w\frac{1}{1-U(X)a}U(X)a\,\mathcyr{sh}_{\hbar}w^{\prime}\frac{1}{1-V(Y)a}\right)V(Y)a
+(w11U(X)ashw11V(Y)a)U(X)V(Y)a.\displaystyle+\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}w^{\prime}\frac{1}{1-V(Y)a}\right)\hbar\,U(X)V(Y)a.

Using (1U(X)a)1=1+(1U(X)a)1U(X)a(1-U(X)a)^{-1}=1+(1-U(X)a)^{-1}U(X)a again, we see that it is equal to

I(X,Y)(U(X)+V(Y)+U(X)V(Y))a\displaystyle I(X,Y)(U(X)+V(Y)+\hbar\,U(X)V(Y))a
(w11U(X)ashw)U(X)a(wshw11V(Y)a)V(Y)a.\displaystyle-\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}w^{\prime}\right)U(X)a-\left(w\,\mathcyr{sh}_{\hbar}w^{\prime}\frac{1}{1-V(Y)a}\right)V(Y)a.

Thus we get the desired equality. ∎

Proposition B.2.

Suppose that U(X)X𝒞[b][[X]]U(X)\in X\mathcal{C}[b][[X]]. Then it holds that

(w11U(X)ashwa)(1U(X)a)\displaystyle\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}\,w^{\prime}a\right)(1-U(X)a)
=wshwa(wshw)a+(w11U(X)ashw)(1+U(X))a\displaystyle=w\,\mathcyr{sh}_{\hbar}\,w^{\prime}a-(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})a+\left(w\frac{1}{1-U(X)a}\,\mathcyr{sh}_{\hbar}\,w^{\prime}\right)(1+\hbar\,U(X))a

for any w,ww,w^{\prime}\in\mathfrak{H}.

Proof.

Set V(Y)=YV(Y)=Y in Proposition B.1, differentiate the both hand sides with respect to YY and set Y=0Y=0. Then we get the desired equality. ∎

Proof of Lemma 3.10.

From Proposition B.1 and Proposition B.2, we see that

(ug111aXshvg111aY)(1(X+Y+XY)a)\displaystyle\left(ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,vg_{1}\frac{1}{1-aY}\right)(1-(X+Y+\hbar XY)a)
=ug1shvg1+(ug111aXshvg1)(1aX)+(ug1shvg111aY)(1aY)\displaystyle=-ug_{1}\,\mathcyr{sh}_{\hbar}\,vg_{1}+\left(ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,vg_{1}\right)(1-aX)+\left(ug_{1}\,\mathcyr{sh}_{\hbar}\,vg_{1}\frac{1}{1-aY}\right)(1-aY)
=ug1shvg1(ug1shv+ushvg1)g1\displaystyle=ug_{1}\,\mathcyr{sh}_{\hbar}\,vg_{1}-(ug_{1}\,\mathcyr{sh}_{\hbar}\,v+u\,\mathcyr{sh}_{\hbar}\,vg_{1})g_{1}
+(1+X)(ug111aXshv)g1+(1+Y)(ushv11aX)g1.\displaystyle+(1+\hbar X)\left(ug_{1}\frac{1}{1-aX}\,\mathcyr{sh}_{\hbar}\,v\right)g_{1}+(1+\hbar Y)\left(u\,\mathcyr{sh}_{\hbar}\,v\frac{1}{1-aX}\right)g_{1}.

Using (3.9), we get the desired equality. ∎

Corollary B.3.

For U(X)X𝒞[b][[X]]U(X)\in X\mathcal{C}[b][[X]], we define the map ρU(X):[[X]]\rho_{U(X)}:\mathfrak{H}\rightarrow\mathfrak{H}[[X]] by

ρU(X)(w)=(11U(X)g1shw)(1U(X)g1).\displaystyle\rho_{U(X)}(w)=\left(\frac{1}{1-U(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,w\right)\left(1-U(X)g_{1}\right).

Then the map ρU(X)\rho_{U(X)} is a 𝒞\mathcal{C}-algebra homomorphism with respect to the concatenation product on \mathfrak{H}, and we have

(B.1) ρU(X)(a)=11U(X)g1(1+bU(X))a,ρU(X)(b)=11U(X)g1b(1U(X)g1).\displaystyle\rho_{U(X)}(a)=\frac{1}{1-U(X)g_{1}}\left(1+\hbar b\,U(X)\right)a,\quad\rho_{U(X)}(b)=\frac{1}{1-U(X)g_{1}}b\left(1-U(X)g_{1}\right).
Proof.

Set w=1w=1 and U(X)U(X) to bU(X)bU(X) in Proposition B.2. Then we see that

ρU(X)(wa)=ρU(X)(w)11U(X)g1(1+bU(X))a.\displaystyle\rho_{U(X)}(w^{\prime}a)=\rho_{U(X)}(w^{\prime})\frac{1}{1-U(X)g_{1}}(1+\hbar bU(X))a.

From the definition of ρU(X)\rho_{U(X)}, we also see that

ρU(X)(wb)=ρU(X)(w)11U(X)g1b(1U(X)g1).\displaystyle\rho_{U(X)}(w^{\prime}b)=\rho_{U(X)}(w^{\prime})\frac{1}{1-U(X)g_{1}}b(1-U(X)g_{1}).

Since ρU(X)(1)=1\rho_{U(X)}(1)=1, we have (B.1) and see that ρU(X)\rho_{U(X)} is an algebra homomorphism. ∎

B.2. Generating function of E1mE_{1^{m}}

Here we calculate the generating function

E(X)=m=0E1mXm.\displaystyle E(X)=\sum_{m=0}^{\infty}E_{1^{m}}X^{m}.

Recall that

R(X)=ebX1b=n=1Xnn!(e1g1)n1.\displaystyle R(X)=\frac{e^{\hbar bX}-1}{\hbar b}=\sum_{n=1}^{\infty}\frac{X^{n}}{n!}(e_{1}-g_{1})^{n-1}.

Note that R(X)R(X) belongs to X𝒞[b][[X]]X\mathcal{C}[b][[X]].

Proposition B.4.

It holds that

(B.2) expsh(g1X)=11R(X)g1.\displaystyle\exp_{\mathcyr{sh}_{\hbar}}{\left(g_{1}X\right)}=\frac{1}{1-R(X)g_{1}}.
Proof.

The power series Ψ(X)=expsh(g1X)\Psi(X)=\exp_{\mathcyr{sh}_{\hbar}}{\left(g_{1}X\right)} is the unique solution of the differential equation Ψ(X)=g1shΨ(X)\Psi^{\prime}(X)=g_{1}\,\mathcyr{sh}_{\hbar}\,\Psi(X) satisfying Ψ(0)=1\Psi(0)=1. Since the right hand side with X=0X=0 is equal to one, it suffices to show that

ddX11R(X)g1=g1sh11R(X)g1.\displaystyle\frac{d}{dX}\frac{1}{1-R(X)g_{1}}=g_{1}\,\mathcyr{sh}_{\hbar}\,\frac{1}{1-R(X)g_{1}}.

