Previous Article in Journal
The Generalized Mehler–Fock Transform over Lebesgue Spaces
Previous Article in Special Issue
Homotopy Perturbation Method for Pneumonia–HIV Co-Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Finite Multiple Mixed Values

Department of Mathematics, The Bishop’s School, La Jolla, CA 92037, USA
Foundations 2024, 4(3), 451-467; https://doi.org/10.3390/foundations4030029
Submission received: 30 June 2024 / Revised: 29 August 2024 / Accepted: 3 September 2024 / Published: 6 September 2024

Abstract

:
In recent years, a variety of multiple zeta values (MZVs) variants have been defined and studied. One way to produce these variants is to restrict the indices in the definition of MZVs to some fixed parity pattern, which include Hoffman’s multiple t-values, Kaneko and Tsumura’s multiple T-values, and Xu and this paper’s author’s multiple S-values. Xu and this paper’s author have also considered the so-called multiple mixed values by allowing all possible parity patterns and have studied a few important relations among these values. In this paper, we turn to the finite analogs and the symmetric forms of the multiple mixed values, motivated by a deep conjecture of Kaneko and Zagier, which relates the finite MZVs and symmetric MZVs, and a generalized version of this conjecture by the author to the Euler sum (i.e., level two) setting. We present a few important relations among these values such as the stuffle, reversal, and linear shuffle relations. We also compute explicitly the (conjecturally smallest) generating set in weight one and two cases. In the appendix, we tabulate some dimension computations for various subspaces of the finite multiple mixed values and propose a conjecture.

1. Introduction

1.1. Multiple Zeta Values and Their Finite Analogs

For any composition of positive integers s = ( s 1 , , s d ) N d , we define the multiple zeta value (MZV) by
ζ ( s ) : = n 1 > > n d > 0 j = 1 d 1 n j s j
and the multiple zeta star value (MZSV) by
ζ ( s ) : = n 1 n d 1 j = 1 d 1 n j s j .
These converge if and only if s 1 2 , in which case we say that s is admissible. As usual, we call | s | : = s 1 + + s d the weight and d the depth. These values were first systematically studied by Zagier [1] and Hoffman [2] independently.
In recent years, a lot of research has been carried out concerning the structure of the different variants of multiple zeta values due to their important applications in both mathematics and theoretical physics. For example, when we allow alternating signs to appear, we then obtain the so-called Euler sums (also called alternating MZVs), which play important roles in the study of knot theory and Witten multiple zeta functions associated with Lie algebra (see, e.g., [3,4]). If we allow not only ± 1 but, more generally, Nth roots of unity, then we can consider colored/cyclotomic MZVs of level N (see [5]), which have also appeared unexpectedly in the study of Feynman diagrams [6].
On the other hand, the modular arithmetic nature of the partial sums of MZVs was first considered by Hoffman [7] and the author of this paper [8] independently. Contrary to the classical cases above, not many variants of these sums exist. To set up the correct theoretical framework for these variants, we define the following adéle-like ring, which was first considered by Kontsevich and then applied to the p-adic setting by Kaneko and Zagier [9]. Let P be the set of primes. Set
A : = p P ( Z / p Z ) / p P ( Z / p Z ) .
Then, we can define the finite multiple zeta values (FMZVs) by the following:
ζ A ( s ) : = p > n 1 > > n d > 0 j = 1 d 1 n j s j ( mod p ) p P A .
In 2014, Kaneko and Zagier proposed a deep conjecture (see Conjecture 1 below for a generalization), relating these values on the p-adic side to MZVs on the Archimedean side via a mysterious connection. This conjecture is far from being proven, but, since then, a plethora of parallel results have been shown to hold on both sides simultaneously (see, e.g., [10,11,12,13,14,15]). In particular, for each positive integer w 2 , the element
β w : = B p w w w < p P A
is the finite analog of ζ ( w ) , where B n s are the Bernoulli numbers defined by
t e t 1 = n 0 B n t n n ! .
And the Fermat quotient
q 2 : = 2 p 1 1 p 2 < p P A
is the analog of ζ ( 1 ¯ ) = log 2 .

1.2. Euler Sums and Their Finite Analogs

As is common in the study of number theory, it is often worthwhile to consider the alternating version of an interesting positive series. We now apply this idea to MZVs. For s 1 , , s d N and ε 1 , , ε d = ± 1 , we define the Euler sums
ζ s 1 , , s d ε 1 , , ε d : = n 1 > > n d > 0 j = 1 d ε j n j n j s j .
For convenience, if ε j = 1 , then s ¯ j is used, and, if a substring S repeats n times in the list, then { S } n is used.
To state the relations between Euler sums concisely, we define a kind of double cover of the set N of positive integers.
Definition 1.
Let D be the set of signed numbers N N ¯ , where
N ¯ : = { s ¯ : s N } .
Define the absolute value function | · | on D by | s | = | s ¯ | = s for all s N and the sign function by sgn ( s ) = 1 and sgn ( s ¯ ) = 1 for all s N .
For any s = ( s 1 , , s d ) D d , we define the n-th partial sum of the Euler sums by
ζ n ( s ) : = n > n 1 > > n d > 0 j = 1 d sgn ( s j ) n j n j | s j | .
Similarly to FMZVs, finite Euler sums (FESs) are defined by
ζ A ( s ) : = ζ p ( s ) ( mod p ) p P A .
For example, it is not hard to compute that, for all prime p > 2 , we have (see [16], Theorem 8.2.7)
ζ p ( 1 ¯ ) : = n = 1 p ( 1 ) n n 2 p 1 1 p ( mod p ) .
Thus, ζ A ( 1 ¯ ) = q 2 , as defined in (6), which is the finite analog of ζ ( 1 ¯ ) = log 2 .
In [16], Conjecture 8.6.9, the author of this paper extended the Kakeko–Zagier conjecture to the setting of the Euler sums. For s = ( s 1 , , s d ) D d , define the symmetrized version of the alternating Euler sums by
ζ S ( s ) : = i = 0 d j = 1 i ( 1 ) | s j | sgn ( s j ) ζ ( s i , , s 1 ) ζ ( s i + 1 , , s d ) , ζ S ( s ) : = i = 0 d j = 1 i ( 1 ) | s j | sgn ( s j ) ζ ( s i , , s 1 ) ζ ( s i + 1 , , s d ) ,
where ζ ( = or ) are regularized values (see [16], Proposition 13.3.8). They are called ♯-symmetric Euler sums. If s N d , then they are called ♯-symmetric multiple zeta values (♯-SMZVs or simply SMZVs if ♯ does not matter).
Conjecture 1.
For any w N , let FES w (resp. ES w ) be the Q -vector space generated by all FESs (resp. Euler sums) of weight w. Then, there is an isomorphism
f ES : FES w ES w ζ ( 2 ) ES w 2 , ζ A ( s ) ζ S ( s ) .
Remark 1.
Note that we can replace ζ S ( s ) by ζ S ( s ) (see [16], Exercise 8.7).
To better understand this mysterious relation is the primary motivation of this paper. We mainly study a few variants of the finite analogs of Euler sums by presenting some results that are analogous to those on the Archimedean side.

