1 The mod 2 Leibniz–Hopf algebra and its dual

Let \(\mathcal {F}_2\) be the free associative algebra over \(\mathbb {F}_2\) generated by the indeterminates \(S^1, S^2, S^3, \ldots \) of degree \(|S^i|=i\). We often denote the unit 1 by \(S^0\). This algebra is equipped with a co-commutative co-product given by

$$\begin{aligned} \Delta (S^n) = \sum _{i=0}^n S^i \otimes S^{n-i}, \end{aligned}$$
(1)

which makes it a graded connected Hopf algebra. This algebra \(\mathcal {F}_2\) is often called the mod 2 Leibniz–Hopf algebra. As an \(\mathbb {F}_2\)-module, \(\mathcal {F}_2\) has the following canonical basis:

$$\begin{aligned} \{ S^I:=S^{i_1}S^{i_2}\cdots S^{i_n} \mid I=(i_1,i_2,\ldots ,i_n)\in \mathbb {N}^n, 0\le n <\infty \}, \end{aligned}$$

where we regard \(S^I=1\) when \(n=0\).

Note that the integral counterpart of \(\mathcal {F}_2\) is called the Leibniz–Hopf algebra and is isomorphic to the ring of non-commutative symmetric functions [7] and the Solomon Descent algebra [17]. Its graded dual is the ring of quasi-symmetric functions with the outer co-product, which has been studied by Hazewinkel, Malvenuto, and Reutenauer in [8,9,10,11,12].

The mod 2 Steenrod algebra \(\mathcal {A}_2\) is defined to be the quotient Hopf algebra of \(\mathcal {F}_2\) by the ideal generated by the Adem relations:

$$\begin{aligned} S^i S^j - \sum _{k=0}^{\lfloor i/2 \rfloor } {{j-k-1}\atopwithdelims (){i-2k}} S^{i+j-k}S^k. \end{aligned}$$
(2)

Denote the quotient map by \(\pi : \mathcal {F}_2\rightarrow \mathcal {A}_2\) and \(Sq^i=\pi (S^i)\). It is well-known (see, for example, [18]) that the admissible monomials

$$\begin{aligned} \{Sq^J:=Sq^{j_1}Sq^{j_2}\cdots Sq^{j_n} \mid J=(j_1,j_2,\ldots ,j_n)\in \mathbb {N}_{>0}^n, 0\le n <\infty , j_{k-1} \ge 2 j_{k} \forall k\} \end{aligned}$$

form a module basis for \(\mathcal {A}_2\). We will adhere to this purely algebraic definition and will not use any other known facts about \(\mathcal {A}_2\).

By taking the graded dual of \(\pi \), we obtain the following inclusion of Hopf algebras

$$\begin{aligned} \pi ^*: \mathcal {A}_2^* \rightarrow \mathcal {F}_2^*. \end{aligned}$$

\(\mathcal {F}_2^*\) is given a module basis \(S_I\) dual to \(S^I\), that is,

$$\begin{aligned} \langle S^{I'}, S_I \rangle = {\left\{ \begin{array}{ll} 1 &{} (I=I') \\ 0 &{} (I\ne I') \end{array}\right. }. \end{aligned}$$

Similarly, we have the dual basis \(\{Sq_J \mid J \text { admissible} \}\) for \(\mathcal {A}_2^*\) determined by

$$\begin{aligned} \langle Sq^{J'}, Sq_J \rangle = {\left\{ \begin{array}{ll} 1 &{} (J=J') \\ 0 &{} (J\ne J') \end{array}\right. }. \end{aligned}$$

The commutative product among the basis elements in \(\mathcal {F}_2^*\) is given by the overlapping shuffle product (see §2) and the co-product is given by

$$\begin{aligned} \Delta (S_{a_1,\ldots ,a_n})=S_{a_1,\ldots ,a_n} \otimes 1 + 1 \otimes S_{a_1,\ldots ,a_n}+ \sum _{i=1}^{n-1} S_{a_1,\ldots ,a_i} \otimes S_{a_{i+1},\ldots ,a_n}. \end{aligned}$$
(3)

The purpose of this paper is to deduce some of the classical results on \(\mathcal {A}_2^*\) and its generalisations by considering it as a subalgebra of \(\mathcal {F}_2^*\). We are particularly interested in the following problems.

Problem 1

  1. (i)

    Determine the coefficients in

    $$\begin{aligned} \pi ^*(Sq_J) = \sum _{I} C^I_J S_I \end{aligned}$$
    (4)

    for all admissible sequences J. This is important since in the dual it is equivalent to computing the coefficients of the Adem relations

    $$\begin{aligned} Sq^I = \sum _{J:\text {admissible}} C^I_J Sq^J \end{aligned}$$
    (5)

    for all sequences I.

