A -analogue of symmetric multiple zeta value
Abstract.
We construct a -analogue of truncated version of symmetric multiple zeta values which satisfies the double shuffle relation. Using it, we define a -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 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) |
for a tuple of positive integers with . We set and regard it as a MZV. We denote by the -linear subspace of spanned by the MZVs. It is known that forms a -algebra with respect to the usual multiplication on .
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 . The purpose of this paper is to construct a -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 , we have two kinds of its regularization, which are called the harmonic regularized MZV and the shuffle regularized MZV (see Section 2.2 for the details). If , they are equal to the MZV . Using them we set
for . It is known that the difference belongs to the ideal for any tuple of positive integers. The SMZV is defined by
as an element of the quotient .
Now we fix a complex parameter satisfying and define the -integer for by . There are various models of a -analogue of the MZV (see, e.g., [11, Chapter 12]). Most of them are of the following form:
(1.2) |
where is a polynomial whose degree is less than or equal to for . If , the infinite sum (1.2) is absolutely convergent. In this paper we call a value of the form (1.2) a -analogue of the MZV (MZV) without specifying the model.
To construct a -analogue of SMZVs, it would be natural to consider two kinds of regularizations of the sum (1.2) with which turn into the harmonic regularized MZV and the shuffle regularized MZV in the limit as . 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 -analogue of SMZVs in a different approach. In [6], Ono, Seki and Yamamoto introduced two kinds of truncations of the -adic symmetric multiple zeta value (-adic SMZV), which is an element of the formal power series ring whose constant term is equal to the SMZV. As a corollary, we have the following expression for . Set
(1.3) | ||||
(1.4) |
Then we have for [6, Corollary 2.7]. Note that we do not need any regularization here. Hence we could follow the above procedure to construct a -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 -adic SMZVs satisfy a generalization of the double shuffle relation [4]. The point is that the truncated version of the -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 -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 -analogue of SMZVs as a limit of the above truncated version. As SMZVs are defined to be the elements of the quotient , our -analogue of SMZVs is an element of the quotient of the space of the MZVs by a sum of two ideals . Roughly speaking, is the ideal generated by the MZVs which turn into zero in the limit as , and is the ideal generated by the values with , which turns into in the limit as . Hence the ideal could be regarded as a -analogue of to the ideal .
Unlike the truncated SMZVs, the double shuffle relation of the truncated SMZVs does not imply that of the SMZVs. It is because not all of the truncated SMZVs converge in the limit as and the space spanned by them which converge is not closed under the shuffle product (see Example 6.3 below). Hence our -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 MZVs. In Section 4 we define a -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 -analogue of SMZVs. In Section 6 we see that our -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 MZVs in the limit as 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 , we define its weight and depth by and , respectively. We call a tuple of positive integers an index. We regard the empty set as an index whose weight and depth are zero. An index is said to be admissible if , or and . We denote the set of indices (resp. admissible indices) by (resp. ). The reversal of an index is defined by . For the empty index we set .
In this paper we often make use of generating functions in proofs. Then we use the operations defined below without mention. Suppose that is a unital algebra over a commutative ring . Then we extend the addition and the multiplication on to the formal power series ring by and for and with , respectively. Similarly, we extend a -linear map to the -linear map by for . 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 and a positive integer , we define the truncated multiple zeta value by
if is not empty, and for the empty index. If , we set .
For an admissible index , we define the multiple zeta value (MZV) by . If is a non-empty admissible index, we have
which converges since .
Let be the non-commutative polynomial ring of two variables and over . We set for , and for a non-empty index . For the empty index we set .
Set . It is a -subalgebra of , which can be identified with the non-commutative polynomial ring over with the set of variables . We also set , which is a -submodule of with a basis .
The harmonic product on is the -bilinear map defined by the following properties:
-
(i)
For any , it holds that and .
-
(ii)
For any and , it holds that .
