Skip to main content
Springer logoLink to Springer
. 2021 Feb 24;382(2):815–874. doi: 10.1007/s00220-021-03969-4

Single-Valued Integration and Superstring Amplitudes in Genus Zero

Francis Brown 1,, Clément Dupont 2
PMCID: PMC7940340  PMID: 33758427

Abstract

We study open and closed string amplitudes at tree-level in string perturbation theory using the methods of single-valued integration which were developed in the prequel to this paper (Brown and Dupont in Single-valued integration and double copy, 2020). Using dihedral coordinates on the moduli spaces of curves of genus zero with marked points, we define a canonical regularisation of both open and closed string perturbation amplitudes at tree level, and deduce that they admit a Laurent expansion in Mandelstam variables whose coefficients are multiple zeta values (resp. single-valued multiple zeta values). Furthermore, we prove the existence of a motivic Laurent expansion whose image under the period map is the open string expansion, and whose image under the single-valued period map is the closed string expansion. This proves the recent conjecture of Stieberger that closed string amplitudes are the single-valued projections of (motivic lifts of) open string amplitudes. Finally, applying a variant of the single-valued formalism for cohomology with coefficients yields the KLT formula expressing closed string amplitudes as quadratic expressions in open string amplitudes.

Introduction

The beta function

As motivation for our results, it is instructive to consider the special case of the Euler beta function (Veneziano amplitude [Ven68])

graphic file with name 220_2021_3969_Equ1_HTML.gif 1

The integral converges for Re(s)>0, Re(t)>0. Less familiar is the complex beta function (Virasoro–Shapiro amplitude [Vir69], [Sha70]), given by

βC(s,t)=-12πiP1(C)|z|2s|1-z|2tdzdz¯|z|2|1-z|2=Γ(s)Γ(t)Γ(1-s-t)Γ(s+t)Γ(1-s)Γ(1-t)· 2

The integral converges in the region Re(s)>0, Re(t)>0,Re(s+t)<1.

The beta function admits the following Laurent expansion

β(s,t)=(1s+1t)exp(n2(-1)n-1ζ(n)n((s+t)n-sn-tn)), 3

and the complex beta function has a very similar expansion

βC(s,t)=(1s+1t)exp(n2nodd(-1)n-12ζ(n)n((s+t)n-sn-tn)). 4

It is important to note that these Laurent expansions are taken at the point (s,t)=(0,0) which lies outside the domain of convergence of the respective integrals.

The coefficients in (4) can be expressed as ‘single-valued’ zeta values which satisfy:

ζsv(2n)=0andζsv(2n+1)=2ζ(2n+1)

for n1. The Laurent expansion (4) can thus be viewed as a ‘single-valued’ version of (3). To make this precise, we define a motivic beta function βm(s,t) which is a formal Laurent expansion in motivic zeta values:

βm(s,t)=(1s+1t)exp(n2(-1)n-1ζm(n)n((s+t)n-sn-tn)), 5

whose coefficients ζm(n) are motivic periods of the cohomology of the moduli spaces of curves M¯0,n+3 relative to certain boundary divisors. It has a de Rham version βm,dR(s,t), obtained from it by applying the de Rham projection term by term. One has

β(s,t)=per(βm(s,t))andβC(s,t)=s(βm,dR(s,t))

where s is the single-valued period map which is defined on de Rham motivic periods. We can therefore conclude that the Laurent expansions of β(s,t) and βC(s,t) are deduced from a single object, namely, the motivic beta function (5).

The first objective of this paper is to generalise all of the above for general string perturbation amplitudes at tree-level.

Cohomology with coefficients

There is another sense in which (1) is a single-valued version of (2) that does not involve expanding in st and uses cohomology with coefficients.

For generic values of st (i.e., s,t,s+tZ), it is known how to interpret β(s,t) as a period of a canonical pairing between algebraic de Rham cohomology and locally finite Betti (singular) homology:

H1(X,s,t)andH1lf(X(C),L-s,-t), 6

where X=P1\{0,1,}, s,t is the integrable connection

s,t=d+sdlogx+tdlog(1-x)

on the rank one algebraic vector bundle OX, and L-s,-t is the rank one local system generated by xs(1-x)t, which is a flat section of -s,-t=s,t (see Example 6.13). An important feature of this situation is Poincaré duality which gives rise to de Rham and Betti pairings between (6) for (st) and for (-s,-t). Compatibility between these pairings amounts to the following functional equation for the beta function:

2πi1s+1t=β(s,t)β(-s,-t)2isin(πs)sin(πt)sin(π(s+t)), 7

where the factor in brackets on the left-hand side is the de Rham pairing of dxx(1-x) with itself and the factor in brackets on the right-hand side is the inverse of the Betti pairing of (0,1)xs(1-x)t with (0,1)x-s(1-x)-t.