  2. (ii)

    Give an expansion of the dual Milnor bases in terms of the dual admissible monomial bases, i.e., determine the coefficient \(B^L_J\) in

    $$\begin{aligned} \xi ^L = \sum _{J:\text {admissible}} B^L_J Sq_J, \end{aligned}$$

    where \(\xi _n=Sq_{2^{n-1},2^{n-2}\cdots 2^1,2^0}\) and \(\xi ^L=\xi _1^{l_1}\xi _2^{l_2}\cdots \xi _n^{l_n}\) for \(L=(l_1,l_2,\ldots , l_n)\).

  3. (iii)

    Generalise Milnor’s conjugation formula [13] in \(\mathcal {A}_2^*\) to \(\mathcal {F}_2^*\). The formula for \(\mathcal {A}_2^*\) is:

    $$\begin{aligned} \chi (\xi _n)= \sum _\alpha \prod _{i=1}^{l(\alpha )} \xi _{\alpha (i)}^{2^{\sigma (i)}}, \end{aligned}$$

    where \(\alpha =(\alpha (1)|\alpha (2)|\ldots |\alpha (l(\alpha ))\) runs through all the compositions of the integer n and \(\sigma (i)=\sum _{j=1}^{i-1} \alpha (j)\).

Several different methods are known for resolving (i) and (ii) (see for example, [15, 19]), but our argument (Sect. 4) is new in that it is purely combinatorial using the overlapping shuffle product on \(\mathcal {F}_2^*\). We implemented our algorithm into a Maple code [16]. In Sect. 5 we discuss the conjugation (or antipode) in \(\mathcal {F}_2^*\) and give an answer to (iii). Finally, we give an explicit duality between the conjugation invariants in \(\mathcal {F}_2\) and \(\mathcal {F}_2^*\) in Sect. 6.

2 Overlapping Shuffle product

We recall the definition of the overlapping shuffle product ([2, Section 2],[8]). Let \(\mathcal {W}\) be the set of finite sequences of natural numbers:

$$\begin{aligned} \mathcal {W}= \{ (i_1,i_2,\ldots ,i_n) \mid 0\le n <\infty \}. \end{aligned}$$

Note that we allow the length 0 sequence. Consider the \(\mathbb {F}_2\)-module \(\mathbb {F}_2\langle \mathcal {W}\rangle \) freely generated by \(\mathcal {W}\). For a sequence \(I=(i_1,i_2,\ldots ,i_n)\), denote its tail partial sequence \((i_k,i_{k+1},\ldots ,i_n)\) by \(I_k\). When \(n<k\), we regard \(I_k\) as the length 0 sequence. We use the convention

$$\begin{aligned}&(a_1,a_2,\ldots ,a_k,(b_1,\ldots ,b_i)+(c_1,\ldots ,c_j))\\&\quad :=(a_1,a_2,\ldots ,a_k,b_1,\ldots ,b_i) +(a_1,a_2,\ldots ,a_k,c_1,\ldots ,c_j). \end{aligned}$$

The overlapping shuffle product on \(\mathbb {F}_2\langle \mathcal {W}\rangle \) is defined as follows:

Definition 1

For \(A=(a_1,a_2,\ldots ,a_n)\) and \(B=(b_1,b_2,\ldots ,b_m)\), define their product inductively by

$$\begin{aligned} A\cdot B:= {\left\{ \begin{array}{ll} A &{} (m=0) \\ B &{} (n=0) \\ \sum _{0\le i \le n} (a_1,\ldots , a_i, b_1, A_{i+1}\cdot B_2)\\ \quad + \sum _{1\le i \le n} (a_1,\ldots , a_i+b_1, A_{i+1}\cdot B_2) &{} (\text {otherwise}). \end{array}\right. } \end{aligned}$$

The product on \(\mathbb {F}_2\langle \mathcal {W}\rangle \) is defined by the linear extension of the above.

We say a term in \(A\cdot B\) is a-first if there exists k such that \(a_k\) goesFootnote 1 to an entry to the left of \(b_k\) and \(a_i\) goes to the same entry as \(b_i\) (that is, the entry makes \(a_i+b_i\)) for all \(i<k\). For example, \((a_1+b_1,a_2,b_2,b_3,a_3)\) is a-first while \((a_1+b_1,b_2,a_2,a_3,b_3)\) is not. Observe that

Lemma 1

For equal length sequences, we have

$$\begin{aligned} (a_1,\ldots ,a_n) \cdot (b_1,\ldots ,b_n)= (a_1+b_1,\ldots ,a_n+b_n) + Z+ \tau (Z), \end{aligned}$$

where Z is a sum of a-first terms and \(\tau \) flips the occurrence of \(a_i\) and \(b_i\) for all i. In particular, the product is commutative.