The shuffle product is similarly defined by the following properties and -linearity:
-
(i)
For any , it holds that and .
-
(ii)
For any and , it holds that .
Then (resp. ) becomes a commutative -algebra with respect to the harmonic (resp. shuffle) product, which we denote by (resp. ). We see that is a -subalgebra of with respect to the harmonic product. We denote it by . We also see that and are -subalgebras of with respect to the shuffle product. We denote them by and , respectively.
For a positive integer , we define the -linear map by for an index . Similarly, we define the -linear map by , for an admissible index .
Proposition 2.1.
-
(i)
For any and , it holds that
Hence, if and belong to , we have
(2.1) -
(ii)
For any , it holds that
(2.2)
2.2. Symmetric multiple zeta value
It is known that for (see [2, 8]). Hence, for , there uniquely exists the -algebra homomorphism from to the polynomial ring such that . For an index , we define and . It is known that
(2.3) |
in the formal power series ring for any index [3, Theorem 1 and Proposition 10].
For an index and , we set
From (2.3), we see that the difference belongs to the ideal of . The symmetric MZV (SMZV) is defined by
as an element of .
2.3. Truncated symmetric multiple zeta value
We define the -algebra anti-automorphism on by for . For , we define the -linear map by and
for and .
For a positive integer and , we define the -linear map by . For an index and a positive integer , the value is expressed as follows:
where and in the condition are set equal to zero in the latter formula. We call the above values the truncated symmetric multiple zeta value.
2.4. Relations of SMZVs
From the definition of , we obtain the following equality, which is called the reversal relation:
Theorem 2.3.
For any index , it holds that
For indices and , we define the non-negative integer by
(2.4) |
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 and , it holds that
In [7], Oyama proved that Theorem 2.4 and the identity for any imply the Ohno-type relation. To write down it, we define the Hoffman dual of an index . We define the -algebra automorphism on by and . For a non-empty index , the monomial is written in the form with a monomial . Then the Hoffman dual index of is defined to be the index satisfying . For example, if , we have and , hence . For the empty index we set .
Theorem 2.5 (Ohno-type relation [7]).
For any index and any non-negative integer , it holds that
where and .
3. A -analogue of multiple zeta value
3.1. Algebraic formulation
Set , where is a formal variable. Let be the non-commutative polynomial ring of two variables and over . For we set
They are related to each other as
for , and
(3.1) |
for . Note that .
We set
which is an algebraically independent set, and denote by the -subalgebra of generated by and . The depth of a monomial is defined to be .
We define the -submodule of by
Then the set consisting of the elements
(3.2) |
with and forms a free basis of .
Let be a complex parameter satisfying . We endow the complex number field with -module structure such that acts as multiplication by .
We denote by the -submodule of spanned by the set . For a positive integer , we define the -linear map by
(3.3) |
for , where is the -integer
Then, from and (3.1) for , we have
(3.4) |
for any because .
For a positive integer , we define the -linear map by and
for . Then we see that, if , converges in the limit as . Hence we can define the -linear map by
for . In this paper we call the value with a -analogue of MZV (MZV).
3.2. Double shuffle relation of MZVs
The harmonic product and the shuffle product associated with the MZV are defined as follows.
First, we define the symmetric -bilinear map by
for . Then we see that
(3.5) |
for any and . The harmonic product is the -bilinear binary operation on uniquely defined by the following properties:
-
(i)
For any , it holds that and .
-
(ii)
For any and , it holds that .
The harmonic product on is commutative and associative.
Proposition 3.2.
The -submodule of is closed under the harmonic product , and it holds that for any and . Therefore, we have for any .
Next, we define the shuffle product. Consider the -linear right action of on a -valued function defined by
Then, for any function and , it holds that
(3.6) |
if all terms are well-defined, where denotes the usual multiplication of functions. Motivated by the above relations we define the shuffle product on as the -bilinear binary operation uniquely defined by the following properties:
-
(i)
For any , it holds that and .