2. Multiple Mixed Values and Their Finite Analogs

We now turn to the main object of study in this paper. For any s = ( s 1 , , s d ) N d , ε = ( ε 1 , , ε d ) { ± 1 } d , and n N , we define the n-th partial sum of multiple mixed values (MMVs) by
M n ( s ; ε ) : = n > n 1 > > n d > 0 j = 1 d 1 + ε j ( 1 ) n j 2 n j s j = n > n 1 > > n d > 0 n j ( 1 ε j ) / 2 ( mod 2 ) j = 1 d 1 n j s j
and the finite multiple mixed values (FMMVs) by
M A ( s ; ε ) : = M p ( s ; ε ) ( mod p ) p P A .
We call | s | : = s 1 + + s d the weight and d the depth.
The motivation for defining the MMVs is to find a common generalization of a few variants of level-two MZVs, including the following. For all admissible s = ( s 1 , , s d ) N d , we define the multiple t-values (MtVs, see [17]), multiple T-values (MTVs, see [18]), and multiple S-values (MSVs, see [19]) by
t ( s ) : = M ( s ; { 1 } d ) = n 1 > > n d > 0 j = 1 d 1 ( 2 n j 1 ) s j ,
T ( s ) : = M s ; ( 1 ) d , ( 1 ) d 1 , , 1 , 1 = n 1 > > n d > 0 n j d j + 1 ( mod 2 ) j = 1 d 1 n j s j ,
S ( s ) : = M s ; ( 1 ) d 1 , ( 1 ) d 2 , , 1 , 1 = n 1 > > n d > 0 n j d j ( mod 2 ) j = 1 d 1 n j s j ,
respectively. Their finite analogs t A ( s ) , T A ( s ) , and S A ( s ) are defined similarly as in (4) and (9). It is clear that
t A ( s ) = 1 2 d ε 1 , , ε d = ± 1 1 j d ε j ζ A s ε , F A ( s ) = 1 2 d ε 1 , , ε d = ± 1 ( 1 j d 2 | d j if F = T 2 d j if F = S ε j ) ζ A s ε .
More recently, another type of level-two FMZVs has been defined by Kaneko et al. [20] as follows (after multiplying by 2 | s | in the following):
ζ A ( 2 ) ( s ) : = p > n 1 > > n d > 0 , 2 | n j j j = 1 d 1 n j s j p P = 1 2 d ε 1 , , ε d = ± 1 ζ A s ε .
It is also clear that
ζ ( 2 ) ( s ) : = n 1 > > n d > 0 , 2 | n j j j = 1 d 1 n j s j = 1 2 | s | ζ ( s ) .
Definition 2.
Let d N and s = ( s 1 , , s d ) N d . We define the -regularized MTVs ( = or ) and MSVs by
F ( s ) : = 1 2 d σ 1 , , σ d = ± 1 1 j d 2 | d j i f F = T 2 d j i f F = S σ j ζ ( s ; σ ) ( F = T or S ) .
We define the -symmetric multiple T-values (SMTVs) and -symmetric multiple S-values (SMSVs) by
F S ( s ) : = i = 0 d = 1 i ( 1 ) s F ( s i , , s 1 ) F ( s i + 1 , , s d ) , i f d i s e v e n ; i = 0 d = 1 i ( 1 ) s F ˜ ( s i , , s 1 ) F ( s i + 1 , , s d ) , i f d i s o d d ,
where F ˜ = S + T F and we set = 1 0 = 1 , as usual.
In [21], we proved that, for = or ⏙ and for all s N d ,
f ES T A ( s ) = T S ( s ) and f ES S A ( s ) = S S ( s ) ( mod ζ ( 2 ) ) .
In comparison to the results in [19], we define the following subspaces of FES :
  • FES : generated by FESs (finite Euler sums);
  • FMM : generated by FMMVs (finite multiple mixed values);
  • FMt : generated by FMtVs (finite multiple t-values);
  • FMT : generated by FMTVs (finite multiple T-values);
  • FMS : generated by FMSVs (finite multiple S-values);
  • FMZ ( 2 ) : generated by level two FMZVs ζ A ( 2 ) ( s ) defined by (15);
  • FMe : generated by the { M A ( s , ε ) : s N d , ε { ± 1 } d , ε d = 1 } ;
  • FMo : generated by the { M A ( s , ε ) : s N d , ε { ± 1 } d , ε d = 1 } .
Proposition 1.
We have FES w = FMM w for all w N .
Proof. 
This follows easily from the identities
( 1 ) n = ( 1 + ( 1 ) n ) ( 1 ( 1 ) n ) 2 , 1 = ( 1 + ( 1 ) n ) + ( 1 ( 1 ) n ) 2 .
For any fixed weight w 1 , we clearly have the inclusion relations between weight w pieces of the above subspaces:
Foundations 04 00029 i001
Kaneko et al. ([20], Conjecture 5) conjectured that FMZ w ( 2 ) = FES w for all w N . We further conjecture the equal signs in the second row above always hold (also see Conjecture A1).
Conjecture 2.
For sufficiently large weight w N , we have
Foundations 04 00029 i002
Except for the middle vertical equal sign, all the other relations are supported by numerical evidence but no formal proofs yet. In contrast, we have the following conjectural relations between the classical subspaces of Euler sums (cf. the Venn diagram at the end of [19]).
Problem 1.
How can we verify numerically the inclusion FMZ w FMS w FMZ w ( 2 ) and FMZ w FMT w FMt w ? We only need to find a basis in each of the four subspaces on the right and show that every FMZV can be expressed as a Q -linear combination of the basis elements. Is it even possible that both (or one) of the inclusions are actually equalities?
Conjecture 3.
For sufficiently large weight w N , we have
Foundations 04 00029 i003
The left vertical equal sign is obvious by (16) and the middle vertical (strict) inclusion follows from [19], Theorem 7.1, if we assume a variant of Grothendieck’s period conjecture. Again, all the other relations are supported by numerical evidence but no formal proofs yet.
In the following, we consider three classes of relations satisfied by FMMVs: stuffle relations, reversal relations, and linear shuffle relations. They play the key roles in the dimension computation of the vector space generated by FMMVs and its various subspaces.