Example 1

$$\begin{aligned} (a_1,a_2)\cdot (b_1,b_2)&=(a_1+b_1,a_2+b_2) \\&\quad +(a_1+b_1,a_2,b_2)+(a_1,b_1,a_2+b_2)+(a_1,b_1+a_2,b_2)\\&\quad +(a_1,a_2,b_1,b_2)+(a_1,b_1,a_2,b_2)+(a_1,b_1,b_2,a_2) \\&\quad +(b_1+a_1,b_2,a_2)+(b_1,a_1,b_2+a_2)+(b_1,a_1+b_2,a_2)\\&\quad +(b_1,b_2,a_1,a_2)+(b_1,a_1,b_2,a_2)+(b_1,a_1,a_2,b_2), \end{aligned}$$

where the second line consists of a-first terms and the third line is the \(\tau \)-image of the second line.

Corollary 1

For \(A=(a_1,\ldots ,a_n)\),

$$\begin{aligned} A\cdot A=(2a_1,\ldots ,2a_n), \quad A^{2^m}=(2^m a_1,\ldots ,2^m a_n). \end{aligned}$$

Proof

In this case, the flip map \(\tau \) in Lemma 1 is the identity. \(\square \)

It is easy to see from the duality relation \(\langle S_I S_J, S^K \rangle =\langle S_I\otimes S_J, \Delta (S^K) \rangle \) that the product on \(\mathcal {F}_2^*\) dual to (1) is given by \(S_I S_J= \sum _{K\in I\cdot J} S_K\).

3 Dual Steenrod algebra as a sub-Hopf algebra of \(\mathcal {F}_2^*\)

To identify the image of the inclusion \(\pi ^*:\mathcal {A}_2^* \rightarrow \mathcal {F}_2^*\), we prove some lemmas in this section. Let \(\xi _n=Sq_{2^{n-1}, 2^{n-2}, \ldots , 2^0}\).

Lemma 2

(cf. [2, 19]) We have

$$\begin{aligned} \pi ^*(Sq_{2^n})= & {} S_{2^n}, \\ \pi ^*(\xi _n)= & {} S_{2^{n-1}, 2^{n-2}, \ldots , 2^0}. \end{aligned}$$

Proof

For the first equation, we have to show that for any non-admissible sequence I, the right-hand side of

$$\begin{aligned} Sq^I = \sum _{J:\text {admissible}} C^I_J Sq^J \end{aligned}$$

does not contain \(Sq^{2^n}\). If there exists such an I, we can assume it has length two, that is, \(I=(i,j)\). (Because the right-hand side is obtained by successively applying the length two relations.) By the Adem relations in Eq. (2), we have \(i+j=2^n\) and

$$\begin{aligned} 1\equiv \left( {\begin{array}{c}j-1\\ i\end{array}}\right) \equiv \left( {\begin{array}{c}2^n-1-i\\ i\end{array}}\right) \mod 2. \end{aligned}$$

However, the binary expressions of \(2^n-1-i\) and i are complementary and the binary expression of \(2^n-1-i\) contains at least one digit with 0. Hence, by Lucas’ Theorem, we have \(\left( {\begin{array}{c}2^n-1-i\\ i\end{array}}\right) \equiv 0 \mod 2\); we arrive at a contradiction.

For the second equation, suppose that there exists an \(I=(i,j)\) such that \(i < 2j\) and

$$\begin{aligned} Sq^{i,j} = \sum _{k=0}^{\lfloor i/2 \rfloor } {{j-k-1}\atopwithdelims (){i-2k}} Sq^{i+j-k} Sq^k \end{aligned}$$

contains \(Sq^{2^{n-k}}\) or \(Sq^{2^{n-k}}Sq^{2^{n-k-1}}\) as a summand. The former case is already ruled out by the first equation. For the latter case to happen, we should have

$$\begin{aligned} i+j=2^{n-k}+2^{n-k-1}, \quad \lfloor i/2 \rfloor \ge 2^{n-k-1}. \end{aligned}$$

But this implies \(j\le 2^{n-k-1}\) so \(i\ge 2j\); we arrive at a contradiction. \(\square \)

Put \(\bar{\xi }_n=\pi ^*(\xi _n) = S_{2^{n-1}, 2^{n-2}, \ldots , 2^0}\). We denote by \(\widetilde{\mathcal {A}}_2^*\) the subalgebra of \(\mathcal {F}_2^*\) generated by \(\{\bar{\xi }_n \mid 0<n \}\). For a sequence \(L=(l_1,l_2,\ldots ,l_n)\) of non-negative integers, we denote \(\bar{\xi }_1^{l_1} \bar{\xi }_2^{l_2} \cdots \bar{\xi }_n^{l_n}\) by \(\bar{\xi }^L\). Then, the monomials \(\bar{\xi }^L\) span \(\widetilde{\mathcal {A}}_2^*\). Now, we identify \(\widetilde{\mathcal {A}}_2^*\) with \(\mathrm {Im}(\pi ^*)\).