-
(ii)
For any , it holds that
The shuffle product is commutative and associative. Note that
(3.7) |
for any since .
Proposition 3.3.
The -submodules and of are closed under the shuffle product . For any and , it holds that
(3.8) |
Therefore, we have for any .
Proof.
Note that for . For and , it holds that
(3.9) | ||||
(3.10) | ||||
(3.11) |
These formulas imply that
(3.12) |
for . From (3.7) and (3.12), we see that and are closed under the shuffle product.
Denote by the -vector space of holomorphic functions on the unit disc. We endow with -module structure such that acts as multiplication by . Let be the constant function. We define the -linear map by
where . We see that
for . Hence, is equal to the sum of the coefficients of in over for . By (3.6), we have for any . Thus we obtain (3.8). ∎
3.3. Limit of MZVs as
We denote by the -submodule of spanned by the elements of the form (3.2) with and for some . We define the -modules and by
For a non-empty index , we set
For the empty index, we set . Note that the quotient is a free -module which has a basis .
Proposition 3.4.
Here we consider a limit as with being real.
-
(i)
If is an admissible index, it holds that .
-
(ii)
For any , it holds that .
See Appendix A for the proof.
We define the unital -algebra homomorphism by
(3.14) |
for . Then we see that the restriction of to is a surjection onto and its kernel is equal to . Therefore, from Proposition 3.4, we obtain the following corollary.
Corollary 3.5.
For any , it holds that
(3.15) |
3.4. Restoration of finite double shuffle relation of MZVs
Proposition 3.7.
The -module is an ideal of with respect to both the harmonic product and the shuffle product.
Proof.
Corollary 3.8.
The -module is closed under both the harmonic product and the shuffle product.
Proof.
Proposition 3.9.
For any and , it holds that
(3.16) |
Proof.
We may assume that and are monomials in . If or , the desired equality (3.16) is trivial. Now we proceed the proof by induction on the sum of the depth of and .
Since for any and for , we see that (3.16) holds for .
We consider the case of . If or is the form of with a monomial of , we see that (3.16) from and (3.7). Hence it suffices to prove the case where and with monomials in and . For that purpose we use the following formula. Note that, in the formal power series ring , we have
Lemma 3.10.
For any , we have
4. A -analogue of truncated SMZV
Let be the -algebra anti-involution on defined by
for . For , we define the -linear map by and
for and .
Now we define the -linear map for and by . We call the value with a -analogue of truncated symmetric multiple zeta value.
We prove the double shuffle relation of the -analogue of truncated SMZVs. The proof is similar to that of the truncated SMZVs in [6].
Proposition 4.1.
For any and , it holds that
(4.3) |
Proof.
Note that is an anti-automorphism. We extend the -linear map for by for . Then, from (4.1), we see that
(4.4) |
for , where is Kontsevich’s order on the set defined by .
Proposition 4.2.
For any and , it holds that
(4.5) |
Proof.
We extend the -linear map for by for . Then we see that (3.3) holds for any nonzero integer and .
Let be the -trilinear map defined by and
where , and . The summation region is the subset of consisting of tuples satisfying the following conditions:
Then we see that for . Hence it suffices to show that
(4.6) |
for any .
From (4.2) and the definition of , we see that (4.6) holds if or . Set and . We prove (4.6) by induction on . Since is symmetric with respect to and , it suffices to consider the two cases: (i) , (ii) with . The case (i) follows from (3.7) and for integers and satisfying and . To show the case (ii), we set
It holds that
for non-zero integers and satisfying . From the above relation and Lemma 3.10, we see that (4.6) holds in the case (ii) under the induction hypothesis. ∎
5. A -analogue of SMZV
In this section, we define a -analogue of the SMZV. To this aim, we use the map , which corresponds to , defined as follows.
First we define the map .
Proposition 5.1.
For any index , the element belongs to the -submodule of .
Proof.
Definition 5.2.