As in the case of relative cohomology with constant coefficients studied in [BD20], there exists a single-valued formalism for cohomology with coefficients in this setting for which we give an integral formula (Theorem 7.8). This formula implies that βC(s,t) is a single-valued period of (6), which amounts to the equality

βC(s,t)=-(1s+1t)β(s,t)β(-s,-t)-1 8

and proves the second equality in (2).

Applying the functional equation (7) we then get the following ‘double copy formula’ expressing a single-valued period as a quadratic expression in ordinary periods:

βC(s,t)=-12πi(2isin(πs)sin(πt)sin(π(s+t)))β(s,t)2. 9

This formula is an instance of the Kawai–Lewellen–Tye (KLT) relations [KLT86].

In conclusion, there are three different ways to deduce the complex beta function from the classical beta function: via (8) or the double copy formula (9), or by applying the single valued period map term by term in its Laurent expansion.

General string amplitudes at tree level

The general N-point genus zero open string amplitude is formally written as an integral which generalises (1):

Iopen(ω,s_)=0<t1<<tN-3<11i<jN-3(tj-ti)sijω

where ω is a meromorphic form with certain logarithmic singularities (see Sect. 3.2), and s_={sij} are Mandelstam variables satisfying momentum conservation equations (30).

It turns out that one can write the closed string amplitudes in the form

Iclosed(ω,s_)=(2πi)3-NCN-31i<jN-3|zj-zi|2sijνSω¯.

Later we shall rewrite the domain of integration as the complex points of the compactified moduli space of curves of genus 0 with N ordered marked points. Then, the form

νS=(-1)N(N-1)2i=0N-3(ti+1-ti)-1dt1dtN-3(t0=0,tN-2=1)

is logarithmic and has poles along the boundary of the domain of integration of the open string amplitude. It is in fact the image of the homology class of this domain under the map c0 defined in [BD20].

The first task is to interpret the open and closed string amplitudes rigourously as integrals over the moduli space of curves M0,N. An immediate problem is that the poles of the integrand lie along divisors which do not cross normally. Using a cohomological interpretation of the momentum conservation equations in Sect. 3.1, we show how to resolve the singularities of the integral by rewriting it in terms of dihedral coordinates. These are certain cross-ratios uc in the ti, indexed by chords c in an N-gon, whose zero loci form a normal crossing divisor. Thus, for example, we write in Sect. 3.3:

Iopen(ω,s_)=Xδcucscω

where Xδ is the locus where all 0<uc<1 and the sc are linear combinations of the sij. This rewriting of the amplitude evinces the divergences of the integrand and the potential poles in the Mandelstam variables. A similar expression holds for the closed string amplitude, in which ucsc is replaced by |uc|2sc and in which the domain of integration is replaced by the complex points of the Deligne–Mumford compactification M¯0,N.

By an inclusion-exclusion procedure close in spirit to renormalisation1 of algebraic integrals in perturbative quantum field theory [BK13], we can explicitly remove all poles using properties of dihedral coordinates and the combinatorics of chords. The renormalisation fundamentally hinges on special properties of morphisms between moduli spaces which play the role of counter-terms and are described in Sect. 4.

Theorem 1.1

There is a canonical ‘renormalisation’

Iopen(ω,s_)=J1sJXJΩJrenwheresJ=cJsc

indexed by sets J of non-crossing chords in an N-gon, where ΩJren is explicitly defined. The integrals on the right-hand side are convergent around sij=0. They are by definition products of convergent integrals over domains Xδ of various dimensions.

This theorem provides an interpretation of the poles in the Mandelstam variables sij in terms of the poles of ω (see for example (51)). A similar statement holds for the closed string amplitude (Theorem 4.24). Having extended the range of convergence of the integrals using the previous theorem, we are then in a position to take a Laurent expansion around sij=0. The coefficients in this expansion, which are canonical, are products of convergent integrals of the form:

Xδclognc(uc)η

where the product ranges over chords c in a polygon and ncN. We then show how to interpret these integrals as periods of moduli spaces M0,N for larger N by replacing the logarithms with integrals (non-canonically). A key, and subtle point, is that they are integrals over a domain Xδ of a global regular form with logarithmic singularities. We can therefore interpret the previous integrals as motivic periods of universal moduli space motives, and hence define a motivic version of the string amplitude.