The right hand side is equal to ρR(X)(g1)(1R(X)g1)1\rho_{R(X)}(g_{1})(1-R(X)g_{1})^{-1}. Hence (B.1) implies that

g1sh11R(X)g1\displaystyle g_{1}\,\mathcyr{sh}_{\hbar}\,\frac{1}{1-R(X)g_{1}} =ρR(X)(b)ρR(X)(a)11R(X)g1\displaystyle=\rho_{R(X)}(b)\rho_{R(X)}(a)\frac{1}{1-R(X)g_{1}}
=11R(X)g1ebXg111R(X)g1=ddX11R(X)g1\displaystyle=\frac{1}{1-R(X)g_{1}}e^{\hbar bX}g_{1}\frac{1}{1-R(X)g_{1}}=\frac{d}{dX}\frac{1}{1-R(X)g_{1}}

since R(X)=ebXR^{\prime}(X)=e^{\hbar bX}. ∎

Proposition B.5.

It holds that

E(X)=11R(X)g1R(X)X.\displaystyle E(X)=\frac{1}{1-R(X)g_{1}}\frac{R(X)}{X}.
Proof.

Note that g1g_{1} and b\hbar b are commutative with respect to the shuffle product sh\mathcyr{sh}_{\hbar}. From the definition of E1mE_{1^{m}}, we see that

E(X)\displaystyle E(X) =m=01(m+1)!s=0mg1sh(ms)sh(g1+b)shsXm\displaystyle=\sum_{m=0}^{\infty}\frac{1}{(m+1)!}\sum_{s=0}^{m}g_{1}^{\mathcyr{sh}_{\hbar}(m-s)}\,\mathcyr{sh}_{\hbar}\,(g_{1}+\hbar b)^{\mathcyr{sh}_{\hbar}s}X^{m}
=msj01(m+1)!(sj)g1sh(mj)(b)jXm.\displaystyle=\sum_{m\geq s\geq j\geq 0}\frac{1}{(m+1)!}\binom{s}{j}g_{1}^{\mathcyr{sh}_{\hbar}(m-j)}(\hbar b)^{j}X^{m}.

Using

s=jm(sj)=s=jm((s+1j+1)(sj+1))=(m+1j+1)\displaystyle\sum_{s=j}^{m}\binom{s}{j}=\sum_{s=j}^{m}\left(\binom{s+1}{j+1}-\binom{s}{j+1}\right)=\binom{m+1}{j+1}

and (B.2), we see that

E(X)=mj0g1sh(mj)(mj)!Xmj(b)j(j+1)!Xj=expsh(g1X)ebX1bX=11R(X)g1R(X)X.\displaystyle E(X)=\sum_{m\geq j\geq 0}\frac{g_{1}^{\mathcyr{sh}_{\hbar}(m-j)}}{(m-j)!}X^{m-j}\frac{(\hbar b)^{j}}{(j+1)!}X^{j}=\exp_{\mathcyr{sh}_{\hbar}}{(g_{1}X)}\frac{e^{\hbar bX}-1}{\hbar bX}=\frac{1}{1-R(X)g_{1}}\frac{R(X)}{X}.

Corollary B.6.

For any index 𝐤\mathbf{k}, it holds that ψsh(E𝐤)(1)wt(𝐤)E𝐤¯\psi^{\mathcyr{sh}}(E_{\mathbf{k}})\equiv(-1)^{\mathrm{wt}(\mathbf{k})}E_{\overline{\mathbf{k}}} modulo 𝔢\hbar\mathfrak{e}. Moreover, if 𝐤\mathbf{k} (resp. 𝐤¯\overline{\mathbf{k}}) is admissible, then ψsh(E𝐤)(1)wt(𝐤)E𝐤¯\psi^{\mathcyr{sh}}(E_{\mathbf{k}})-(-1)^{\mathrm{wt}(\mathbf{k})}E_{\overline{\mathbf{k}}} belongs to j2ej𝔢\hbar\sum_{j\geq 2}e_{j}\mathfrak{e} (resp. j2𝔢ej\hbar\sum_{j\geq 2}\mathfrak{e}e_{j}).

Proof.

For k2k\geq 2, using (3.1), we see that

(B.3) ψsh(ek)=ψsh(gk+gk1)=(1)kba(a+)k1=(1)kj=2k(k2j2)kjej.\displaystyle\psi^{\mathcyr{sh}}(e_{k})=\psi^{\mathcyr{sh}}(g_{k}+\hbar g_{k-1})=(-1)^{k}ba(a+\hbar)^{k-1}=(-1)^{k}\sum_{j=2}^{k}\binom{k-2}{j-2}\hbar^{k-j}e_{j}.

Next we calculate ψsh(E1m)\psi^{\mathcyr{sh}}(E_{1^{m}}) using Proposition B.5. Since ψsh(g1)=e1\psi^{\mathcyr{sh}}(g_{1})=-e_{1} and ψsh(b)=b\psi^{\mathcyr{sh}}(\hbar b)=\hbar b, it holds that

ψsh(E(X))=R(X)X11+e1R(X)=11+R(X)e1R(X)X.\displaystyle\psi^{\mathcyr{sh}}(E(X))=\frac{R(X)}{X}\frac{1}{1+e_{1}R(X)}=\frac{1}{1+R(X)e_{1}}\frac{R(X)}{X}.

We have

(B.4) 1+R(X)e1=1+R(X)(e1g1)+R(X)g1=ebX+R(X)g1=ebX(1R(X)g1).\displaystyle 1+R(X)e_{1}=1+R(X)(e_{1}-g_{1})+R(X)g_{1}=e^{\hbar bX}+R(X)g_{1}=e^{\hbar bX}\left(1-R(-X)g_{1}\right).

Thus we see that ψsh(E(X))=E(X)\psi^{\mathcyr{sh}}(E(X))=E(-X). Hence

(B.5) ψsh(E1m)=(1)mE1m\displaystyle\psi^{\mathcyr{sh}}(E_{1^{m}})=(-1)^{m}E_{1^{m}}

for m0m\geq 0. Therefore, for an index 𝐤\mathbf{k} of the form (5.1), we have

ψsh(E𝐤)=(1)wt(𝐤)j1=0t1jr=0tr{l=1rtljl(tljl)}E1srejr+2E1s1ej1+2E1s0,\displaystyle\psi^{\mathcyr{sh}}(E_{\mathbf{k}})=(-1)^{\mathrm{wt}(\mathbf{k})}\sum_{j_{1}=0}^{t_{1}}\cdots\sum_{j_{r}=0}^{t_{r}}\left\{\prod_{l=1}^{r}\hbar^{t_{l}-j_{l}}\binom{t_{l}}{j_{l}}\right\}E_{1^{s_{r}}}e_{j_{r}+2}\cdots E_{1^{s_{1}}}e_{j_{1}+2}E_{1^{s_{0}}},

and it implies the statement. ∎

B.3. Proof of Proposition 5.4

Note that

(B.6) (𝒞A𝒞)a0\displaystyle(\mathcal{C}\langle A\rangle\setminus\mathcal{C})a\subset\mathfrak{H}^{0}

because (e1g1)a=g1(e_{1}-g_{1})a=\hbar g_{1} and gka=gk+1g_{k}a=g_{k+1} for k1k\geq 1.

We define the 𝒞\mathcal{C}-trilinear map Ksh:𝒞A×𝒞A×𝒞A1^K_{\hbar}^{\mathcyr{sh}}:\mathcal{C}\langle A\rangle\times\mathcal{C}\langle A\rangle\times\mathcal{C}\langle A\rangle\rightarrow\widehat{\mathfrak{H}^{1}} by

Ksh(w,u1ur,w)=i=0rwu1uishψsh(ui+1urw)\displaystyle K_{\hbar}^{\mathcyr{sh}}(w,u_{1}\cdots u_{r},w^{\prime})=\sum_{i=0}^{r}wu_{1}\cdots u_{i}\mathcyr{sh}_{\hbar}\psi^{\mathcyr{sh}}(u_{i+1}\cdots u_{r}w^{\prime})

for w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle and u1,,urAu_{1},\ldots,u_{r}\in A. It suffices to show that Ksh(E𝐦,E1r,E𝐦¯)K_{\hbar}^{\mathcyr{sh}}(E_{\mathbf{m}},E_{1^{r}},E_{\overline{\mathbf{m}^{\prime}}}) belongs to 0\mathfrak{H}^{0} for any admissible index 𝐦,𝐦\mathbf{m},\mathbf{m}^{\prime} and r0r\geq 0.