Set
which is a -submodule of . We define the -linear map by
for an index . More explicitly, we have
for a non-empty index .
We want to define the map similarly by taking the limit of as . For that purpose, however, we should determine the domain carefully because it is not closed under the concatenation product unlike the -module in Definition 5.2 as seen by the following example.
Example 5.3.
We have and . Hence a -linear combination of and whose image by converges in the limit as should be proportional to . However, we see that
Therefore, the concatenation product of does not belong to the domain of the map . Here we note that, because , it holds that , which clearly converges as .
To describe the domain of the map , we introduce the element of defined by
where and for and . For example, we have
Let be a non-empty index. It can be written uniquely in the form
(5.1) |
with and , where the right hand side reads if . Then we set
For the empty index, we set .
Proposition 5.4.
For any index , the element belongs to .
Definition 5.5.
Set
We define the -linear map by
for an index .
Example 5.6.
If , it holds that for . Hence, for a non-empty index whose all components are larger than one, we have
Using the two maps and , we define a -analogue of the SMZV. Recall that we need the relation (2.3) to define the SMZV. We show the corresponding relation in the -analogue case.
For and satisfying , we define the exponential with respect to the product by
Theorem 5.8.
For any admissible index , it holds that
(5.2) |
in modulo .
Proof.
We start from the following equality. See Appendix B.4 for the proof.
Lemma 5.9.
Set
(5.3) |
For any admissible index , it holds that
(5.4) |
modulo the -submodule of .
We denote by the ideal of generated by the set .
Corollary 5.10.
For any index , the difference belongs to the ideal .
Proof.
Now we are in a position to define a -analogue of the SMZV.
Definition 5.11.
For an index , we define a -analogue of SMZV (SMZV) as an element of the quotient by
modulo .
Example 5.12.
Example 5.13.
We consider the SMZV of depth two. Set . If and are even, then . If and are odd, we have . In both cases we see that belongs to from Proposition 3.2. Therefore if is even.
We consider the case where is odd. To this aim we calculate the MZV whose weight is odd modulo . The calculation is similar to that for MZV in [1, 10].
Suppose that is odd and . For , we have
From Lemma 3.10, we also see that
Since and belongs to if , we have
modulo . Note that or is even. Hence, from Proposition 3.2 and Proposition 3.3, we see that
(5.8) | |||
(5.9) |
modulo . Set
Then (5.8) and (5.9) imply that
modulo , respectively. Using these relations and , we obtain
Hence we find that
(5.10) |
modulo if and is odd.
Now suppose that and is odd. Since or is even, we see that
From the above arguments we see that
in the quotient for any . It is a -analogue of the formula for the SMZV of depth two
modulo (see, e.g., [5]).
We check that our SMZV is really a -analogue of the SMZV.
Theorem 5.14.
For any index , it holds that
Proof.
The limit as of any element of is zero and that of is contained in because for . Therefore, we have the well-defined map which sends the equivalent class of to that of . Theorem 5.14 implies that the map sends the SMZV to the SMZV . In this sense we may regard as a -analogue of .
6. Relations of the -analogue of symmetric multiple zeta value
6.1. Reversal relation
Theorem 6.1.
For any index , we have
and
modulo . Therefore, for any index , it holds that
Proof.
Since is an anti-involution, we see that for any and . From the definition of , we have . We also have modulo (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 SMZVs.
Proposition 6.2.
For any index and , it holds that
Next we consider the shuffle relation. Note that the product does not necessarily belong to as follows.
Example 6.3.
We have
which is not a -linear combination of and .
However, we have the following proposition. We set
Proposition 6.4.
Let and be an index. If at least one of and is admissible, then belongs to . If both and are admissible, then belongs to .
See Appendix B.5 for the proof.
Proposition 6.5.
Let and be an index and suppose that at least one of them is admissible. Then it holds that
modulo , where is the concatenation of and .
From Proposition 6.2 and Proposition 6.5, we obtain the double shuffle relation of SMZVs as follows.