Theorem 1.2

There is a motivic string amplitude:

Im(ω,s_)PMT(Z)m,+((sc))

which is a Laurent expansion with coefficients in the ring of motivic multiple zeta values of homogeneous weight. Its period is the open string amplitude

perIm(ω,s_)=Iopen(ω,s_).

It follows that the coefficients of Iopen(ω,s_) are multiple zeta values.

The first statement has been used implicitly in [SS13], [SS19] by assuming the period conjecture for multiple zeta values. The fact that the Laurent coefficients are multiple zeta values is folklore. A subtlety in the previous theorem is that the motivic lift Im(ω,s_) is a priori not unique, as there are many possible ways to express the logarithms log(uc) as integrals. We believe that one could fix these choices if one wished. In any case, the period conjecture suggests that the motivic amplitude Im(ω,s_) is independent of these choices.

By applying the general theorems on single-valued integration proved in the prequel to this paper [BD20] we deduce that the closed string amplitude is the single-valued version of the motivic amplitude.

Theorem 1.3

Let πm,dR denote the de Rham projection map from effective mixed Tate motivic periods to de Rham motivic periods (which maps ζm to ζm,dR), and s the single-valued period map (which maps ζm,dR to ζsv). Then

Iclosed(ω,s_)=sπm,dRIm(ω,s_).

It follows that the coefficients in the canonical Laurent expansion of the closed string amplitudes are single-valued multiple zeta values.

This theorem, for periods (i.e., assuming the period conjecture) was conjectured in [Sti14], [ST14] and proved independently by a very different method from our own in [SS19]. Since the first draft of this paper was written, yet another approach to computing the closed string amplitudes appeared in [VZ18]. An interesting consequence of Theorem 1.3 is that it suggests that the space generated by closed string amplitudes might be closed under the action of the de Rham motivic Galois group. It is important to note that the proof of the previous theorem, in contrast to the approach sketched in [SS19], uses no prior knowledge of multiple zeta values or polylogarithms, and merely involves an application of our general results on single-valued integrals.

String amplitudes from the point of view of cohomology with coefficients and double copy formulae

In the final parts of this paper Sect. 6, 7, we consider the open string amplitude as a period of the canonical pairing between algebraic de Rham cohomology with coefficients in a certain universal (Koba–Nielsen) algebraic vector bundle with connection, and locally finite homology with coefficients in its dual local system:

HN-3(M0,N,s_)HN-3lf(M0,N,L-s_)C.

As in the case of the beta function, Poincaré duality exchanges s_ and -s_ and leads to quadratic functional equations for open string amplitudes generalising (7).

It is important to note that this interpretation of the open string amplitude, as a function of generic Mandelstam variables, is quite different from its interpretation as a Laurent series. After defining the single-valued period map, our main theorem (Theorem 7.8) provides an interpretation of the closed string amplitudes Iclosed(ω,s_) as its single-valued periods. Theorem 7.8 is in no way logically equivalent to the previous results since it is not obvious that the two notions of ‘single-valuedness’, namely as a function of the sij, or term-by-term in their Laurent expansion, coincide. The paper [BD19] provides yet another connection between these two different cohomological points of view.

As a consequence of Theorem 7.8, we immediately deduce an identity relating closed and open string amplitudes which involves the period matrix, its inverse, and the de Rham pairing. By the compatibility between the de Rham and Betti pairings, it in turn implies a ‘double copy formula’ which generalizes (9). It expresses closed string amplitudes as quadratic expressions in open string amplitudes but this time using the Betti intersection pairing (Corollary 7.10). Since Mizera has recently shown [Miz17] that the inverse transpose matrix of Betti intersection numbers coincides with the matrix of KLT coefficients, our formula implies the KLT relations.

Because our results for genus zero string amplitudes are in fact instances of a more general mathematical theory [BD20], valid for all algebraic varieties, we expect that many of these results may carry through in some form to higher genera. It remains to be seen, in the light of [Wit12], if this has a chance of leading to a possible double copy formalism for higher genus string amplitudes.