2.1. Stuffle Relations

Stuffle relations hold in all of the above subalgebras of FES appearing in Conjecture 2, except for FMT and FMS . For convenience, we make D a semi-group by equipping it with a commutative and associative binary operation ⊕ (called O-plus) as follows: for all a , b D ,
a b : = | a | + | b | ¯ , if sgn ( a ) sgn ( b ) ; | a | + | b | , if sgn ( a ) = sgn ( b ) .
For example, for all a , b D ,
ζ A ( a ) ζ A ( b ) = ζ A ( a , b ) + ζ A ( b , a ) + ζ A ( a b ) ,
and if a , b , c D
M A ( a , b ) M A ( c ) = M A ( a , b , c ) + M A ( a , c , b ) + M A ( c , a , b ) + δ sgn ( a ) , sgn ( c ) M A ( a c , b ) + δ sgn ( b ) , sgn ( c ) M A ( a , b c )
where δ s , t is the Kronecker symbol satisfying δ s , t = 1 if s = t and δ s , t = 0 otherwise. Hence, the stuffing for FMMVs occurs only when the two merging components have the same sign.
Problem 2.
Do FMT and FMS form subalgebras of FES ? If not, what is the first instance of a product that is not closed? Note, if the answer is negative, then it cannot be verified rigorously with the current level of knowledge. This is similar to the situation for classical MTV and MSV , and this is for the same type of reason. In the classical setting, counterexamples can only be verified numerically approximately, but they cannot be proven rigorously due to the difficulty in proving transcendence results in general. In the p-adic cases, the transcendence problem for A -numbers may be even harder. For example, we do not even know if β w A vanishes or not for any fixed odd w 3 , although, conjecturally, the density of primes p > w , such that B p w 0 ( mod p ) should be 0.
The question is interesting to us because the corresponding problem for the classical MTVs has an affirmative answer due to their iterated integral expression (see [18], Theorem 2.1).
Here is the process to find such a product of two FMTVs of total weight w numerically, under the assumption that dim Q FES w = F w , which is the same as dim Q FMM w = F w by Proposition 1. First, find a generating set T B w of FMT w , which is possible by using linear shuffle, reversal, and stuffle relations. Second, expand the set into a a generating set M B w of FMM w . If | M B w | = F w , then it is a basis by the assumption dim Q FMM w = F w . Finally, for each product of FMTVs of weight w, we can express it using the conjectural basis M B w . If elements outside of T B w are needed for such a product, then it does not lie in FMT w , meaning that the product of FMT is not closed.

2.2. Reversal Relations

The reversal relations in the next proposition are a class of relations satisfied by the FMMVs but not by the classical MMVs.
Proposition 2.
For all d N , s N d , and ε { ± 1 } d ,
M A ( s , ε ) = ( 1 ) | s | M A ( s , ε ) .
In particular, for all s N d ,
t A ( s ) = ( 1 ) | s | ζ A ( 2 ) ( s ) ,
T A ( s ) = ( 1 ) | s | T A ( s ) a n d S A ( s ) = ( 1 ) | s | S A ( s ) i f 2 | d ,
T A ( s ) = ( 1 ) | s | S A ( s ) a n d S A ( s ) = ( 1 ) | s | T A ( s ) i f 2 d .
Proof. 
The relations follow immediately from the change of indices n p n . □
Proposition 3.
For all d N and s N d ,
T S ( s ) = ( 1 ) | s | T S ( s ) a n d S S ( s ) = ( 1 ) | s | S S ( s ) i f 2 | d ,
T S ( s ) = ( 1 ) | s | S S ( s ) a n d S S ( s ) = ( 1 ) | s | T S ( s ) i f 2 d .
Proof. 
When d is even, we have
T S ( s ) = i = 0 d = i + 1 d ( 1 ) s T ( s i + 1 , , s d ) T ( s i , , s 1 ) = ( 1 ) w i = 0 d = 1 i ( 1 ) s T ( s i , , s 1 ) T ( s i + 1 , , s d ) = ( 1 ) w T S ( s ) .
The same argument works for SMSVs and odd d cases. □
Proposition 4.
For all s N ,
t A ( s , s ) = ζ A ( 2 ) ( s , s ) = q 2 2 / 2 , i f s = 1 ; ( 1 2 1 s ) 2 β s 2 / 2 , i f s 2 ,
and
t A ( s , s , s ) = ζ A ( 2 ) ( s , s , s ) = q 2 3 / 6 + β 3 / 8 , i f s = 1 ; ( 1 2 1 s ) 3 β s 3 / 6 + β 3 s / 8 , i f s 2 .
More generally, for all d N ,
t A ( { s } d ) = ζ A ( 2 ) ( { s } d ) δ s , 1 q 2 s Q + k 1 + + k = d , N β s k 1 β s k Q .
Moreover, we may assume that all k j s are odd.
Proof. 
By the stuffle relations,
2 t A ( s , s ) = t A ( s ) 2 t A ( 2 s ) .
Thus, (25) follows from (37) and (20) by noticing that β w = 0 if w is even. Similarly, by the stuffle relations,
6 t A ( s , s , s ) = t A ( s ) t A ( s , s ) t A ( 2 s , s ) t A ( s , 2 s ) = t A ( s ) ( t A ( s ) 2 s t A ( 2 s ) ) t A ( s ) t A ( 2 s ) + t A ( 3 ) .
Hence, (26) follows from (37) and (20). The general homogeneous FMtVs can be computed similarly by induction or by [7], Theorem 2.3 (see also [22,23], Lemma 5.1). The statement for ζ A ( 2 ) is an easy application of the reversal relation (20) so that the sign in the equation is ( 1 ) s d . But the values could be nonzero only when both s and d are odd, resulting in the negative sign. □