Theorem 6.6.
Let be the non-negative integer defined by (2.4).
-
(i)
For any index and , it holds that
-
(ii)
Moreover, if at least one of and is admissible, it holds that
Proof.
From the definition of the harmonic product , we see that . Hence (i) is true because of Proposition 6.2.
We prove (ii) by using the map defined by (3.14). Proposition 6.4 implies that there exists such that . As shown in the proof of Theorem 5.14, we have for any index . Hence we see that
On the other hand, Proposition 3.9 implies that
Hence and
Therefore we see that (ii) is true from Proposition 6.5 and . ∎
6.3. Ohno-type relation
From Theorem 6.6, we see that the SMZVs satisfy a part of the Ohno-type relation.
Theorem 6.7.
Let be a non-empty admissible index. Then it holds that
(6.1) |
for any , where and .
Proof.
We define the -linear map by for an index . Theorem 6.6 and (5.7) imply the following properties:
-
(i)
For any index and , it holds that .
-
(ii)
For any admissible index and , it holds that .
For a non-empty index and , we define the element of by
where the sum is over the set
We also set
for . Then we have
for any (see equation (3) in [7]).
Appendix A Proof of Proposition 3.4
Lemma A.1.
Suppose that and . Then the function is non-decreasing on the interval .
Proof.
If , the statement is trivial. We assume that . From the assumption we see that . Set
Since and , it suffices to show that on the interval . Set
Then we see that and, from the assumption,
for . Therefore we see that , and hence , on the interval . ∎
Corollary A.2.
Proof.
It is because as . ∎
Proposition A.3.
Suppose that . Set
Then, for any non-negative integer , it holds that
Proof.
From Corollary A.2, we see that
for any . Therefore, it holds that
Set for . Then the right-hand side is dominated by since . ∎
Proposition A.4.
Suppose that and are non-negative integers. Then it holds that
(A.1) |
Moreover, if and for some , then we have
(A.2) |
Proof.
For a non-negative integer , we define
We have
Hence as for any . If , we see that
for . Hence as for any .
Proof of Proposition 3.4.
From Lemma A.1 and the monotone convergence theorem, we see that as for any admissible index . Suppose that is an element of of the form (3.2). Proposition A.3 and (A.1) imply that as . Hence in the limit as . If , then there exist and such that and . Then Proposition A.3 and (A.2) imply that in the limit as . Thus we see that for any . ∎
Appendix B Proofs
B.1. Formulas of shuffle product
Proposition B.1.
Suppose that and belong to the ideal of the formal power series ring generated by . Then it holds that
for any .
Proof.
Set
Using , we see that
Since all the coefficients of and are polynomials in , the fourth term of the right hand side is equal to
Using again, we see that it is equal to
Thus we get the desired equality. ∎
Proposition B.2.
Suppose that . Then it holds that
for any .
Proof.
Set in Proposition B.1, differentiate the both hand sides with respect to and set . Then we get the desired equality. ∎
Proof of Lemma 3.10.
Corollary B.3.
For , we define the map by
Then the map is a -algebra homomorphism with respect to the concatenation product on , and we have
(B.1) |
B.2. Generating function of
Here we calculate the generating function
Recall that
Note that belongs to .
Proposition B.4.
It holds that
(B.2) |
Proof.
The power series is the unique solution of the differential equation satisfying . Since the right hand side with is equal to one, it suffices to show that
The right hand side is equal to . Hence (B.1) implies that
since . ∎
Proposition B.5.
It holds that
Proof.
Note that and are commutative with respect to the shuffle product . From the definition of , we see that
Using
and (B.2), we see that
∎
Corollary B.6.
For any index , it holds that modulo . Moreover, if (resp. ) is admissible, then belongs to (resp. ).
B.3. Proof of Proposition 5.4
Note that
(B.6) |
because and for .
We define the -trilinear map by
for and . It suffices to show that belongs to for any admissible index and .