Contents

In §2 we review the geometry of the moduli spaces M0,N, dihedral coordinates, and the forgetful maps which play a key role in the regularisation of singularities. In §3 we recall the definitions of tree-level string amplitudes, their interpretation as moduli space integrals, and discuss their convergence. Section 4 defines the ‘renormalisation’ of string amplitudes via the subtraction of counter-terms which uses the natural maps between moduli spaces. It uses in an essential way the fact that the zeros of dihedral coordinates are normal crossing. Lastly, in Sect. 5 we construct the motivic amplitude and prove the main theorems using [BD20]. The final sections Sect. 6, 7 treat cohomology with coefficients as discussed above. In an appendix, we prove a folklore result that the Parke–Taylor forms are a basis of cohomology with coefficients.

Dihedral Coordinates and Geometry of M0,S

Let n0 and let S be a set with n+3 elements, which we frequently identify with {1,,n+3}. Let M0,S denote the moduli space of curves of genus zero with marked points labelled by S. It is a smooth scheme over Z whose points correspond to sets of n+3 distinct points psP1, for sS, modulo the action of PGL2. Since this action is simply triply transitive, we can place pn+1=1,pn+2=,pn+3=0 and define the simplicial coordinates (t1,,tn) to be the remaining n points. In other words, they are defined for 1in as the cross-ratios

ti=(pi-pn+3)(pn+1-pn+2)(pi-pn+2)(pn+1-pn+3)· 10

Note that the indexing differs slightly from that in [Bro09]. These coordinates identify M0,S as the hyperplane complement (P1\{0,1,})n minus diagonals, and are widespread in the physics literature. We also use cubical coordinates:

x1=t1/t2,,xn-1=tn-1/tn,xn=tn. 11

Dihedral extensions of moduli spaces

A dihedral structure δ for S is an identification of S with the edges of an (n+3)-gon (which we call (S,δ), or simply S when δ is fixed) modulo dihedral symmetries. When we identify S with {1,,n+3} we take δ to be the ‘standard’ dihedral structure that is compatible with the linear order on S. Let χS,δ denote the set of chords of (S,δ). The dihedral extension M0,Sδ of M0,S is a smooth affine scheme over Z of dimension n defined in [Bro09]. Its affine ring O(M0,Sδ) is the ring over Z generated by ‘dihedral coordinates’ uc, for each chord cχS,δ, modulo the ideal generated by the relations

cAuc+cBuc=1 12

for all sets of chords A,BχS,δ which cross completely (defined in [Bro09, §2.2]). We frequently use the following special case: if c, c are crossing chords, then

uc=1-xuc 13

where x is a product of dihedral coordinates which depends on c,c.

The zero locus of uc is denoted DcM0,Sδ. We have

M0,S=M0,Sδ\DwhereD=cχS,δDc

and D (also denoted by M0,Sδ) is a simple normal crossing divisor. Two components Dc, Dc intersect if and only if c, c do not cross. In the case |S|=4, M0,Sδ=A1, and the divisor D has two components, 0 and 1, corresponding to the two chords in a square. The case |S|=5 is pictured in Fig. 1.

Fig. 1.

Fig. 1

On the left: three out of the five chords in a pentagon, corresponding to the dihedral coordinates u24 (dashed), u35, u13 (dotted). The figure illustrates the relation u24=1-u13u35. On the right: the five divisors on M0,5δ defined by uij=0 form a pentagon. Two divisors intersect if and only if the corresponding chords do not cross

Let us write M¯0,S for the Deligne–Mumford compactification of M0,S. The open subspace M0,SδM¯0,S can be obtained by removing all boundary divisors which are not compatible with the dihedral structure δ. The set of M0,Sδ as δ ranges over all dihedral structures form an open affine cover of M¯0,S.

Morphisms

Given a subset TS with |T|3, let δ|T denote the dihedral structure on T induced by δ. There is a partially defined map fT:χS,δχT,δ|T induced by contracting all edges in (S,δ) not in T. Since some chords map to the outer edges of the polygon (T,δ|T) under this operation, it is only defined on the complementary set of such chords in χS,δ. It gives rise to a ‘forgetful map’

fT:M0,SδM0,Tδ|T

whose associated morphism of affine rings fT:O(M0,Tδ|T)O(M0,Sδ) is

fT(uc)=fT(c)=cuc 14

where cχT,δ|T, and c ranges over its preimages in χS,δ. The forgetful map restricts to a morphism fT:M0,SM0,T between the open moduli spaces.