2.3. Linear Shuffle Relations

The most nontrivial relations among finite MZVs and finite Euler sums are provided by the linear shuffle relations, which are closely related to the shuffle relations among the classical MZVs and Euler sums. In this subsection, we extend this to FMMVs and their alternating versions.
For any n N , s = ( s 1 , , s d ) N d , ε = ( ε 1 , , ε d ) { ± 1 } d , and σ = ( σ 1 , , σ d ) { ± 1 } d , we define the partial sums of alternating MMVs by
M n ( s ; ε ; σ ) : = n > k 1 > > k d > 0 ( 1 + ε 1 ( 1 ) k 1 ) σ 1 ( 2 k 1 + 1 ε 1 ) / 4 ( 1 + ε d ( 1 ) k d ) σ d ( 2 k d + 1 ε d ) / 4 k 1 s 1 k d s d .
When n and ( s d , σ d ) ( 1 , 1 ) , we recover the alternating MMVs studied by Xu, Yan, and the present author in [24]. By allowing n to range over the set of primes, we can now define the finite alternating multiple mixed values
M A ( s ; ε ; σ ) : = M p ( s ; ε ; σ ) ( mod p ) p P .
It turns out that, at depth two, all the non-alternating MMVs already have special names, which we now recall. Let d = 2 . Then, we set
ζ A ( 2 ) ( s ; σ ) : = M A ( s ; 1 , 1 ; σ ) , T A ( s ; σ ) : = M A ( s ; 1 , 1 ; σ ) , S A ( s ; σ ) : = M A ( s ; 1 , 1 ; σ ) , t A ( s ; σ ) : = M A ( s ; 1 , 1 ; σ ) .
Moreover, to save space, if an alternating sign σ j = 1 , then we put a bar on top of s j correspondingly. For example,
S A ( 2 ¯ , 3 ) = m > n > 0 , 2 m , 2 | n ( 1 ) ( m + 1 ) / 2 m 2 n 3 .
We recall briefly the main setup for the integral expression of alternating MMVs. Let
a = d t t , b = w + 1 1 : = d t 1 t 2 , c = w 1 1 : = d t 1 + t 2 , β = w + 1 + 1 : = t d t 1 t 2 , γ = w 1 + 1 : = t d t 1 + t 2 .
Set
w σ ε 1 , ε 2 : = max { σ , sgn ( 1 + ε 2 ε 1 ) } w σ ε 1 ε 2 = w σ ε 1 ε 2 , if σ = ε 2 = ε 1 = 1 ; w σ ε 1 ε 2 , otherwise .
It is straightforward to deduce that alternating MMVs can be expressed by the following iterated integrals:
M ( s ; ε ; σ ) = 0 1 w 0 s 1 1 w σ 1 ε 1 , ε 2 w 0 s 2 1 w σ 1 σ 2 ε 2 , ε 3 w 0 s r 1 w σ 1 σ 2 σ r ε r .
For σ , ε { ± 1 } r , define
sgn ( σ , ε ) : = ( 1 ) { i < r σ i = ε i = ε i + 1 ε i + 2 ε r = 1 } .
Then, for all s = ( s 1 , , s r ) N r with ( s 1 , σ 1 ) ( 1 , 1 ) , we have
0 1 w 0 s 1 1 w σ 1 ε 1 w 0 s r 1 w σ r ε r = sgn ( σ , ε ) M ( s ; ε ˜ ; σ ˜ )
where σ ˜ = ( σ 1 , σ 2 σ 1 , , σ r σ r 1 ) and ε ˜ = ( ε 1 ε r , ε 2 ε r , , ε r 1 ε r , ε r ) . See [24] for more details, where the definition of alternating MMVs differs from the one used in this paper by a power of 2.
We now apply the above integral expressions to our finite situations. For example,
0 t b β γ = 0 t i 0 x 2 i d x 0 x j 0 y 2 j + 1 d y 0 y k 0 ( 1 ) k + 1 z 2 k + 1 d z = i , j , k 0 t 2 i + 2 j + 2 k + 5 ( 1 ) k + 1 ( 2 i + 2 j + 2 k + 5 ) ( 2 j + 2 k + 4 ) ( 2 k + 2 ) = m > n > l > 0 , 2 m , 2 | n , 2 | l t m ( 1 ) l / 2 m n l .
Taking the coefficient of t p for any prime p, we obtain
Coeff t p 0 t b β γ = 1 p ζ p ( 2 ) ( 1 , 1 ¯ ) .
Then, from the shuffle relation
0 t b 0 t β γ = 0 t b β γ = 0 t ( b β γ + β b γ + β γ b )
we obtain
p · Coeff t p 0 t b 0 t β γ = ζ p ( 2 ) ( 1 , 1 ¯ ) + S p ( 1 , 1 ¯ ) + t p ( 1 ¯ , 1 ¯ ) .
Hence, we arrive at one of the linear shuffle relations:
ζ A ( 2 ) ( 1 , 1 ¯ ) + S A ( 1 , 1 ¯ ) + t A ( 1 ¯ , 1 ¯ ) = 0 .
Similarly, we can obtain the following linear shuffle relations by (30):
b b b 3 T A ( 1 , 1 ) = 0 ,
b b c 2 T A ( 1 , 1 ¯ ) T A ( 1 ¯ , 1 ¯ ) = 0 ,
γ b γ ( 1 ) p S A ( 1 ¯ , 1 ¯ ) 2 ζ A ( 2 ) ( 1 ¯ , 1 ) = 0 ,
γ γ b 2 t A ( 1 , 1 ¯ ) + S A ( 1 ¯ , 1 ¯ ) = 0 ,
b γ β ζ A ( 2 ) ( 1 ¯ , 1 ¯ ) ( 1 ) p S A ( 1 ¯ , 1 ) ( 1 ) p t A ( 1 ¯ , 1 ) = 0 .
Here and in the rest of this section, we put p = ( p 1 ) / 2 to save space. We observe that the number of b and c must be either one or three in order to have nontrivial relations.