Lemma B.7.

For any w,w𝒞Aw,w^{\prime}\in\mathcal{C}\langle A\rangle, it holds that

(B.7) Ksh(w,E(X),w)=w11R(X)g1shψsh(w)11R(X)g1.\displaystyle K_{\hbar}^{\mathcyr{sh}}(w,E(X),w^{\prime})=w\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,\psi^{\mathcyr{sh}}(w^{\prime})\frac{1}{1-R(-X)g_{1}}.
Proof.

Decompose (1R(X)g1)1=1+(1R(X)g1)1R(X)g1(1-R(X)g_{1})^{-1}=1+(1-R(X)g_{1})^{-1}R(X)g_{1}. Then we obtain

Ksh(w,E(X),w)=Ksh(w,R(X)X,w)+Ksh(w,11R(X)g1R(X)g1R(X)X,w).\displaystyle K_{\hbar}^{\mathcyr{sh}}(w,E(X),w^{\prime})=K_{\hbar}^{\mathcyr{sh}}(w,\frac{R(X)}{X},w^{\prime})+K_{\hbar}^{\mathcyr{sh}}(w,\frac{1}{1-R(X)g_{1}}R(X)g_{1}\frac{R(X)}{X},w^{\prime}).

From the definition of KshK_{\hbar}^{\mathcyr{sh}}, we see that the second term in the right hand side is equal to

Ksh(w,11R(X)g1R(X),g1R(X)Xw)+Ksh(w11R(X)g1R(X)g1,R(X)X,w)\displaystyle K_{\hbar}^{\mathcyr{sh}}(w,\frac{1}{1-R(X)g_{1}}R(X),g_{1}\frac{R(X)}{X}w^{\prime})+K_{\hbar}^{\mathcyr{sh}}(w\frac{1}{1-R(X)g_{1}}R(X)g_{1},\frac{R(X)}{X},w^{\prime})
=Ksh(w,E(X),g1R(X)w)+Ksh(w(11R(X)g11),R(X)X,w).\displaystyle=K_{\hbar}^{\mathcyr{sh}}(w,E(X),g_{1}R(X)w^{\prime})+K_{\hbar}^{\mathcyr{sh}}(w(\frac{1}{1-R(X)g_{1}}-1),\frac{R(X)}{X},w^{\prime}).

Hence, it holds that

Ksh(w,E(X),(1g1R(X))w)=Ksh(w11R(X)g1,R(X)X,w).\displaystyle K_{\hbar}^{\mathcyr{sh}}(w,E(X),(1-g_{1}R(X))w^{\prime})=K_{\hbar}^{\mathcyr{sh}}(w\frac{1}{1-R(X)g_{1}},\frac{R(X)}{X},w^{\prime}).

Since ψsh(b)=b\psi^{\mathcyr{sh}}(\hbar b)=\hbar b, we see that the right hand side above is equal to

(w11R(X)g1shψsh(w))ebX=w11R(X)g1sh(ψsh(w)ebX).\displaystyle(w\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,\psi^{\mathcyr{sh}}(w^{\prime}))e^{\hbar bX}=w\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,(\psi^{\mathcyr{sh}}(w^{\prime})e^{\hbar bX}).

Therefore, by replacing ww^{\prime} with (1g1R(X))1w(1-g_{1}R(X))^{-1}w^{\prime}, we obtain

Ksh(w,E(X),w)=w11R(X)g1sh(ψsh(11g1R(X)w)ebX).\displaystyle K_{\hbar}^{\mathcyr{sh}}(w,E(X),w^{\prime})=w\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,(\psi^{\mathcyr{sh}}(\frac{1}{1-g_{1}R(X)}w^{\prime})e^{\hbar bX}).

Since ψsh(R(X))=R(X)\psi^{\mathcyr{sh}}(R(X))=R(X) and ψsh(g1)=e1\psi^{\mathcyr{sh}}(g_{1})=-e_{1}, it holds that

ψsh(11g1R(X)w)=ψsh(w)11+R(X)e1.\displaystyle\psi^{\mathcyr{sh}}(\frac{1}{1-g_{1}R(X)}w^{\prime})=\psi^{\mathcyr{sh}}(w^{\prime})\frac{1}{1+R(X)e_{1}}.

Now the desired equality (B.7) follows from (B.4). ∎

Since R(X)+R(X)+bR(X)R(X)=0R(X)+R(-X)+\hbar b\,R(X)R(-X)=0, from Proposition B.1, we see that the right hand side of (B.7) is equal to

(B.8) wshψsh(w)+(w11R(X)g1shψsh(w))(1R(X)g1)\displaystyle{}-w\,\mathcyr{sh}_{\hbar}\,\psi^{\mathcyr{sh}}(w^{\prime})+\left(w\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,\psi^{\mathcyr{sh}}(w^{\prime})\right)(1-R(X)g_{1})
+(wshψsh(w)11R(X)g1)(1R(X)g1).\displaystyle+\left(w\,\mathcyr{sh}_{\hbar}\,\psi^{\mathcyr{sh}}(w^{\prime})\frac{1}{1-R(-X)g_{1}}\right)(1-R(-X)g_{1}).

Now set w=E𝐦w=E_{\mathbf{m}} and w=E𝐦¯w^{\prime}=E_{\overline{\mathbf{m}^{\prime}}} with admissible indices 𝐦\mathbf{m} and 𝐦\mathbf{m}^{\prime}. We have E𝐦shψsh(E𝐦¯)(1)wt(𝐦)E𝐦shE𝐦E_{\mathbf{m}}\mathcyr{sh}_{\hbar}\psi^{\mathcyr{sh}}(E_{\overline{\mathbf{m}^{\prime}}})\equiv(-1)^{\mathrm{wt}(\mathbf{m}^{\prime})}E_{\mathbf{m}}\mathcyr{sh}_{\hbar}E_{\mathbf{m}^{\prime}} modulo 0^\hbar\widehat{\mathfrak{H}^{0}} because of Proposition 3.3, Corollary B.6 and E𝐦0^E_{\mathbf{m}}\in\widehat{\mathfrak{H}^{0}}. From the definition of the shuffle product and (B.6), we see that E𝐦shE𝐦E_{\mathbf{m}}\mathcyr{sh}_{\hbar}E_{\mathbf{m}^{\prime}} belongs to 0\mathfrak{H}^{0}, hence so does the first term of (B.8) with w=E𝐦w=E_{\mathbf{m}} and w=E𝐦¯w^{\prime}=E_{\overline{\mathbf{m}^{\prime}}}. Since 𝐦\mathbf{m} and 𝐦\mathbf{m}^{\prime} are admissible, we can write E𝐦=uaE_{\mathbf{m}}=ua and ψsh(E𝐦¯)=va\psi^{\mathcyr{sh}}(E_{\overline{\mathbf{m}^{\prime}}})=va with some u,v0^u,v\in\widehat{\mathfrak{H}^{0}} because of (B.3). Then, from Proposition B.2, the second term of (B.8) is equal to

{ushva+ushv+(ua11R(X)g1shv)ebX}a,\displaystyle\left\{u\,\mathcyr{sh}_{\hbar}\,va+\hbar\,u\,\mathcyr{sh}_{\hbar}\,v+\left(ua\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,v\right)e^{\hbar bX}\right\}a,

which belongs to 0[[X]]\mathfrak{H}^{0}[[X]] because of (B.6). Similarly, the third term of (B.8) also belongs to 0[[X]]\mathfrak{H}^{0}[[X]]. Thus we find that Ksh(E𝐦,E(X),E𝐦¯)K_{\hbar}^{\mathcyr{sh}}(E_{\mathbf{m}},E(X),E_{\overline{\mathbf{m}^{\prime}}}) belongs to 0[[X]]\mathfrak{H}^{0}[[X]], and this completes the proof of Proposition 5.4.