Recall the definition of the excess vector of an admissible sequence \(J=(j_1,j_2,\ldots ,j_n)\):

$$\begin{aligned} \gamma (j_1,j_2,\ldots ,j_n)=(j_1-2j_2, j_2-2j_3,\ldots ,j_{n-1}-2j_n,j_n). \end{aligned}$$

This gives a bijection between admissible sequences and sequences of non-negative integers. The inverse is given by

$$\begin{aligned} \gamma ^{-1}(l_1,l_2,\ldots ,l_n) = (l_1+2l_2+2^2l_3+\cdots +2^{n-1}l_n,\ldots ,l_{n-1}+2l_n,l_n). \end{aligned}$$

We put the right lexicographic order on \(\mathcal {W}\), i.e.,

$$\begin{aligned} (a_1,a_2,\ldots ,a_n)> (b_1,b_2,\ldots ,b_m) \Leftrightarrow (n>m) \text { or } (\exists k, a_k>b_k \text { and } a_i=b_i \forall i>k). \end{aligned}$$

This induces an ordering on the basis elements \(S_I\) which is compatible with the overlapping shuffle product. Observe that the lowest term in the product \(S_I \cdot S_{I'}\) for \(I=(i_1,i_2,\ldots )\) and \(I'=(i'_1,i'_2,\ldots )\) is \(S_{(i_1+i'_1,i_2+i'_2,\ldots )}\).

Lemma 3

For an admissible sequence J,

$$\begin{aligned} \langle \bar{\xi }^{\gamma (J)}, S^I \rangle = {\left\{ \begin{array}{ll} 1 &{} (I=J) \\ 0 &{} (I<J). \end{array}\right. } \end{aligned}$$

Proof

We proceed by induction on \(J=(j_1,\ldots ,j_n)\). Put \(J'=(j_1-2^{n-1},j_2-2^{n-2},\ldots ,j_n-2^0)\). Then by induction hypothesis,

$$\begin{aligned} \bar{\xi }^{\gamma (J')}=S_{J'} + (\text {terms higher than } S_{J'}). \end{aligned}$$

It follows that

$$\begin{aligned} \bar{\xi }^{\gamma (J)}= & {} \bar{\xi }^{\gamma (J')} \cdot \bar{\xi }_n\\= & {} (S_{J'} + (\text {terms higher than } S_{J'}))\cdot S_{2^{n-1},2^{n-2},\ldots ,2^0} \\= & {} S_{J} + (\text {terms higher than } S_J). \end{aligned}$$

\(\square \)

By this upper-triangularity, the monomials \(\bar{\xi }^L\) are linearly independent and we have

Theorem 1

$$\begin{aligned} \mathrm {Im}(\pi ^*) = \widetilde{\mathcal {A}}_2^* = \mathbb {F}_2[\bar{\xi }_1,\bar{\xi }_2,\ldots ,]. \end{aligned}$$

Proof

By Lemma 3 in each degree \(\widetilde{\mathcal {A}}_2^*\) has the same dimension as \(\mathcal {A}_2^*\) (the number of admissible sequences). \(\square \)

This is nothing but the well-known fact:

Corollary 2

[13]

$$\begin{aligned} \mathcal {A}_2^*=\mathbb {F}_2[\xi _1,\xi _2,\ldots ,], \end{aligned}$$

where

$$\begin{aligned} \xi ^{\gamma (J)}= Sq_J + (\text {terms higher than}\; Sq_J). \end{aligned}$$

4 Computation with \(\pi ^*\)

Recall from [19, Section 4] the linear left inverse \(r:\mathcal {F}_2^* \rightarrow \mathcal {A}_2^*\) of \(\pi ^*\):

$$\begin{aligned} r(S_I)={\left\{ \begin{array}{ll} Sq_I &{} (I:\text {admissible}) \\ 0 &{}(\text {otherwise}). \end{array}\right. } \end{aligned}$$

For (ii) of Problem 1, we can compute

$$\begin{aligned} \xi ^{(l_1,l_2,\ldots , l_n)}= & {} r\pi ^*(\xi ^{(l_1,l_2,\ldots ,l_n)})\nonumber \\= & {} r({\bar{\xi }}_{1}^{l_1}{\bar{\xi }}_{2}^{l_2}\cdots {\bar{\xi }}_{n}^{l_n})\nonumber \\= & {} r((S_{2^0})^{l_1} (S_{2^1, 2^0})^{l_2}\cdots (S_{2^{n-1}, 2^{n-2}, \ldots , 2^0})^{l_n}) \end{aligned}$$
(6)

and it reduces to computing admissible sequences occurring in the overlapping shuffle product.