Lemma B.7.
For any , it holds that
(B.7) |
Proof.
Decompose . Then we obtain
From the definition of , we see that the second term in the right hand side is equal to
Hence, it holds that
Since , we see that the right hand side above is equal to
Therefore, by replacing with , we obtain
Since and , it holds that
Now set and with admissible indices and . We have modulo because of Proposition 3.3, Corollary B.6 and . From the definition of the shuffle product and (B.6), we see that belongs to , hence so does the first term of (B.8) with and . Since and are admissible, we can write and with some because of (B.3). Then, from Proposition B.2, the second term of (B.8) is equal to
which belongs to because of (B.6). Similarly, the third term of (B.8) also belongs to . Thus we find that belongs to , and this completes the proof of Proposition 5.4.
B.4. Proof of Lemma 5.9
We use the map and defined in Corollary B.3 with and , respectively. Let . Since is a -algebra homomorphism and , we have
Similarly, we have
Note that is a two-sided ideal of with respect to the concatenation product. Hence, to prove Lemma 5.9, it suffices to show that
modulo for any non-empty admissible index .
First we calculate . For , we have
modulo . Moreover,
modulo . Therefore, it holds that
modulo for . If , each coefficient of the formal power series belongs to . For any and , we see that . Therefore, if is a non-empty admissible index, it holds that
modulo , where denotes the ordered product.
Next we calculate . For , we have
modulo . Note that
which belongs to . Since , it holds that
(B.9) |
modulo for . Now we set and . Since is admissible, it holds that , and each coefficient of belongs to because of (B.9). Therefore, we may calculate modulo . Then we see that
modulo , and hence
for . As a result we find that
modulo . This completes the proof of Lemma 5.9.
B.5. Proof of Proposition 6.4
We set
For , we set
where . Note that is symmetric with respect to and , and
(B.10) |
We also set
Proposition B.8.
Set
It holds that
for any .
Proof.
Set
Since , we have
We calculate the second term of the right hand side by using Proposition B.2 and
(B.11) |
Then we obtain
By changing and , we obtain a similar formula for . Thus we get
where
From Proposition B.1 and (B.11), we see that
Using (B.11) we see that
Because , we have
Thus we get the desired formula. ∎
Lemma B.9.
For any , the following formulas hold.
(B.12) | |||
(B.13) |
Proof.
Note that and for any . Hence
(B.14) |
Since , we see that
Therefore we have
since . Thus we get (B.12).
Proposition B.10.
For any , it holds that
Proof.
Proposition B.11.
For any , it holds that
Proof.
Now we prove Proposition 6.4. For a subset of and a formal power series , we say that belongs to if all the coefficients of belong to . We show that
-
(i)
belongs to ,
-
(ii)
and belong to ,
-
(iii)
belongs to
for any . 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 -module is generated by the coefficients of the formal power series
with . Hence we may assume that and for some , where the prime symbol indicates that contains a different family of variables from . We proceed the proof of (i) and (ii) by induction on .
First we consider the case of . From Proposition B.8, we see that belongs to , and it implies that also belongs to because of Proposition B.10. Hence (i) and (ii) are true in the case of . Now we consider the case where and . Set and . Proposition B.8 and Proposition B.10 imply that
From Lemma B.9 and the induction hypothesis, we see that and belong to . Hence belongs to , and it implies that belongs to . Moreover, since belongs to because of the induction hypothesis (iii), we see that belongs to . Thus we obtain (i), (ii) with and for any . Since is symmetric with respect to and , they are also true in the case where with and .
Next we consider the case where . Set and . From Proposition B.8, Proposition B.10 and (B.13), we see that
and
From (B.12) and the induction hypothesis, we see that and belong to . Hence belongs to . The induction hypothesis says that belongs to , and hence belongs to . Similarly we find that belongs to . Therefore (i) and (ii) hold for and , 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 -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 -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 , 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.