A dihedral coordinate uc is such a morphism fTc:M0,SδM0,Tcδ|TcA1, where Tc is the set of four edges which meet the endpoints of c.

Strata

Cutting along cχS,δ breaks the polygon S into two smaller polygons, (S,δ) and (S,δ), with S=(S\{c})(S\{c}) (see e.g. [Bro09, Figure 3]). There is a canonical isomorphism

DcM0,Sδ×M0,Sδ. 15

In particular, the restriction of a dihedral coordinate uc to the divisor Dc, where c and c do not cross, is the dihedral coordinate uc on either (S,δ) or (S,δ), depending on which component c lies in.

Definition 2.1

Let JχS,δ be a set of k non-crossing chords. Cutting (S,δ) along J decomposes it into polygons (Si,δi), 0ik. Write

M0,S/Jδ/J=M0,S0δ0××M0,Skδk

and similarly, M0,S/J=M0,S0××M0,Sk. If J={j1,,jk} then set

DJ=Dj1Djk.

There is a canonical isomorphism DJM0,S/Jδ/J.

Trivialisation maps

A crucial ingredient in our ‘renormalisation’ of differential forms is to use dihedral coordinates to define a canonical trivialisation of the normal bundles of the divisors Dc in a compatible manner. In order to define this, we shall fix a cyclic order γ on S, which is compatible with δ. Such a cyclic structure is simply a choice of orientation of the polygon (S,δ).

Definition 2.2

Let c be a chord as above. Let T,T be the subsets of S consisting of the edges in S\{c},S\{c}, respectively, together with the next edge in S with respect to the cyclic ordering γ. Let δ,δ be the induced dihedral structures on T,T. There are natural bijections ST and ST where in each case we identify the chord c with the next edge after S or S in the cyclic ordering. Consider the map

fcγ:M0,SδA1×M0,Sδ×M0,Sδ 16

induced by fcγ=fTc×fT×fT, where MTcδ|Tc is identified with A1. An illustration of this map is given in Fig. 2.

Fig. 2.

Fig. 2

An illustration of the trivialisation map fcγ (the cyclic orientation γ is clockwise, induced by the numbering)

The first component fTc is simply the dihedral coordinate uc, and hence the restriction of (16) to Dc induces the isomorphism (15). Note that (15) is canonical, but fcγ depends on the choice of cyclic structure γ.

Lemma 2.3

If c1,c2χS,δ do not cross, then fc1γfc2γ=fc2γfc1γ.

Proof

Cutting along c1,c2 decomposes the oriented polygon (S,γ) into three smaller polygons (S1,γ1), (S12,γ12), (S2,γ2), where S1 has one edge labelled by c1, S2 has one edge labelled by c2, and S12 two edges labelled c1,c2. The graphs S1S and S2S each have one connected component and S12S has exactly two components (one of which may reduce to a single vertex). Extending each such component by the next edge in the cyclic order defines sets T1,T12,T2S where |T1|=|S1|+1, |T2|=|S2|+1, and |T12|=|S12|+2. One checks from the definitions that

fc1γfc2γ=fTc1×fTc2×fT1×fT12×fT2,

which is symmetric in c1,c2. The point is that the operation of ‘extending by adding the next edge in the cyclic order’ does not depend on the order in which one cuts along the chords c1,c2.

Definition 2.4

Let JχS,δ be a set of non-crossing chords, and define

fJγ:M0,SδAJ×M0,S/Jδ/J

for the composite of the maps fcγ, for cJ, in any order, where AJ=(A1)J. Its restriction to DJ gives the canonical isomorphism DJM0,S/Jδ/J.

When the cyclic ordering is fixed, we shall drop the γ from the notation.

Domains

Let XδM0,Sδ(R) be the subset defined by the positivity of the dihedral coordinates uc>0, for all cχS,δ. In simplicial coordinates (10) it is the open simplex {0<t1<<tn<1}. In cubical coordinates (11) it is the open hypercube (0,1)n. It serves as a domain of integration. On the domain Xδ, every dihedral coordinate uc takes values in (0, 1) by (13). Given a cyclic ordering γ on S, (16) defines a homeomorphism

fcγ:Xδ(0,1)×Xδ×Xδ.