For (i) of Problem 1, by Corollary 2 we have

$$\begin{aligned} \pi ^*(\xi ^{\gamma (J)})=\pi ^*(Sq_J + (\text {terms higher than} \,Sq_J)) \end{aligned}$$

and the left-hand side can be computed by the overlapping shuffle product. Thus, we can compute inductively the coefficients \(C^I_J\) in

$$\begin{aligned} \pi ^*(Sq_J) = \sum _{I} C^I_J S_I. \end{aligned}$$

We implemented the algorithm into a Maple code [16].

Example 2

We demonstrate the above algorithm in low degrees. First, compute \(\pi ^*\)-image of monomials \(\xi ^L\):

$$\begin{aligned} \pi ^*(\xi _2^2)&= S_{2, 1} S_{2, 1}=S_{4, 2}\\ \pi ^*(\xi _1^3\xi _2)&=(S_3+S_{1, 2}+S_{2, 1})S_{2, 1}\\&=S_{5, 1}+S_{4, 2}+S_{3, 3}+S_{2, 4}+S_{2, 3, 1}+S_{1, 4, 1}\\&\quad +\,S_{3, 1, 2}+S_{2, 2, 2}+S_{1, 2, 3}+S_{2, 1, 2, 1}+S_{1, 2, 1, 2} \\ \pi ^*(\xi _1^6)&= S_{6}+S_{4,2}+S_{2,4}. \end{aligned}$$

Taking r on the both sides of equations, we obtain

$$\begin{aligned} \xi _2^2 =Sq_{4,2},\quad \xi _1^3\xi _2 =Sq_{5,1}+Sq_{4,2}, \quad \xi _1^6=Sq_6+Sq_{4,2}. \end{aligned}$$

Again taking \(\pi ^*\) on the both sides of the equations, we obtain

$$\begin{aligned} \pi ^*(Sq_{4, 2})&= S_{4, 2} \\ \pi ^*(Sq_{5, 1}+Sq_{4, 2})&= S_{5, 1}+S_{4, 2}+S_{3, 3}+S_{2, 4}+S_{2, 3, 1}+S_{1, 4, 1}\\&\quad + S_{3, 1, 2}+S_{2, 2, 2}+S_{1, 2, 3}+S_{2, 1, 2, 1}+S_{1, 2, 1, 2} \\ \pi ^*(Sq_6+Sq_{4,2})&= S_{6}+S_{4,2}+S_{2,4}. \end{aligned}$$

Finally, by using the upper-triangularity, we obtain

$$\begin{aligned} \pi ^*(Sq_{4,2})&= S_{4,2} \\ \pi ^*(Sq_{5,1})&= S_{5, 1}+S_{3, 3}+S_{2, 4}+S_{2, 3, 1}+S_{1, 4, 1}+S_{3, 1, 2}+S_{2, 2, 2}\\&\quad +S_{1, 2, 3}+S_{2, 1, 2, 1}+S_{1, 2, 1, 2} \\ \pi ^*(Sq_{6})&= S_6+S_{2,4}. \end{aligned}$$

5 Formula for the conjugation

Any connected commutative or co-commutative Hopf algebra has a unique conjugation \(\chi \) satisfying

$$\begin{aligned} \chi (1)=1, \quad \chi (xy)=\chi (y)\chi (x), \quad \chi ^2(x)=x, \quad \sum x'\chi (x'')=0, \end{aligned}$$

where \(\Delta (x)=\sum x'\otimes x''\) and \(\deg (x)>0\) [14]. The conjugation invariants in \(\mathcal {A}_2^*\) is studied in [5] because it is relevant to the commutativity of ring spectra [1, Lecture 3]. The same problem in \(\mathcal {F}_2^*\) has been also studied in [3, 4]. Here we investigate them through our point of view.

Since \(\pi ^*\) is a Hopf algebra homomorphism, we have \(\pi ^*\circ \chi _{\mathcal {A}_2^*} = \chi _{\mathcal {F}_2^*}\circ \pi ^*\), where \(\chi _{\mathcal {A}_2^*}\) and \(\chi _{\mathcal {F}_2^*}\) denote the conjugation operations in \(\mathcal {A}_2^*\) and \(\mathcal {F}_2^*\) respectively. For the module basis \(S_I\) in \(\mathcal {F}_2^*\), the conjugation \(\chi _{\mathcal {F}_2^*}\) is calculated combinatorially.