B.4. Proof of Lemma 5.9

We use the map ρR(X)\rho_{R(X)} and ρX\rho_{X} defined in Corollary B.3 with U(X)=R(X)U(X)=R(X) and U(X)=XU(X)=X, respectively. Let ww\in\mathfrak{H}. Since ρR(X)\rho_{R(X)} is a 𝒞\mathcal{C}-algebra homomorphism and ρR(X)(g1)=(1R(X)g1)1ebXg1\rho_{R(X)}(g_{1})=(1-R(X)g_{1})^{-1}e^{\hbar bX}g_{1}, we have

11R(X)g1shw11g1X\displaystyle\frac{1}{1-R(X)g_{1}}\,\mathcyr{sh}_{\hbar}\,w\frac{1}{1-g_{1}X} =ρR(X)(w11g1X)11R(X)g1\displaystyle=\rho_{R(X)}\left(w\frac{1}{1-g_{1}X}\right)\frac{1}{1-R(X)g_{1}}
=ρR(X)(w)(111R(X)g1ebXg1X)1(1R(X)g1)1\displaystyle=\rho_{R(X)}(w)\left(1-\frac{1}{1-R(X)g_{1}}e^{\hbar bX}g_{1}X\right)^{-1}\left(1-R(X)g_{1}\right)^{-1}
=ρR(X)(w)(1(R(X)+ebXX)g1)1.\displaystyle=\rho_{R(X)}(w)\left(1-(R(X)+e^{\hbar bX}X)g_{1}\right)^{-1}.

Similarly, we have

11g1Xshw11R(X)g1\displaystyle\frac{1}{1-g_{1}X}\,\mathcyr{sh}_{\hbar}\,w\frac{1}{1-R(X)g_{1}} =ρX(w11R(X)g1)11g1X\displaystyle=\rho_{X}\left(w\frac{1}{1-R(X)g_{1}}\right)\frac{1}{1-g_{1}X}
=ρX(w)(111g1XR(X)(1+bX)g1)1(1g1X)1\displaystyle=\rho_{X}(w)\left(1-\frac{1}{1-g_{1}X}R(X)(1+\hbar bX)g_{1}\right)^{-1}(1-g_{1}X)^{-1}
=ρX(w)(1(R(X)+ebXX)g1)1.\displaystyle=\rho_{X}(w)\left(1-(R(X)+e^{\hbar bX}X)g_{1}\right)^{-1}.

Note that 𝔫\mathfrak{n} is a two-sided ideal of 0^\widehat{\mathfrak{H}^{0}} with respect to the concatenation product. Hence, to prove Lemma 5.9, it suffices to show that

ρR(X)(g𝐤)ρX(E𝐤)\displaystyle\rho_{R(X)}(g_{\mathbf{k}})\equiv\rho_{X}(E_{\mathbf{k}})

modulo 𝔫[[X]]\mathfrak{n}[[X]] for any non-empty admissible index 𝐤\mathbf{k}.

First we calculate ρR(X)(g𝐤)\rho_{R(X)}(g_{\mathbf{k}}). For k1k\geq 1, we have

ρR(X)(gk)\displaystyle\rho_{R(X)}(g_{k}) =ρR(X)(ba)(ρR(X)(a))k1\displaystyle=\rho_{R(X)}(ba)\left(\rho_{R(X)}(a)\right)^{k-1}
=11R(X)g1ebXg1(11R(X)g1(a+R(X)g1))k1\displaystyle=\frac{1}{1-R(X)g_{1}}e^{\hbar bX}g_{1}\left(\frac{1}{1-R(X)g_{1}}(a+\hbar R(X)g_{1})\right)^{k-1}
11R(X)g1ebXg1(11R(X)g1a)k1\displaystyle\equiv\frac{1}{1-R(X)g_{1}}e^{\hbar bX}g_{1}\left(\frac{1}{1-R(X)g_{1}}a\right)^{k-1}

modulo 0^[[X]]\hbar\,\widehat{\mathfrak{H}^{0}}[[X]]. Moreover,

11R(X)g1a=a+n1(R(X)g1)n1R(X)g2a+n1(Xg1)n1Xg2=11g1Xa\displaystyle\frac{1}{1-R(X)g_{1}}a=a+\sum_{n\geq 1}(R(X)g_{1})^{n-1}R(X)g_{2}\equiv a+\sum_{n\geq 1}(Xg_{1})^{n-1}Xg_{2}=\frac{1}{1-g_{1}X}a

modulo 𝔫0[[X]]\mathfrak{n}_{0}[[X]]. Therefore, it holds that

ρR(X)(gk)11R(X)g1ebXg1(11g1Xa)k1\displaystyle\rho_{R(X)}(g_{k})\equiv\frac{1}{1-R(X)g_{1}}e^{\hbar bX}g_{1}\left(\frac{1}{1-g_{1}X}a\right)^{k-1}

modulo 𝔫[[X]]\mathfrak{n}[[X]] for k1k\geq 1. If k2k\geq 2, each coefficient of the formal power series g1((1g1X)1a)k1g_{1}((1-g_{1}X)^{-1}a)^{k-1} belongs to j2𝒞Agj\sum_{j\geq 2}\mathcal{C}\langle A\rangle g_{j}. For any w𝒞Aw\in\mathcal{C}\langle A\rangle and j2j\geq 2, we see that (b)wgj=(e1g1)wgj𝔫(\hbar b)wg_{j}=(e_{1}-g_{1})wg_{j}\in\mathfrak{n}. Therefore, if 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) is a non-empty admissible index, it holds that

ρR(X)(g𝐤)1ir(11g1Xg1(11g1Xa)ki1)\displaystyle\rho_{R(X)}(g_{\mathbf{k}})\equiv\prod_{1\leq i\leq r}^{\curvearrowright}\left(\frac{1}{1-g_{1}X}\,g_{1}\left(\frac{1}{1-g_{1}X}a\right)^{k_{i}-1}\right)

modulo 𝔫[[X]]\mathfrak{n}[[X]], where 1irAi=A1Ar\prod_{1\leq i\leq r}^{\curvearrowright}A_{i}=A_{1}\cdots A_{r} denotes the ordered product.