3. Depth-One and -Two Values

In this short section, we compute a few special forms of FMMVs of depth one and two by using the relations we have found in the previous sections.
First, we observe that, since ζ A ( s ) = 0 for all s N , by [16], Theorem 8.2.7,
ζ A ( 2 ) ( s ) = S A ( s ) = t A ( s ) = T A ( s ) = 1 2 ζ A ( s ¯ ) = q 2 , if s = 1 ; ( 2 1 s 1 ) β s , if s 2 ,
where q 2 is the Fermat quotient (6) and β s is given by (5). For weight one and depth one, we have the following result.
Proposition 5.
We have
S A ( 1 ) = ζ A ( 2 ) ( 1 ) = q 2 , ζ A ( 2 ) ( 1 ¯ ) = S A ( 1 ¯ ) = q 2 / 2 ,
t A ( 1 ) = T A ( 1 ) = q 2 , t A ( 1 ¯ ) = T A ( 1 ¯ ) = ( 1 ) p q 2 / 2
where we regard ( 1 ) p as ( 1 ) p p 3 A .
Proof. 
We only need to prove that S A ( 1 ) = 2 S A ( 1 ¯ ) . Indeed,
2 S p ( 1 ) = k = 1 p ( 1 ) k + 1 k + p < k < p , 2 | k 2 k = 2 S p ( 1 ¯ ) + k = 1 p 1 k + 0 < k p , 2 k 1 p k + p < k < p , 2 | k 1 k 2 S p ( 1 ¯ ) + 0 < k p , 2 | k 1 k + p < k < p , 2 | k 1 k ( mod p ) 2 S p ( 1 ¯ ) + S p ( 1 ) ( mod p ) .
This proposition follows easily from (37) and the substitution k p k for t A ( 1 ¯ ) and T A ( 1 ¯ ) . □
For depth-two values, we have the following results.
Proposition 6.
For all a , b N , if w = a + b is odd, then
ζ A ( 2 ) , ( a , b ) = t A ( a , b ) = 1 2 2 1 w 1 ( 1 ) a 2 w w a β w ,
ζ A ( 2 ) ( a , b ) = t A ( a , b ) = 1 2 1 2 1 w ( 1 ) a 2 w w a β w ,
S A ( a , b ) = T A ( a , b ) = ( 1 ) a 2 1 2 w w a β w .
Remark 2.
The formulas for ζ A ( 2 ) ( s ) in (37) and for ζ A ( 2 ) ( a , b ) in (40) are consistent with [20], Proposition 2.1. Note that the ordering in this paper is opposite to that of [20]. Also, our definition of ζ A ( 2 ) ( s ) is 2 | s | times that in [20].
Proof. 
By [16], Theorem 8.6.4,
ζ A ( a , b ) = ζ A ( a , b ) = ( 1 ) a w a β w , ζ A ( a ¯ , b ¯ ) = ζ A ( a ¯ , b ¯ ) = ( 1 ) a ( 2 1 w 1 ) w a β w , ζ A ( a ¯ , b ) = ζ A ( a , b ¯ ) = ( 1 2 1 w ) β w , ζ A ( a ¯ , b ) = ζ A ( a , b ¯ ) = ( 2 1 w 1 ) β w .
The proposition follows easily. □

4. Weight-Two Finite Alternating MMVs

In the previous sections, we have mainly studied FMMVs and some of their special forms. As we mentioned earlier, it is often fruitful to study the alternating version of any interesting positive series. Hence, we turn to the finite alternating MMVs and find the corresponding algebraic structures when the weight is two.
We first recall that the Euler polynomials E n ( x ) are defined by the generating function
2 e t x e t + 1 = n = 0 E n ( x ) t n n ! .
Moreover, E 0 ( 0 ) = 1 and, for all j N ,
E j ( 0 ) = 2 j + 1 j + 1 B j + 1 1 2 B j + 1 = 2 j + 1 ( 1 2 j + 1 ) B j + 1
by [16], p. 242, (8.11). The Euler numbers E n are defined by the generating function
2 e t + e t = n = 0 E n t n n ! ,
which then satisfy
E n = 2 n E n 1 2 .
Let G be the traditional Catalan’s constant defined by
G : = n 1 ( 1 ) n 1 ( 2 n 1 ) 2 .
Then,
T ( 2 ¯ ) = n 1 ( 1 ) n ( 2 n 1 ) 2 = G .
Motivated by the next proposition, we define the finite Catalan’s constant by
G A : = E p 3 2 3 < p P A .
Proposition 7.
We have
T A ( 2 ¯ ) = G A .
Proof. 
For any prime p 5 , we have modulo p
T p ( 2 ¯ ) = 0 < k < p , 2 k ( 1 ) ( k + 1 ) / 2 k 2 = 0 < k < p , 2 | k ( 1 ) ( p k + 1 ) / 2 ( p k ) 2 ( 1 ) ( p + 1 ) / 2 0 < k < p , 2 | k ( 1 ) k / 2 k 2 ( 1 ) ( p + 1 ) / 2 n = 1 p ( 1 ) n 4 n 2 ( 1 ) ( p + 1 ) / 2 4 n = 1 p ( 1 ) n n p 3 ( 1 ) ( p + 1 ) / 2 8 ( 1 ) p E p 3 p + 1 2 + E p 3 0 1 8 E p 3 1 2 1 2 E p 3
by [16], Lemma 8.2.5 and (45) since E p 3 ( 0 ) = 0 for all odd primes p by (44). This completes the proof of the proposition. □
Proposition 8.
We have
S A ( 2 ) = T A ( 2 ) = T A ( 1 , 1 ) = 0 , S A ( 1 , 1 ) = q 2 2 , t A ( 1 , 1 ) = ζ A ( 2 ) ( 1 , 1 ) = q 2 2 2 .
Proof. 
The vanishing of the first three values follows from (37) and (32) quickly. Then, by Proposition 5, we obtain
S A ( 1 , 1 ) = S A ( 1 , 1 ) + T A ( 1 , 1 ) = S A ( 1 ) T A ( 1 ) = q 2 2 .
By (37), we see that
2 t A ( 1 , 1 ) = t A ( 1 ) 2 t A ( 2 ) = q 2 2
and, similarly,
2 ζ A ( 2 ) ( 1 , 1 ) = S A ( 1 ) 2 S A ( 2 ) = q 2 2 .
We have completed the proof of the proposition. □
Proposition 9.
We have
T A ( 1 ¯ , 1 ¯ ) = 2 T A ( 1 , 1 ¯ ) , S A ( 1 ¯ , 1 ¯ ) = 2 t A ( 1 , 1 ¯ ) , t A ( 1 ¯ , 1 ¯ ) = ζ A ( 2 ) ( 1 ¯ , 1 ¯ ) = q 2 2 8 .
Proof. 
The first identity is just (33) and the second is (35). The other two follow immediately from the stuffle relations by Proposition 5. □
Proposition 10.
We have
T A ( 2 ¯ ) = G A , T A ( 1 ¯ , 1 ) = ( 1 ) p 2 G A , S A ( 1 ¯ , 1 ) = ( 1 ) p 2 q 2 2 1 2 G A ,
S A ( 2 ¯ ) = ( 1 ) p G A , T A ( 1 , 1 ¯ ) = 1 2 G A , S A ( 1 , 1 ¯ ) = 1 2 q 2 2 + ( 1 ) p 2 G A ,
t A ( 1 ¯ , 1 ) = 3 ( 1 ) p 8 q 2 2 + 1 2 G A , ζ A ( 2 ) ( 1 ¯ , 1 ) = 1 8 q 2 2 ( 1 ) p 2 G A ,
t A ( 1 , 1 ¯ ) = ( 1 ) p 8 q 2 2 + 1 2 G A , ζ A ( 2 ) ( 1 , 1 ¯ ) = 3 8 q 2 2 ( 1 ) p 2 G A .
Proof. 
Fix a large prime p so that the identities in both Propositions 8 and 9 hold for T p , S p , t p , and ζ p ( 2 ) . Throughout the rest of this proof, we drop the subscript p to save space.
By stuffle relation and Proposition 5, we have
( 1 ) p q 2 2 = S ( 1 ) T ( 1 ¯ ) = S ( 1 ¯ , 1 ) + T ( 1 , 1 ¯ ) ,
( 1 ) p q 2 4 = S ( 1 ¯ ) T ( 1 ¯ ) = S ( 1 ¯ , 1 ¯ ) + T ( 1 ¯ , 1 ¯ ) = 2 T ( 1 , 1 ¯ ) 2 t ( 1 , 1 ¯ ) ,
by Proposition 9. Plugging (54) into (53),
S ( 1 ¯ , 1 ) = 3 T ( 1 , 1 ¯ ) 4 t ( 1 , 1 ¯ ) .
By the change of index k p k for ζ ( 2 ) in (36) (or by (31)), we obtain
t ( 1 ¯ , 1 ) = ( 1 ) p t ( 1 ¯ , 1 ¯ ) S ( 1 ¯ , 1 ) = ( 1 ) p q 2 2 8 S ( 1 ¯ , 1 ) = 3 t ( 1 , 1 ¯ ) 2 T ( 1 , 1 ¯ )
by (54) and (55).
Next, by Proposition 5 and (54),
T ( 2 ¯ ) = t ( 1 ) t ( 1 ¯ ) t ( 1 , 1 ¯ ) t ( 1 ¯ , 1 ) = ( 1 ) p 2 q 2 2 t ( 1 , 1 ¯ ) t ( 1 ¯ , 1 ) = 2 T ( 1 , 1 ¯ )
by (56) and (56). Thus,
T A ( 1 , 1 ¯ ) = 1 2 T A ( 2 ¯ ) = 1 2 G A
by Proposition 7. On the other hand, by the index substitution k p k , we see that
S A ( 2 ¯ ) = ( 1 ) p T ( 2 ¯ ) = ( 1 ) p G A , T A ( 1 ¯ , 1 ) = ( 1 ) p T A ( 1 ¯ , 1 ) = ( 1 ) p 2 G A .
Further, (34) yields that
ζ ( 2 ) ( 1 ¯ , 1 ) = ( 1 ) p 2 S ( 1 ¯ , 1 ¯ ) = q 2 2 8 ( 1 ) p 2 T ( 1 ¯ , 1 ¯ ) = q 2 2 8 ( 1 ) p T ( 1 , 1 ¯ ) = q 2 2 8 + T ( 1 ¯ , 1 )
by (54) and Proposition 9. Hence,
t ( 1 , 1 ¯ ) = ( 1 ) p ζ ( 2 ) ( 1 ¯ , 1 ) = ( 1 ) p 8 q 2 2 + T ( 1 , 1 ¯ ) = ( 1 ) p 8 q 2 2 + 1 2 G A .
All the other identities in the proposition can be now derived by applying the index substitution k p k . □
By combining Propositions 8–10, we immediately obtain the following theorem.
Theorem 1.
Let FAM w be the Q -vector space generated by finite alternating MMVs of weight w. Let ( 1 ) p = ( 1 ) ( p 1 ) / 2 3 p P A . Then,
FAM 2 = q 2 2 , G A 2 , ( 1 ) p q 2 2 , ( 1 ) p G A 2 .
We have carried out some extensive computations of FAM w for small weights w and tabulated the result in Appendix A in this paper.