Definition 2

The coarsening set C(I) of a sequence \(I=(i_1,\ldots ,i_l)\) is defined recursively as

$$\begin{aligned} C(I) := \{ (i_1, I'), (i_1+i'_1,I'_2) \mid I'\in C((i_2,\ldots ,i_l))\} \quad \text { and } \quad C((i))=\{ (i) \}, \end{aligned}$$

where \(I'_2\) is the tail partial sequence \((i'_2,\ldots ,i'_{l'})\) of \(I'=(i'_1,i'_2,\ldots ,i'_{l'})\).

Example 3

\(C((a,b,c))=\{ (a,b,c), (a+b,c), (a,b+c), (a+b+c) \}\).

A formula for the conjugation operation in the dual Leibniz–Hopf algebra is given by Ehrenborg [6, Proposition 3.4]. We now give a simple proof for its mod 2 reduction.

Proposition 1

$$\begin{aligned} \chi _{\mathcal {F}_2^*}(S_I)= \sum _{I'\in C(I^{-1})} S_{I'}, \end{aligned}$$

where \(I^{-1}=(i_l,\ldots ,i_1)\) is the reverse sequence of \(I=(i_1,\ldots ,i_l)\).

Proof

The conjugation is uniquely characterised by

$$\begin{aligned} \chi _{\mathcal {F}_2^*}(1)=1, \quad \sum x'\chi _{\mathcal {F}_2^*}(x'')=0, \end{aligned}$$

where \(\Delta (x)=\sum x'\otimes x''\) and \(\deg (x)>0\). We put \(\chi '(S_I)=\sum _{I'\in C(I^{-1})} S_{I'}\) and show that it satisfies the above equations. It is obvious that \(\chi '(1)=1\). Since the co-product is given in (3), the second equation reads

$$\begin{aligned} \sum _{k=0}^{l} S_{i_1,\ldots ,i_k} \chi '(S_{i_{k+1},\ldots ,i_n})=0 \qquad (\forall I=(i_1,i_2,\ldots ,i_n)). \end{aligned}$$

We regard an element of \(\mathbb {F}_2\langle \mathcal {W}\rangle \) with a finite subset of \(\mathcal {W}\) in the obvious way. We investigate relation between coarsening and the overlapping shuffle product. Define

$$\begin{aligned} C_k(I) =\sum _{ I' \in C((I_{k+1})^{-1})} I' \cdot (i_1, \ldots , i_k). \end{aligned}$$

We observeFootnote 2 that \(C(I^{-1}) \subset C_1(I)\) and \(C'_1(I):=C_1(I) {\setminus } C(I^{-1})\) consists of those sequences that \(i_1\) appears to the left of \(i_2\). In turn, \(C'_1(I) \subset C_2(I)\) and \(C'_2(I):=C_2(I) {\setminus } C'_1(I)\) consists of those sequences that \(i_2\) appears to the left of \(i_3\). Continuing similarly, we obtain

$$\begin{aligned} C(I^{-1}) =\sum _{k=1}^l C_k(I). \end{aligned}$$

It follows that

$$\begin{aligned} \chi '(S_I)=\sum _{k=1}^l \sum _{I' \in C((I_{k+1})^{-1})} S_{(i_1, \ldots , i_k)}\cdot S_{I'} =\sum _{k=0}^{l} S_{i_1,\ldots ,i_k} \chi '(S_{i_{k+1},\ldots ,i_n})-\chi '(S_{I}) \end{aligned}$$

and \(\sum _{k=0}^{l} S_{i_1,\ldots ,i_k} \chi '(S_{i_{k+1},\ldots ,i_n})=0\). \(\square \)

We give another formula for \(\chi _{\mathcal {F}_2^*}(S_I)\).