Next we calculate ρX(E𝐤)\rho_{X}(E_{\mathbf{k}}). For k1k\geq 1, we have

ρX(ek)\displaystyle\rho_{X}(e_{k}) =ρX(b(a+))(ρX(a))k1=11g1Xe1(11g1X(a+g1X))k1\displaystyle=\rho_{X}(b(a+\hbar))(\rho_{X}(a))^{k-1}=\frac{1}{1-g_{1}X}\,e_{1}\left(\frac{1}{1-g_{1}X}(a+\hbar g_{1}X)\right)^{k-1}
11g1Xe1(11g1Xa)k1\displaystyle\equiv\frac{1}{1-g_{1}X}\,e_{1}\left(\frac{1}{1-g_{1}X}a\right)^{k-1}

modulo 0^[[X]]\hbar\,\widehat{\mathfrak{H}^{0}}[[X]]. Note that

(e1g1)11g1Xa=(e1g1)(1+11g1Xg1X)a=g1+(e1g1)11g1Xg2X,\displaystyle(e_{1}-g_{1})\frac{1}{1-g_{1}X}a=(e_{1}-g_{1})\left(1+\frac{1}{1-g_{1}X}g_{1}X\right)a=\hbar g_{1}+(e_{1}-g_{1})\frac{1}{1-g_{1}X}g_{2}X,

which belongs to 𝔫[[X]]\mathfrak{n}[[X]]. Since 𝔫a𝔫\mathfrak{n}a\subset\mathfrak{n}, it holds that

(B.9) ρX(ek)11g1Xg1(11g1Xa)k1\displaystyle\rho_{X}(e_{k})\equiv\frac{1}{1-g_{1}X}g_{1}\left(\frac{1}{1-g_{1}X}a\right)^{k-1}

modulo 𝔫[[X]]\mathfrak{n}[[X]] for k2k\geq 2. Now we set 𝐤=(k1,,kr)\mathbf{k}=(k_{1},\ldots,k_{r}) and 𝐤=(k1,,kr1)\mathbf{k}^{\prime}=(k_{1},\ldots,k_{r-1}). Since 𝐤\mathbf{k} is admissible, it holds that ρX(E𝐤)=ρX(E𝐤)ρX(ekr)\rho_{X}(E_{\mathbf{k}})=\rho_{X}(E_{\mathbf{k}^{\prime}})\rho_{X}(e_{k_{r}}), and each coefficient of ρX(ekr)\rho_{X}(e_{k_{r}}) belongs to j2𝒞Agj\sum_{j\geq 2}\mathcal{C}\langle A\rangle g_{j} because of (B.9). Therefore, we may calculate ρX(E𝐤)\rho_{X}(E_{\mathbf{k}^{\prime}}) modulo (𝒞A(e1g1)𝒞A)[[X]](\mathcal{C}\langle A\rangle(e_{1}-g_{1})\mathcal{C}\langle A\rangle)[[X]]. Then we see that

ρX(E(Y))(111g1XY)1\displaystyle\rho_{X}(E(Y))\equiv\left(1-\frac{1}{1-g_{1}X}Y\right)^{-1}

modulo (𝒞A(e1g1)𝒞A)[[X]](\mathcal{C}\langle A\rangle(e_{1}-g_{1})\mathcal{C}\langle A\rangle)[[X]], and hence

ρX(E1m)(11g1Xg1)m\displaystyle\rho_{X}(E_{1^{m}})\equiv\left(\frac{1}{1-g_{1}X}g_{1}\right)^{m}

for m0m\geq 0. As a result we find that

ρX(E𝐤)1ir(11g1Xg1(11g1Xa)ki1)\displaystyle\rho_{X}(E_{\mathbf{k}})\equiv\prod_{1\leq i\leq r}^{\curvearrowright}\left(\frac{1}{1-g_{1}X}\,g_{1}\left(\frac{1}{1-g_{1}X}a\right)^{k_{i}-1}\right)

modulo 𝔫[[X]]\mathfrak{n}[[X]]. This completes the proof of Lemma 5.9.

B.5. Proof of Proposition 6.4

We set

e(X)=k1ekXk1=e111aX,e0(X)=k2ekXk2=e211aX=e(X)a.\displaystyle e(X)=\sum_{k\geq 1}e_{k}X^{k-1}=e_{1}\frac{1}{1-aX},\quad e^{0}(X)=\sum_{k\geq 2}e_{k}X^{k-2}=e_{2}\frac{1}{1-aX}=e(X)a.

For w,ww,w^{\prime}\in\mathfrak{H}, we set Ξ(w,w)\Xi(w,w^{\prime})

Ξ(w,w)=wshwE1+wE1shw,\displaystyle\Xi(w,w^{\prime})=w\,\mathcyr{sh}_{\hbar}\,w^{\prime}E_{1}+wE_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime},

where E1=(e1+g1)/2E_{1}=(e_{1}+g_{1})/2. Note that Ξ(w,w)\Xi(w,w^{\prime}) is symmetric with respect to ww and ww^{\prime}, and

(B.10) we1shwe1=Ξ(w,w)e1.\displaystyle we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e_{1}=\Xi(w,w^{\prime})\,e_{1}.

We also set

(w,w)=wg1shw(wshw)g1=we1shw(wshw)e1,\displaystyle\partial(w,w^{\prime})=wg_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}-(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})g_{1}=we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}-(w\,\mathcyr{sh}_{\hbar}\,w^{\prime})e_{1},
ΛY(w,w)=(w11R(Y)g1shw)(1R(Y)g1)=(wE(Y)shw)E(Y)1.\displaystyle\Lambda_{Y}(w,w^{\prime})=\left(w\frac{1}{1-R(Y)g_{1}}\,\mathcyr{sh}_{\hbar}\,w^{\prime}\right)\left(1-R(Y)g_{1}\right)=\left(wE(Y)\mathcyr{sh}_{\hbar}\,w^{\prime}\right)E(Y)^{-1}.
Proposition B.8.

Set

E(Y1,Y2)=1Y1Y2j=12Yj(E(Y1+Y2)E(Yj)).\displaystyle E(Y_{1},Y_{2})=\frac{1}{Y_{1}Y_{2}}\sum_{j=1}^{2}Y_{j}\left(E(Y_{1}+Y_{2})-E(Y_{j})\right).

It holds that

Ξ(wE(Y1),wE(Y2))\displaystyle\Xi(wE(Y_{1}),w^{\prime}E(Y_{2})) =(wE(Y2),w)E(Y1)+(wE(Y1),w)E(Y2)\displaystyle=\partial(w^{\prime}E(Y_{2}),w)E(Y_{1})+\partial(wE(Y_{1}),w^{\prime})E(Y_{2})
+(wshw+ΛY1(w,w)+ΛY2(w,w))E(Y1,Y2).\displaystyle+\left(-w\,\mathcyr{sh}_{\hbar}w^{\prime}+\Lambda_{Y_{1}}(w,w^{\prime})+\Lambda_{Y_{2}}(w^{\prime},w)\right)E(Y_{1},Y_{2}).

for any w,ww,w^{\prime}\in\mathfrak{H}.

Proof.

Set

P(Y1,Y2)=w11R(Y1)g1shw11R(Y2)g1.\displaystyle P(Y_{1},Y_{2})=w\frac{1}{1-R(Y_{1})g_{1}}\,\mathcyr{sh}_{\hbar}\,w^{\prime}\frac{1}{1-R(Y_{2})g_{1}}.

Since E1=b/2+baE_{1}=\hbar b/2+ba, we have

wE(Y1)shwE(Y2)E1\displaystyle wE(Y_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}E(Y_{2})E_{1} =P(Y1,Y2)12Y1Y2bR(Y1)R(Y2)\displaystyle=P(Y_{1},Y_{2})\frac{1}{2Y_{1}Y_{2}}\hbar bR(Y_{1})R(Y_{2})
+(w11R(Y1)g1shwE(Y2)ba)(1R(Y1)g1)E(Y1).\displaystyle+\left(w\frac{1}{1-R(Y_{1})g_{1}}\,\mathcyr{sh}_{\hbar}\,w^{\prime}E(Y_{2})ba\right)(1-R(Y_{1})g_{1})E(Y_{1}).