5. Sum Formulas of FMMVs

One of the most intriguing formulas of multiple zeta values is the so-called sum formula, proven independently by Granville [25] and Zagier around 1997: for all positive integers w > d 1 , let
I w , d : = { s N d : | s | = w } .
Then,
s I w , d ζ ( s ) = ζ ( w ) .
There have been many variations and generalizations of this formula since then. See, for example, refs. [12,26,27,28,29,30,31]. In this section, we prove a few formulas of the same flavor for some special types of FMMVs.

5.1. Sum Formulas of Symmetric and Finite MTVs and MSVs of Even Depth

The first class of formulas we consider concerns only MTVs and MSVs. They reflect clearly the philosophy that the symmetric version should provide the bridge between the p-adic world and the Archimedean world.
Proposition 11.
Suppose w d N with w odd and d even. Then, for F = T and S, we have
s I w , d F A ( s ) = 0 , s I w , d F S ( s ) = 0 .
Proof. 
By reversal relation, we have
s I w , d T A ( s ) = s I w , d ( 1 ) w T A ( s ) = s I w , d T A ( s ) = 0 .
The same argument works for FMSVs. For the symmetric values, by (23), we have
s I w , d T S ( s ) = s I w , d T S ( s ) = ( 1 ) w s I w , d T S ( s ) = s I w , d T S ( s ) = 0 .
The same argument works for SMSVs. This concludes the proof of the proposition. □

5.2. Restricted Sum Relations

In this subsection, we prove the following results, which involve all FMMVs.
Theorem 2.
Let w , d N with d w . Let
I w , d , i : = { s : | s | = w , dep ( s ) = d , s i 2 } 1 i d , I w , d , i , i : = { s : | s | = w , dep ( s ) = d , s i 3 } 1 i d , I w , d , i , j : = { s : | s | = w , dep ( s ) = d , s i 2 , s j 2 } 1 j < i d .
Then,
s I w , d M A ( s ) = s I w , d M A ( s ) = 0 , s I w , d , i M A ( s ) = ( 1 ) i 1 w 1 i 1 + ( 1 ) d w 1 d i β w , s I w , d , i M A ( s ) = ( 1 ) i 1 ( 1 ) d w 1 i 1 + w 1 d i β w , s I w , d , i , j M A ( s ) = ( 1 ) d s I w , d , i , j M A ( s ) = 1 2 N w , d , i , j β w ,
where β w is defined by (5) and N w , d , i , j is an integer explicitly given by [31], Theorem 3.1.
Proof. 
For any prime p and s I w , d , recall that
ζ p ( s ) : = p > n 1 > > n d > 0 j = 1 d 1 n j s j .
By [7], Theorem 4.1, we have
σ S d ζ p ( σ ( s ) ) 0 ( mod p )
where S d is the group of symmetry of d letters. By partitioning I w , d into equivalent classes under permutation, we see quickly that
s I w , d ζ A ( s ) = 0 .
Thus,
s I w , d M A ( s ) = s I w , d 0 < n 1 < < n d < p j = 1 d ( 1 + ( 1 ) n j ) + ( 1 ( 1 ) n j ) 2 n j s j = s I w , d ζ A ( s ) = 0 .
The second and third equations follow from the sum formulas of Saito and Wakabayashi ([12], Theorem 1.4), and the last two from [31], Theorem 3.1 immediately. □