Definition 3

For a sequence \(a_1,a_2,\ldots ,a_n\), the set of ordered block partitions \(\mathcal {P}(a_1,a_2,\ldots ,a_n)\) consists of elements of the form

$$\begin{aligned} \beta =((a_1,a_2,\ldots ,a_{i_1}) | (a_{i_1+1},\ldots ,a_{i_2}) | \ldots | (a_{i_{l-1}+1},\ldots ,a_{i_l})), \end{aligned}$$

where \(1\le i_1<i_2<\cdots <i_l=n\). We denote \(l(\beta )=l\) and \(\beta (k)=(a_{i_{k-1}+1},\ldots ,a_{i_k})\). Or inductively, we can define

$$\begin{aligned} \mathcal {P}(a_1,a_2,\ldots ,a_n) = \bigcup _{k=1}^n \bigg \{ ((a_1,\ldots ,a_k)| \beta ) \bigg |\beta \in \mathcal {P}(a_{k+1},a_{k+2},\ldots ,a_n) \bigg \}. \end{aligned}$$
(7)

Theorem 2

$$\begin{aligned} \chi _{\mathcal {F}_2^*}(S_I)=\sum _{\beta \in \mathcal {P}(I)} \prod _{k=1}^{l(\beta )} S_{\beta (k)}. \end{aligned}$$

Proof

Let \(I=(a_1,a_2,\ldots ,a_n)\). Put

$$\begin{aligned} \chi '(S_{a_1,a_2,\ldots ,a_n})= \sum _{\beta \in \mathcal {P}(a_1,a_2,\ldots ,a_n)} \prod _{k=1}^{l(\beta )} S_{\beta (k)} \end{aligned}$$

and we check that

$$\begin{aligned} \chi '(1)=1,\quad \sum _{k=0}^{n} S_{a_1,\ldots ,a_k} \chi '(S_{a_{k+1},\ldots ,a_n})=0. \end{aligned}$$

Then, by the uniqueness of the conjugation, we have \(\chi _{\mathcal {F}_2^*}=\chi '\). The first assertion is trivial. For the second, observe that by (7)

$$\begin{aligned} \chi '(S_{a_1,a_2,\ldots ,a_n})= & {} \sum _{k=1}^n S_{a_1,\ldots ,a_k} \left( \sum _{\beta \in \mathcal {P}(a_{k+1},a_{k+2},\ldots ,a_n)} \prod _{j=1}^{l(\beta )} S_{\beta (j)} \right) \\= & {} \sum _{k=1}^n S_{a_1,\ldots ,a_k} \chi '(S_{a_{k+1},\ldots ,a_n}). \end{aligned}$$

Hence, we have

$$\begin{aligned} \sum _{k=0}^{n} S_{a_1,\ldots ,a_k} \chi '(S_{a_{k+1},\ldots ,a_n})&= \chi '(S_{a_1,a_2,\ldots ,a_n})+\sum _{k=1}^{n} S_{a_1,\ldots ,a_k} \chi '(S_{a_{k+1},\ldots ,a_n}) \\&= 2\chi '(S_{a_1,a_2,\ldots ,a_n}) \\&=0. \end{aligned}$$

\(\square \)

Example 4

$$\begin{aligned} \chi _{\mathcal {F}_2^*}(S_{1,2,3})&= S_{3,2,1}+S_{5,1}+S_{3,3}+S_{6} \\&= S_{1,2,3}+S_1 S_{2,3} +S_{1,2}S_3 + S_1 S_2 S_3. \end{aligned}$$

The first line is computed by Proposition 1, and the second by Theorem 2.

Theorem 2 can be thought of as a generalisation of Milnor’s conjugation formula in \(\mathcal {A}_2^*\). To see this, we first show a small lemma:

Lemma 4

$$\begin{aligned} \bar{\xi }_n^{2^m}=(S_{2^{n-1}, 2^{n-2}, \ldots , 2^0})^{2^m}=S_{2^{n+m-1}, 2^{n+m-2}, \ldots , 2^m}. \end{aligned}$$

Proof

This is a direct consequence of Corollary 1 combined with Lemma 2. \(\square \)

Corollary 3

[13, Lemma 10]

$$\begin{aligned} \chi _{\mathcal {A}_2^*}(\xi _n)= \sum _\alpha \prod _{k=1}^{l(\alpha )} \xi _{\alpha (k)}^{2^{\sigma (k)}}, \end{aligned}$$
(8)

where \(\alpha =(\alpha (1)|\alpha (2)|\ldots |\alpha (l(\alpha ))\) runs through all the compositions of the integer n and \(\sigma (k)=\sum _{j=1}^{k-1} \alpha (j)\).