We calculate the second term of the right hand side by using Proposition B.2 and

(B.11) R(Y1)+R(Y2)+bR(Y1)R(Y2)=R(Y1+Y2).\displaystyle R(Y_{1})+R(Y_{2})+\hbar bR(Y_{1})R(Y_{2})=R(Y_{1}+Y_{2}).

Then we obtain

{(wE(Y2),w)+P(Y1,Y2)1Y2(R(Y1+Y2)R(Y1))g1}E(Y1).\displaystyle\left\{\partial(w^{\prime}E(Y_{2}),w)+P(Y_{1},Y_{2})\frac{1}{Y_{2}}(R(Y_{1}+Y_{2})-R(Y_{1}))g_{1}\right\}E(Y_{1}).

By changing www\leftrightarrow w^{\prime} and Y1Y2Y_{1}\leftrightarrow Y_{2}, we obtain a similar formula for wE(Y1)E1shwE(Y2)wE(Y_{1})E_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}E(Y_{2}). Thus we get

Ξ(wE(Y1),wE(Y2))=(wE(Y2),w)E(Y1)+(wE(Y1),w)E(Y2)+1Y1Y2P(Y1,Y2)Q(Y1,Y2),\displaystyle\Xi(wE(Y_{1}),w^{\prime}E(Y_{2}))=\partial(w^{\prime}E(Y_{2}),w)E(Y_{1})+\partial(wE(Y_{1}),w^{\prime})E(Y_{2})+\frac{1}{Y_{1}Y_{2}}P(Y_{1},Y_{2})Q(Y_{1},Y_{2}),

where

Q(Y1,Y2)=bR(Y1)R(Y2)+j=12Yj(R(Y1+Y2)R(Yj))g1E(Yj).\displaystyle Q(Y_{1},Y_{2})=\hbar bR(Y_{1})R(Y_{2})+\sum_{j=1}^{2}Y_{j}(R(Y_{1}+Y_{2})-R(Y_{j}))g_{1}E(Y_{j}).

From Proposition B.1 and (B.11), we see that

P(Y1,Y2)=(wshw+ΛY1(w,w)+ΛY2(w,w))11R(Y1+Y2)g1.\displaystyle P(Y_{1},Y_{2})=\left(-w\,\mathcyr{sh}_{\hbar}w^{\prime}+\Lambda_{Y_{1}}(w,w^{\prime})+\Lambda_{Y_{2}}(w^{\prime},w)\right)\frac{1}{1-R(Y_{1}+Y_{2})g_{1}}.

Using (B.11) we see that

Q(Y1,Y2)=R(Y1+Y2)(1+g1j=12YjE(Yj))j=12R(Yj)(1+Yjg1E(Yj)).\displaystyle Q(Y_{1},Y_{2})=R(Y_{1}+Y_{2})\left(1+g_{1}\sum_{j=1}^{2}Y_{j}E(Y_{j})\right)-\sum_{j=1}^{2}R(Y_{j})(1+Y_{j}g_{1}E(Y_{j})).

Because R(Yj)(1+Yjg1E(Yj))=YjE(Yj)R(Y_{j})(1+Y_{j}g_{1}E(Y_{j}))=Y_{j}E(Y_{j}), we have

Q(Y1,Y2)=R(Y1+Y2)+(R(Y1+Y2)g11)j=12YjE(Yj)=(1R(Y1+Y2)g1)Y1Y2E(Y1,Y2).\displaystyle Q(Y_{1},Y_{2})=R(Y_{1}+Y_{2})+(R(Y_{1}+Y_{2})g_{1}-1)\sum_{j=1}^{2}Y_{j}E(Y_{j})=(1-R(Y_{1}+Y_{2})g_{1})Y_{1}Y_{2}E(Y_{1},Y_{2}).

Thus we get the desired formula. ∎

Lemma B.9.

For any w,ww,w^{\prime}\in\mathfrak{H}, the following formulas hold.

(B.12) (w,we0(X))=(Ξ(w,w)+Xwshwe0(X))e0(X),\displaystyle\partial(w,w^{\prime}e^{0}(X))=\left(\Xi(w,w^{\prime})+Xw\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)\right)e^{0}(X),
(B.13) ΛY(w,we0(X))=wshwe0(X)+Y(wE(Y),we0(X)).\displaystyle\Lambda_{Y}(w,w^{\prime}e^{0}(X))=w\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)+Y\partial(wE(Y),w^{\prime}e^{0}(X)).
Proof.

Note that e0(X)=e(X)ae^{0}(X)=e(X)a and (w,wa)=(we1shw)a\partial(w,w^{\prime}a)=(we_{1}\,\mathcyr{sh}_{\hbar}w^{\prime})a for any w,ww,w^{\prime}\in\mathfrak{H}. Hence

(B.14) (w,we0(X))=(w,we(X)a)=(we1shwe(X))a.\displaystyle\partial(w,w^{\prime}e^{0}(X))=\partial(w,w^{\prime}e(X)a)=(we_{1}\,\mathcyr{sh}_{\hbar}w^{\prime}e(X))a.

Since e(X)=e1+e(X)aXe(X)=e_{1}+e(X)aX, we see that

we1shwe(X)\displaystyle we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X) =Ξ(w,w)e1+Xwe1shwe(X)a\displaystyle=\Xi(w,w^{\prime})e_{1}+Xwe_{1}\,\mathcyr{sh}_{\hbar}w^{\prime}e(X)a
=Ξ(w,w)e1+X{(wshwe(X)a)e1+(we1shwe(X))a}\displaystyle=\Xi(w,w^{\prime})e_{1}+X\left\{(w\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X)a)e_{1}+(we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X))a\right\}
=(Ξ(w,w)e1+Xwshwe0(X))e1+(we1shwe(X))aX.\displaystyle=\left(\Xi(w,w^{\prime})e_{1}+Xw\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)\right)e_{1}+(we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X))aX.

Therefore we have

we1shwe(X)=(Ξ(w,w)e1+Xwshwe0(X))e(X)\displaystyle we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X)=\left(\Xi(w,w^{\prime})e_{1}+Xw\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)\right)e(X)

since e1(1aX)1=e(X)e_{1}(1-aX)^{-1}=e(X). Thus we get (B.12).

Next we prove (B.13). Using e0(X)=e(X)ae^{0}(X)=e(X)a and Proposition B.2, we see that

ΛY(w,we0(X))\displaystyle\Lambda_{Y}(w,w^{\prime}e^{0}(X)) =wshwe0(X)(wsh,we(X))a+(w11R(Y)g1(1+bR(Y))shwe(X))a\displaystyle=w\,\mathcyr{sh}_{\hbar}w^{\prime}e^{0}(X)-(w\,\mathcyr{sh}_{\hbar}^{,}w^{\prime}e(X))a+\left(w\frac{1}{1-R(Y)g_{1}}(1+\hbar bR(Y))\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X)\right)a
=wshwe0(X)+(w11R(Y)g1R(Y)e1shwe(X))a.\displaystyle=w\,\mathcyr{sh}_{\hbar}w^{\prime}e^{0}(X)+\left(w\frac{1}{1-R(Y)g_{1}}R(Y)e_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X)\right)a.

The second term of the right hand side is equal to

Y(wE(Y)e1shwe(X))a=Y(wE(Y),we0(X))\displaystyle Y\left(wE(Y)e_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X)\right)a=Y\partial(wE(Y),w^{\prime}e^{0}(X))

because of (B.14). Thus we get (B.13). ∎

Proposition B.10.