6. Concluding Remarks

In this paper, motivated by the Kaneko–Zagier conjecture that relates MZVs and FMZVs, and the present author’s similar conjecture relating Euler sums and finite Euler sums (which are alternating versions of MZVs and FMZVs, respectively), we have defined and studied the finite versions of multiple mixed values (MMVs), which are further variations of MZVs, by allowing arbitrary parity patterns on the summation indices. As the set of MZVs is included in the set of Euler sums, which, in turn, is a subset of MMVs, we can hope to gain some insight on MZVs and Euler sums by studying MMVs. The same holds true for their finite analogs in the p-adic world. And the bridge between the two worlds is the symmetric version. Philosophically speaking, any Q -linear relation that holds for the symmetric version (modulo ζ ( 2 ) ) should hold for the finite version.
By introducing several crucial algebraic relations, including the stuffle, reversal, and linear shuffle relations, which govern the structure of FMMVs and their various special forms, we were able to carry out explicit computations for generating sets in weight-one and -two cases and, more generally, analyze the dimensions of the various subspaces generated by these values. Furthermore, we proposed a couple of new conjectures (one of which is in Appendix A) based on the numerical evidence, the confirmation of which will greatly deepen our understanding of the algebraic structure and relationships among these mathematical objects.
It is well known that Euler sums have played an important role in theoretical physics. Since all of the MMVs are real numbers, one may wonder if there is any such MMV (and not an Euler sum) that appears essentially in the computation of Feynman integrals. Note that MMVs are level-four objects whose motivic structure has been investigated by Deligne in his influential paper [5], while the motivic nature of some of the Feynman integrals has been revealed in Brown’s paper [32].
However, our research is limited by our current (lack of) knowledge of the still highly mysterious relations between the classical Archimedean world and the p-adic world—in particular, our incomplete understanding of Grothendieck’s theory of motives (especially the crystalline aspect), even though we have a much better understanding of the de Rham and Betti side of the story by the seminal work of Deligne [33], Deligne and Goncharov [34], and Brown [35]. Progress in the study of motives will undoubtedly allow us to see a more holistic and complete picture of the various entities appearing in this paper.

Funding

Jianqiang Zhao was supported by the Jacobs Prize from The Bishop’s School.

Data Availability Statement

All data that are related to this research are provided in this paper.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
(F)MZV(finite) multiple zeta values
(F)ES(finite) Euler sums
(F)MMV(finite) multiple mixed values
(F)MtV(finite) multiple t-values
(F)MTV(finite) multiple T-values
(F)MSV(finite) multiple S-values

Appendix A. Dimensions of Subspaces of Finite MMVs

by Jeremy Feng, Angelina Kim, Sienna Li, Ryan Qin, Logan Wang, and J. Zhao
The following table provides the conjectural dimensions of the various subspaces of FMM s considered in this paper. We obtained these by computing the largest number of independent Q -linear relations among these FMMVs, numerically aided by Maple with the code contained in [16], Appendix D. When a dimension is not followed by a question mark then it means we have verified it numerically by finding the correct number of basis elements.
Table A1. Conjectural dimensions of the various subspaces of FMM s over Q .
Table A1. Conjectural dimensions of the various subspaces of FMM s over Q .
w012345678910111213
FMZ w 10010111223457
FMZ w ( 2 ) 01123581321345589144233
FES w 01123581321345589144233
FMt w 01123581321345589144233
FMT w 0101233691517324476
FMS w 011124571219283966109
FMM w 01123581321345589144?233?
FMe w 01123581321345589?144?233?
FMo w 01123581321345589?144?233?
From numerical computation, we can formulate the following conjecture.
Conjecture A1.
(i) For all w 1 ,
FMZ w ( 2 ) = FMe w = FMM w = FES w = FMo w = FMt w
all have dimension F w .
(ii) For all w 1 ,
FMZ w ( 2 ) ζ A ( 2 ) ( { 1 } w ) Q = FMe w ζ A ( 2 ) ( { 1 } w ) Q = FES w ζ A ( { 1 ¯ } w ) Q = FMo w t A ( { 1 } w ) Q = FMt w t A ( { 1 } w ) Q
all have dimension F w 1 .
(iii) For all k 1 ,
dim Q FMT 2 k + 1 = dim Q FMT 2 k + dim Q FMT 2 k 1 .
Note that Conjecture A1(ii) and [19], Theorem 7.1, can partially explain why the conjectured dimension for FMM w (and FMe w ) computed in this paper differs by 1 from the corresponding dimension for classical MMV w (and MVe w ) that was numerically computed in [19].