More generally, for any set J of k non-crossing chords (Definition 2.1) set

XJ=Xδ0××Xδk. 17

It follows by iterating the above that

fJγ:Xδ(0,1)k×XJ. 18

The closure Xδ¯M0,Sδ(R) for the analytic topology is a compact manifold with corners which has the structure of an associahedron. Note that the maps fJγ do not extend to homeomorphisms of the closed polytopes Xδ¯.

Logarithmic differential forms

We define ΩS to be the graded Q-subalgebra of regular forms on M0,S generated by the dloguc, cχS,δ. These are functorial with respect to forgetful maps, i.e.

fT:ΩTΩS 19

which follows from (14). One knows that all algebraic relations between the forms dloguc are generated by quadratic relations and furthermore, by Arnol’d–Brieskorn, that M0,S is formal, i.e., the natural map

ΩSHdR(M0,S/Q) 20

is an isomorphism of Q-algebras. Consequently, one has [Bro09, §6.1]

HdR1(M0,S/Q)=cχS,δQ[ducuc]. 21

Finally, it follows from mixed Hodge theory [Del71] (see, e.g., [BD20, §4]), that

ΩSr=Γ(M¯0,S,ΩM¯0,S/Qr(logM¯0,S))

are the global sections of the sheaf of regular r-forms over Q, with logarithmic singularities along M¯0,S=M¯0,S\M0,S.

Residues

Taking the residue of logarithmic differential forms defines a map

ResDc:ΩS|S|-3ΩS|S|-3ΩS|S|-3.

It can be represented graphically by cutting S along the chord c (see e.g. [DV17, Proposition 4.4 and Remark 4.5] where ResDc is denoted by Δ{c} up to a sign). Residues are functorial with respect to forgetful maps:

Lemma 2.5

Let TS as in Sect. 2.2 and cχT,δ|T. Let c be a chord in χS,δ in the preimage of c with respect to fT:χS,δχT,δ|T. Suppose that cutting along c breaks (T,δ|T) into polygons T,T, and cutting along c breaks (S,δ) into polygons S, S. Then the following diagram commutes:

ΩT|T|-3fTΩS|S|-3ResDcResDcΩT|T|-3ΩT|T|-3fTfTΩS|S|-3ΩS|S|-3

Proof

This is simply the functoriality of the residue. It can also be checked explicitly using (14) and (21) which implies that

ResDcfT(dlogucω)=ResDc((dloguc+xdlogux)fT(ω))=fT(ω)|uc=0,

where x ranges over chords in the preimage of c not equal to c. Thus the statement reduces to the equation (fTfT)(ω|uc=0)=fT(ω)|uc=0, which is clear. For a form ω which does not have a pole along Dc, we have ResDcω=0. One checks using (14) and the fact that fT(c)=c that ResDcfT(ω)=0.

In the opposite direction, a cyclic structure γ defines maps

(fcγ):QdxxΩS|S|-3ΩS|S|-3ΩS|S|-3 22

which send dxxωω to dlog(uc)fT(ω)fT(ω).

Lemma 2.6

We have

ResDc(fcγ)(dxxωω)=ωω 23

and

ResDc(fcγ)(dxxωω)=0 24

if c and c cross.

Proof

The first equality follows from the definition. Suppose that cutting (S,δ) along c decomposes it into S,S. Since c crosses c, (13) implies that dlog(uc)=dlog(1-xuc) vanishes along uc=0. Hence forms in (fcγ)(QdxxΩS|S|-3ΩS|S|-3) have no poles along Dc, which proves the second equality.

Summary of structures

With a view to generalisations we briefly list the geometric ingredients in our renormalisation procedure. We have a simple normal crossing divisor DM0,Sδ whose induced stratification defines an operad structure (15). More precisely, this is a dihedral operad in the sense of [DV17]. We have spaces of global regular logarithmic forms equipped with

  • (Residues)
    ResDc:ΩS|S|-3ΩS|S|-3ΩS|S|-3.
  • (Trivialisations, depending on a choice of cyclic structure on S)
    (fcγ):QdxxΩS|S|-3ΩS|S|-3ΩS|S|-3

satisfying a certain number of compatibilities. In this paper, we also use the property uc|Dc=1 if Dc and Dc do not intersect, but we plan to return to the renormalisation of integrals in a more general context with a leaner set of axioms.