For any w,ww,w^{\prime}\in\mathfrak{H}, it holds that

wE(Y)shwe0(X)\displaystyle wE(Y)\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)
={wshwe0(X)+YΞ(wE(Y),w)e0(X)}E(Y)11XYe0(X)E(Y).\displaystyle=\left\{w\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)+Y\Xi(wE(Y),w^{\prime})e^{0}(X)\right\}E(Y)\frac{1}{1-XYe^{0}(X)E(Y)}.
Proof.

We denote the left hand side by J(X,Y)J(X,Y). We see that

J(X,Y)\displaystyle J(X,Y) =ΛY(w,we0(X))E(Y)\displaystyle=\Lambda_{Y}(w,w^{\prime}e^{0}(X))E(Y)
={wshwe0(X)+Y(Ξ(wE(Y),w)+XJ(X,Y))e0(X)}E(Y)\displaystyle=\left\{w\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X)+Y\left(\Xi(wE(Y),w^{\prime})+XJ(X,Y)\right)e^{0}(X)\right\}E(Y)

using (B.12) and (B.13). It implies the desired equality. ∎

Proposition B.11.

For any w,ww,w^{\prime}\in\mathfrak{H}, it holds that

we0(X1)shwe0(X2)\displaystyle we^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X_{2})
=1X1X2{(X1+X2+X1X2)Ξ(w,w)e2\displaystyle=\frac{1}{X_{1}X_{2}}\biggl{\{}-(X_{1}+X_{2}+\hbar X_{1}X_{2})\Xi(w,w^{\prime})e_{2}
+X1(1+X2)(w,we0(X2))+X2(1+X1)(w,we0(X1))}\displaystyle\qquad\qquad\qquad{}+X_{1}(1+\hbar X_{2})\,\partial(w,w^{\prime}e^{0}(X_{2}))+X_{2}(1+\hbar X_{1})\,\partial(w^{\prime},we^{0}(X_{1}))\biggr{\}}
×11(X1+X2+X1X2)a.\displaystyle{}\times\frac{1}{1-(X_{1}+X_{2}+\hbar X_{1}X_{2})a}.
Proof.

Since e0(X)=(e(X)e1)/Xe^{0}(X)=(e(X)-e_{1})/X, we have

X1X2we0(X1)shwe0(X2)=w(e(X1)e1)shw(e(X2)e1).\displaystyle X_{1}X_{2}\,we^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X_{2})=w(e(X_{1})-e_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}(e(X_{2})-e_{1}).

Calculate we(X1)shwe(X2)we(X_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X_{2}) using Proposition B.1, and we see that

X1X2(we0(X1)shwe0(X2))(1(X1+X2+X1X2)a)\displaystyle X_{1}X_{2}\,(we^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X_{2}))(1-(X_{1}+X_{2}+\hbar X_{1}X_{2})a)
=(X1+X2+X1X2)(we1shwe1)a\displaystyle=-(X_{1}+X_{2}+\hbar X_{1}X_{2})(we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e_{1})a
+X1(1+X2)(we1shwe(X2))a+X2(1+X1)(we(X1)shwe1)a.\displaystyle+X_{1}(1+\hbar X_{2})(we_{1}\,\mathcyr{sh}_{\hbar}\,w^{\prime}e(X_{2}))a+X_{2}(1+\hbar X_{1})(we(X_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}e_{1})a.

Now the desired equality follows from (B.10) and (B.14). ∎

Now we prove Proposition 6.4. For a subset 𝔞\mathfrak{a} of \mathfrak{H} and a formal power series f(X1,,Xn)[[X1,,Xn]]f(X_{1},\ldots,X_{n})\in\mathfrak{H}[[X_{1},\ldots,X_{n}]], we say that f(X1,,Xn)f(X_{1},\ldots,X_{n}) belongs to 𝔞\mathfrak{a} if all the coefficients of f(X1,,Xn)f(X_{1},\ldots,X_{n}) belong to 𝔞\mathfrak{a}. We show that

  1. (i)

    Ξ(w,w)\Xi(w,w^{\prime}) belongs to 𝔢\mathfrak{e},

  2. (ii)

    wshwe0(X)w\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X) and we0(X)shwwe^{0}(X)\,\mathcyr{sh}_{\hbar}\,w^{\prime} belong to 𝔢\mathfrak{e},

  3. (iii)

    we0(X1)shwe0(X2)we^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,w^{\prime}e^{0}(X_{2}) belongs to 𝔢0\mathfrak{e}^{0}

for any w,w𝔢w,w^{\prime}\in\mathfrak{e}. Proposition 6.4 follows from (ii) and (iii). The third property (iii) follows from (i) and (ii) because of Proposition B.11 and (B.12). Hence it suffices to prove (i) and (ii). Note that the 𝒞\mathcal{C}-module 𝔢\mathfrak{e} is generated by the coefficients of the formal power series

Cr=Cr(X1,,Xr;Y0,,Yr)=E(Y0)1jr(e0(Xj)E(Yj))\displaystyle C_{r}=C_{r}(X_{1},\ldots,X_{r};Y_{0},\ldots,Y_{r})=E(Y_{0})\prod_{1\leq j\leq r}^{\curvearrowright}\left(e^{0}(X_{j})E(Y_{j})\right)

with r0r\geq 0. Hence we may assume that w=Crw=C_{r} and w=Csw^{\prime}=C_{s}^{\prime} for some r,s0r,s\geq 0, where the prime symbol indicates that ww^{\prime} contains a different family of variables from ww. We proceed the proof of (i) and (ii) by induction on r+sr+s.

First we consider the case of r=0r=0. From Proposition B.8, we see that Ξ(E(Y1),E(Y2))=E(Y1,Y2)\Xi(E(Y_{1}),E(Y_{2}))=E(Y_{1},Y_{2}) belongs to 𝔢\mathfrak{e}, and it implies that E(Y1)shE(Y2)e0(X)E(Y_{1})\,\mathcyr{sh}_{\hbar}\,E(Y_{2})e^{0}(X) also belongs to 𝔢\mathfrak{e} because of Proposition B.10. Hence (i) and (ii) are true in the case of r=s=0r=s=0. Now we consider the case where r=0r=0 and s1s\geq 1. Set w=C0=E(Y1)w=C_{0}=E(Y_{1}) and w=Cs=Cs1e0(X2)E(Y2)w^{\prime}=C_{s}^{\prime}=C_{s-1}^{\prime}e^{0}(X_{2})E(Y_{2}). Proposition B.8 and Proposition B.10 imply that

Ξ(C0,Cs)\displaystyle\Xi(C_{0},C_{s}^{\prime}) =(C0,Cs1e0(X2))E(Y2)+ΛY1(1,Cs1e0(X2))E(Y1,Y2),\displaystyle=\partial(C_{0},C_{s-1}^{\prime}e^{0}(X_{2}))E(Y_{2})+\Lambda_{Y_{1}}(1,C_{s-1}^{\prime}e^{0}(X_{2}))E(Y_{1},Y_{2}),
C0shCse0(X)\displaystyle C_{0}\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime}e^{0}(X) ={Cs+Y1Ξ(C0,Cs)}e0(X)E(Y1)11XY1e0(X)E(Y1),\displaystyle=\left\{C_{s}^{\prime}+Y_{1}\,\Xi(C_{0},C_{s}^{\prime})\right\}e^{0}(X)E(Y_{1})\frac{1}{1-XY_{1}e^{0}(X)E(Y_{1})},
C0e0(X)shCs\displaystyle C_{0}e^{0}(X)\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime} ={C0e0(X)shCs1e0(X2)+Y2Ξ(C0,Cs)e0(X)}E(Y2)\displaystyle=\left\{C_{0}e^{0}(X)\,\mathcyr{sh}_{\hbar}\,C_{s-1}^{\prime}e^{0}(X_{2})+Y_{2}\,\Xi(C_{0},C_{s}^{\prime})e^{0}(X)\right\}E(Y_{2})
×11XY2e0(X)E(Y2).\displaystyle{}\times\frac{1}{1-XY_{2}e^{0}(X)E(Y_{2})}.