References

  1. Zagier, D. Values of zeta functions and their applications. In First European Congress of Mathematics; (Paris, 1992), Vol. II; Joseph, A., Mignot, F., Murat, F., Prum, B., Rentschler, R., Eds.; Birkhäuser: Basel, Switzerland, 1994; pp. 497–512. [Google Scholar]
  2. Hoffman, M.E. Multiple harmonic series. Pac. J. Math. 1992, 152, 275–290. [Google Scholar] [CrossRef]
  3. Broadhurst, D.J. Conjectured enumeration of irreducible multiple zeta values, from knots and Feynman diagrams. arXiv 1996, arXiv:hep-th/9612012. [Google Scholar]
  4. Zhao, J. Alternating Euler sums and special values of Witten multiple zeta function attached to so (5). J. Aust. Math. Soc. 2011, 89, 419–430. [Google Scholar] [CrossRef]
  5. Deligne, P. Le groupe fondamental de la G m μ N , pour N = 2, 3, 4, 6 ou 8. Publ. Math. Inst. Hautes Etudes Sci. 2010, 112, 101–141. (In French) [Google Scholar] [CrossRef]
  6. Broadhurst, D.J. Massive 3-loop Feynman diagrams reducible to SC* primitives of algebras of the sixth root of unity. Eur. Phys. J. C (Fields) 1999, 8, 311–333. [Google Scholar] [CrossRef]
  7. Hoffman, M.E. Quasi-symmetric functions and mod p multiple harmonic sums. Kyushu J. Math. 2015, 69, 345–366. [Google Scholar] [CrossRef]
  8. Zhao, J. Wolstenholme type Theorem for multiple harmonic sums. Intl. J. Number Thyory 2008, 4, 73–106. [Google Scholar] [CrossRef]
  9. Kaneko, M. Finite multiple zeta values. RIMS Kôkyûroku Bessatsu 2017, B68, 175–190. (In Japanese) [Google Scholar]
  10. Jarossay, D. Double mélange des multizêtas finis et multizêtas symétrisés. C. R. Acad. Sci. Paris Ser. I 2014, 352, 767–771. (In French) [Google Scholar] [CrossRef]
  11. Murahara, H. A note on finite real multiple zeta values. Kyushu J. Math. 2016, 70, 197–204. [Google Scholar] [CrossRef]
  12. Saito, S.; Wakabayashi, N. Sum formula for finite multiple zeta values. J. Math. Soc. Jpn. 2015, 67, 1069–1076. [Google Scholar] [CrossRef]
  13. Sakurada, K. Duality for finite/symmetric multiple zeta values of fixed weight, depth, and height. Int. J. Number Theory 2023, 19, 2299–2307. [Google Scholar] [CrossRef]
  14. Singer, J.; Zhao, J. Finite and symmetrized colored multiple zeta values. Finite Fields Their Appl. 2020, 65, 101676. [Google Scholar] [CrossRef]
  15. Yasuda, S. Finite real multiple zeta values generate the whole space Z . Int. J. Number Theory 2015, 12, 787–812. [Google Scholar] [CrossRef]
  16. Zhao, J. Multiple Zeta Functions, Multiple Polylogarithms and Their Special Values; Series on Number Theory and Its Applications: Volume 12; World Scientific Publishing: Singapore, 2016; 620p. [Google Scholar]
  17. Hoffman, M.E. An odd variant of multiple zeta values. Commun. Number Theory Phys. 2019, 13, 529–567. [Google Scholar] [CrossRef]
  18. Kaneko, M.; Tsumura, H. On multiple zeta values of level two. Tsukuba J. Math. 2020, 44, 213–234. [Google Scholar] [CrossRef]
  19. Xu, C.; Zhao, J. Variants of multiple zeta values with even and odd summation indices. Math. Z. 2022, 300, 3109–3142. [Google Scholar] [CrossRef]
  20. Kaneko, M.; Murakami, T.; Yoshihara, A. On finite multiple zeta values of level two. arXiv 2021, arXiv:2109.12501. [Google Scholar] [CrossRef]
  21. Zhao, J. Finite and symmetric Euler sums and finite and symmetric (alternating) multiple T-values. Axioms 2024, 13, 210. [Google Scholar] [CrossRef]
  22. Sun, Z.-W. Congruences involving Bernoulli and Euler numbers. J. Number Theory 2008, 128, 280–312. [Google Scholar] [CrossRef]
  23. Tauraso, R.; Zhao, J. Congruences of alternating multiple harmonic sums. J. Comb. Number Theory 2010, 2, 129–159. [Google Scholar] [CrossRef]
  24. Xu, C.; Yan, L.; Zhao, J. Alternating multiple mixed values, regularization, parity, and dimension conjecture. Indag. Math. 2024; in press. [Google Scholar] [CrossRef]
  25. Granville, A. A decomposition of Riemann’s zeta-function. In Analytic Number Theory; London Mathematical Society Lecture Note Series, Series Number 247; Motohashi, Y., Ed.; Cambridge University Press: Cambridge, UK, 1997; pp. 95–101. [Google Scholar]
  26. Eie, M.; Liaw, W.; Wei, C.S. Double weighted sum formulas of multiple zeta values. Abh. Math. Semin. Univ. Hambg. 2015, 85, 23–41. [Google Scholar] [CrossRef]
  27. Guo, L.; Xie, B. Weighted sum formula for multiple zeta values. J. Number Theory 2009, 129, 2747–2765. [Google Scholar] [CrossRef]
  28. Hirose, H.; Murahara, M.; Saito, S. Weighted sum formula for multiple harmonic sums modulo primes. Proc. Am. Math. Soc. 2019, 147, 3357–3366. [Google Scholar] [CrossRef]
  29. Kawashima, G.; Tanaka, T.; Wakabayashi, N. Cyclic sum formula for multiple L-values. J. Algebra 2011, 348, 336–349. [Google Scholar] [CrossRef]
  30. Murahara, H. A combinatorial proof of the weighted sum formula for finite and symmetric multiple zeta(-star) values. Kobe J. Math. 2021, 38, 73–81. [Google Scholar]
  31. Murahara, H.; Saito, S. Restricted sum formula for finite and symmetric multiple zeta values. Pac. J. Math. 2019, 303, 325–335. [Google Scholar] [CrossRef]
  32. Brown, F. Multiple zeta values and periods: From moduli spaces to Feynman integrals. In Combinatorics and Physics; Contemporary Mathematics; American Mathematical Society: Providence, RI, USA, 2011; Volume 539, pp. 27–52. [Google Scholar]
  33. Deligne, P. Le groupe fondamental de la droite projective moins trois points. In Galois Groups over Q : Proceedings of a Workshop Held March 23–27, 1987; Mathematical Sciences Research Institute Publications; Ihara, Y., Ribet, K., Serre, J.-P., Eds.; Springer: New York, NY, USA, 1989; Volume 16, pp. 79–297. (In French) [Google Scholar]
  34. Deligne, P.; Goncharov, A. Groupes fondamentaux motiviques de Tate mixte. Ann. Sci. de l’Ecole Norm. Supérieure 2005, 38, 1–56. (In French) [Google Scholar] [CrossRef]
  35. Brown, F. Mixed Tate motives over Z . Ann. Math. 2012, 175, 949–976. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, J. Finite Multiple Mixed Values. Foundations 2024, 4, 451-467. https://doi.org/10.3390/foundations4030029

AMA Style

Zhao J. Finite Multiple Mixed Values. Foundations. 2024; 4(3):451-467. https://doi.org/10.3390/foundations4030029

Chicago/Turabian Style

Zhao, Jianqiang. 2024. "Finite Multiple Mixed Values" Foundations 4, no. 3: 451-467. https://doi.org/10.3390/foundations4030029

Article Metrics

Back to TopTop