Proof

We apply the injection \(\pi ^*\) to the both sides of (8) and show that they coincide. For the left-hand side, by Lemma 4 we have

$$\begin{aligned} \pi ^*(\chi _{\mathcal {A}_2^*}(\xi _n))=\chi _{\mathcal {F}_2^*}(\pi ^*(\xi _n))= \chi _{\mathcal {F}_2^*}(S_{2^{n-1},2^{n-2},\ldots ,2^0}). \end{aligned}$$

Since \(\pi ^*(\xi _{\alpha (k)}^{2^{\sigma (k)}})=S_{2^{\alpha (k)+\sigma (k)-1},\ldots ,2^{\sigma (k)}}\) by Lemma 4, we see

$$\begin{aligned} \pi ^* \left( \prod _{k=1}^{l(\alpha )} \xi _{\alpha (k)}^{2^{\sigma (k)}} \right) = S_{2^{n-1},\ldots ,2^{n-\alpha (l(\alpha ))}} \cdot S_{2^{n-1-\alpha (l(\alpha ))},\ldots ,2^{n-\alpha (l(\alpha ))-\alpha (l(\alpha )-1)}} \cdots S_{2^{\alpha (1)-1},\ldots ,2^0}. \end{aligned}$$

So when \(\alpha \) ranges over all compositions of n, we get all the ordered block partitions of the sequence \(2^{n-1},2^{n-2},\ldots ,2^0\). The assertion follows from Theorem 2. \(\square \)

6 Duality between \(\mathcal {F}_2\) and \(\mathcal {F}_2^*\)

In the previous section we discussed how to compute the conjugation in \(\mathcal {F}_2^*\). Here, we relate the conjugation in \(\mathcal {F}_2\) with that in \(\mathcal {F}_2^*\) by using a self-duality of \(\mathcal {W}\). Denote \(I \preceq I'\) if \(I\in C(I')\). We think of \(I\in \mathcal {W}\) as a string of 1’s separated by ‘\(+\)’ and commas; \((\underbrace{1+1+\cdots +1}_{i_1}, \underbrace{1+1+\cdots +1}_{i_2},\ldots ,\underbrace{1+1+\cdots +1}_{i_l})\).

Definition 4

We define the dual \(\bar{I}\in \mathcal {W}\) of I by switching \(+\) and the commas.

Example 5

For \(I=(1,3,2)=(1,1+1+1,1+1)\), its dual is

$$\begin{aligned} \bar{I}=(1+1,1,1+1,1)=(2,1,2,1). \end{aligned}$$

It is easily seen that \(\bar{\bar{I}}=I\) and \(I \preceq I' \Leftrightarrow \bar{I} \succeq \bar{I}'\). Extend the duality to one between \(\mathcal {F}_2\) and \(\mathcal {F}_2^*\) by

$$\begin{aligned} D(S^I) = S_{\bar{I}}, \quad D^{-1}(S_I) = S^{\bar{I}}. \end{aligned}$$

Theorem 3

We have \(D\circ \chi _{\mathcal {F}_2}= \chi _{\mathcal {F}_2^*}\circ D\). In particular, \(f \in \mathcal {F}_2\) is a conjugation invariant if and only if so is \(\bar{f} \in \mathcal {F}_2^*\).

Proof

We compute

$$\begin{aligned} D^{-1}\circ \chi _{\mathcal {F}_2^*}\circ D ( S^I )&= D^{-1} \chi _{\mathcal {F}_2^*}(S_{\bar{I}}) = D^{-1} \left( \sum _{I' \preceq (\bar{I})^{-1}} S_{I'}\right) = D^{-1} \left( \sum _{\bar{I}' \succeq I^{-1}} S_{I'}\right) \\&= \sum _{\bar{I}'\succeq I^{-1}} S^{\bar{I}'} = \sum _{I'\succeq I^{-1}} S^{I'}. \end{aligned}$$

Put \(\chi '(S^I)=\sum _{I'\succeq I^{-1}} S^{I'}\). Then, one can check \(\chi '(1)=1\) and \(\sum x' \chi '(x'')=0\) for \(\Delta x= \sum x'\otimes x''\) as in Proposition 1. Hence, by the uniqueness of the conjugation, we have \(\chi '=\chi _{\mathcal {F}_2}\). \(\square \)

Example 6

\(f=S^{1,1,2}+S^{2,1,1}+S^{1,1,1,1}\) is a \(\chi _{\mathcal {F}_2}\)-invariant, whilst \(D(f)=S_{3,1}+S_{1,3}+S_4\) is a \(\chi _{\mathcal {F}_2^*}\)-invariant.

Remark 1

The sub-module of the conjugation invariants in \(\mathcal {F}_2\) is \(\ker (\chi _{\mathcal {F}_2}-1)\) and that in \(\mathcal {F}_2^*\) is \(\ker (\chi _{\mathcal {F}_2^*}-1)\). The conjugations in \(\mathcal {F}_2\) and \(\mathcal {F}_2^*\) are dual to each other, and hence, the linear map \(\chi _{\mathcal {F}_2}-1\) is transpose to \(\chi _{\mathcal {F}_2^*}-1\) with the kernel of same dimension [3]. Theorem 3 gives more information by specifying an explicit correspondence between their elements.