From Lemma B.9 and the induction hypothesis, we see that (C0,Cs1e0(X2))\partial(C_{0},C_{s-1}^{\prime}e^{0}(X_{2})) and ΛY1(1,Cs1e0(X2))\Lambda_{Y_{1}}(1,C_{s-1}^{\prime}e^{0}(X_{2})) belong to 𝔢0\mathfrak{e}^{0}. Hence Ξ(C0,Cs)\Xi(C_{0},C_{s}^{\prime}) belongs to 𝔢\mathfrak{e}, and it implies that C0shCse0(X)C_{0}\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime}e^{0}(X) belongs to 𝔢\mathfrak{e}. Moreover, since C0e0(X)shCs1e0(X2)C_{0}e^{0}(X)\,\mathcyr{sh}_{\hbar}\,C_{s-1}^{\prime}e^{0}(X_{2}) belongs to 𝔢0\mathfrak{e}^{0} because of the induction hypothesis (iii), we see that C0e0(X)shCsC_{0}e^{0}(X)\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime} belongs to 𝔢\mathfrak{e}. Thus we obtain (i), (ii) with w=C0w=C_{0} and w=Csw^{\prime}=C_{s}^{\prime} for any s0s\geq 0. Since Ξ(w,w)\Xi(w,w^{\prime}) is symmetric with respect to ww and ww^{\prime}, they are also true in the case where w=Crw=C_{r} with r0r\geq 0 and w=C0w^{\prime}=C_{0}^{\prime}.

Next we consider the case where r,s1r,s\geq 1. Set w=Cr=Cr1e0(X1)E(Y1)w=C_{r}=C_{r-1}e^{0}(X_{1})E(Y_{1}) and w=Cs=Cs1e0(X2)E(Y2)w^{\prime}=C_{s}^{\prime}=C_{s-1}^{\prime}e^{0}(X_{2})E(Y_{2}). From Proposition B.8, Proposition B.10 and (B.13), we see that

Ξ(Cr,Cs)\displaystyle\Xi(C_{r},C_{s}^{\prime}) =(Cr,Cs1e0(X2)){E(Y1)+Y2E(Y1,Y2)}\displaystyle=\partial(C_{r},C_{s-1}^{\prime}e^{0}(X_{2}))\left\{E(Y_{1})+Y_{2}E(Y_{1},Y_{2})\right\}
+(Cs,Cr1e0(X1)){E(Y2)+Y1E(Y1,Y2)}\displaystyle+\partial(C_{s}^{\prime},C_{r-1}e^{0}(X_{1}))\left\{E(Y_{2})+Y_{1}E(Y_{1},Y_{2})\right\}
+(Cr1e0(X1)shCs1e0(X2))E(Y1,Y2)\displaystyle+\left(C_{r-1}e^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,C_{s-1}^{\prime}e^{0}(X_{2})\right)E(Y_{1},Y_{2})

and

CrshCse0(X)={Cr1e0(X1)shCse0(X)+Y1Ξ(Cr,Cs)e0(X)}E(Y1)11XY1e0(X)E(Y1).\displaystyle C_{r}\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime}e^{0}(X)=\left\{C_{r-1}e^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime}e^{0}(X)+Y_{1}\,\Xi(C_{r},C_{s}^{\prime})e^{0}(X)\right\}E(Y_{1})\frac{1}{1-XY_{1}e^{0}(X)E(Y_{1})}.

From (B.12) and the induction hypothesis, we see that (Cr,Cs1e0(X2)),(Cs,Cr1e0(X1))\partial(C_{r},C_{s-1}^{\prime}e^{0}(X_{2})),\partial(C_{s}^{\prime},C_{r-1}e^{0}(X_{1})) and Cr1e0(X1)shCs1e0(X2)C_{r-1}e^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,C_{s-1}^{\prime}e^{0}(X_{2}) belong to 𝔢0\mathfrak{e}^{0}. Hence Ξ(Cr,Cs)\Xi(C_{r},C_{s}^{\prime}) belongs to 𝔢\mathfrak{e}. The induction hypothesis says that Cr1e0(X1)shCse0(X)C_{r-1}e^{0}(X_{1})\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime}e^{0}(X) belongs to 𝔢0\mathfrak{e}^{0}, and hence CrshCse0(X)C_{r}\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime}e^{0}(X) belongs to 𝔢\mathfrak{e}. Similarly we find that Cre0(X)shCsC_{r}e^{0}(X)\,\mathcyr{sh}_{\hbar}\,C_{s}^{\prime} belongs to 𝔢\mathfrak{e}. Therefore (i) and (ii) hold for w=Crw=C_{r} and w=Csw^{\prime}=C_{s}, and this completes the proof of Proposition 6.4.

References

  • [1] Cartier, P., On the double zeta values, in Galois-Teichmüller theory and arithmetic geometry, 91–119, Adv. Stud. Pure Math., 63, Math. Soc. Japan, Tokyo, 2012.
  • [2] Hoffman, E., The algebra of multiple harmonic series, J. Algebra 194 (1997), no. 2, 477–495.
  • [3] Ihara, K., Kaneko, M. and Zagier, D., Derivation and double shuffle relations for multiple zeta values, Compos. Math. 142 (2006), no. 2, 307–338.
  • [4] Jarossay, D., Adjoint cyclotomic multiple zeta values and cyclotomic multiple harmonic values, preprint (2019), arXiv:1412.5099v5.
  • [5] Kaneko, M., An introduction to classical and finite multiple zeta values, Publications mathématiques de Besançon. Algébre et théorie des nombres. 2019/1, 103–129, Publ. Math. Besançon Algébre Théorie Nr., 2019/1, Presses Univ. Franche-Comté, Besançon, 2020.
  • [6] Ono, M., Seki, S. and Yamamoto, S., Truncated tt-adic symmetric multiple zeta values and double shuffle relations, Res. Number Theory 7 (2021), no. 1, Paper No. 15, 28 pp. DOI: 10.1007/s40993-021-00241-5.
  • [7] Oyama, K., Ohno-type relation for finite multiple zeta values, Kyushu J. Math. 72 (2018), no. 2, 277–285.
  • [8] Reutenauer, C., Free Lie algebras, Oxford Science Publications, 1993.
  • [9] Takeyama, Y., The algebra of a qq-analogue of multiple harmonic series, Symmetry, Integrability and Geometry: Methods and Applications (SIGMA) 9 (2013), 061, 15 pages, DOI: 10.3842/SIGMA.2013.061.
  • [10] Zagier, D., Evaluation of the multiple zeta values ζ(2,,2,3,2,,2)\zeta(2,\ldots,2,3,2,\ldots,2), Ann. of Math. (2) 175 (2012), no. 2, 977–1000.
  • [11] Zhao, J., Multiple zeta functions, multiple polylogarithms and their special values, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2016.