Examples

Let |S|=4, and set S={s1,s2,s3,s4} with the natural dihedral structure δ. The square (S,δ) has two chords, and two dihedral coordinates

v24=xandv13=1-x.

Here, and in the next example, a subscript ij denotes the chord meeting the edges labelled {si,si+1} and {sj,sj+1} [Bro09, Figure 2]. The scheme M0,4 is isomorphic, via the coordinate x, to P1\{0,1,} and its dihedral extension is M0,4δ=SpecZ[v24,v13]/(v24+v13=1)A1. The domain XδR\{0,1} is the open interval (0, 1). Its closure Xδ¯A1(R)=R is [0, 1].

Let |S|=5, and set S={s1,s2,s3,s4,s5} with the natural dihedral structure δ. The five chords in the pentagon (S,δ) give rise to five dihedral coordinates which satisfy equations given in [Bro09, §2.2]. These equations define the affine scheme M0,5δ. The pair x=u24,y=u25 are cubical coordinates (11), and embed

(x,y):M0,5M0,4×M0,4P1\{0,1,}×P1\{0,1,}

Its image is the complement of the hyperbola xy=1. We can write all other dihedral coordinates using (12) in terms of these two to give:

u13=1-xy,u24=x,u35=1-x1-xy,u14=1-y1-xy,u25=y.

The domain Xδ maps to the open unit square {(x,y):0<x,y<1}. The first coordinate, x, is the forgetful map which forgets the edge s5:

(x,y)x:M0,5δπM0,4δ

The induced map on affine rings satisfies π(v24)=u24, π(v13)=u13u35.

de Rham projection

We now fix a dihedral structure δ on S and write S for (S,δ). There is a volume form αS,δ on M0,Sδ which is canonical up to a sign [Bro09, §7.1]. A cyclic structure on S defines an orientation on the cell Xδ and fixes the sign of αS,δ if we demand that its integral over Xδ be positive. In simplicial coordinates it is given by [Bro09, (7.1)]:

αS,δ=i=0n-1(ti+2-ti)-1dt1dtn, 25

with the convention t0=0, tn+1=1.

Definition 2.7

Writing uS,δ=cχS,δuc we define

νS,δ=(-1)n(n+1)2uS,δ-1αS,δ 26

Note that the sign is the same as in [BD20, §4.5]. When the dihedral structure is clear from the context, we write νS for νS,δ.

Lemma 2.8

The form νS,δ defines a meromorphic form on M¯0,S with logarithmic singularities, and has simple poles only along those divisors which bound the cell Xδ. In simplicial coordinates (10), and using the convention t0=0, tn+1=1,

νS,δ=(-1)n(n+1)2i=0n(ti+1-ti)-1dt1dtn. 27

Proof

Using the equation 1-uc=cAuc, where A is the set of chords which cross c, which as an instance of (12), we deduce that

uS,δ2=imodn+3(1-u{i,i+1,i+2,i+3}).

Using the definition of dihedral coordinates as cross-ratios [Bro09, (2.6) and §2.1],

uS,δ2=imodn+3(zi-zi+1)(zi+2-zi+3)(zi-zi+2)(zi+1-zi+3)=imodn+3zi-zi+1zi-zi+22.

Using the fact that uS,δ is positive on Xδ one obtains

uS,δ=i=0n(ti+1-ti)i=0n-1(ti+2-ti)-1

after passing to simplicial coordinates. Equation (27) then follows from (25). From (27), we see that νS,δ is, up to a sign, the cellular differential form [BCS10, §2] corresponding to (S,δ). The rest follows from [BCS10, Proposition 2.7].

After passing to cubical coordinates one obtains the more symmetric expression

νS,δ=(-1)n(n+1)2i=1ndxixi(1-xi)· 28

The following proposition follows from the computations in [BD20, §4].

Proposition 2.9

Let [Xδ¯]H0B(M0,Sδ,M0,Sδ) denote the class of the closure of the domain Xδ. Then, with c0 as defined in [BD20, §4], we have

c0[Xδ¯]=(2πi)-nνS,δ.

Working in cubical coordinates and using (28) we get the following compatibility between the νS,δ and the maps fcγ:M0,SM0,Tc×M0,S×M0,S. We set

νc=-ducuc(1-uc)=uc(νTc,δ|Tc).

Lemma 2.10

We have (fcγ)(νcνS,δνS,δ)=(-1)(|S|-1)(|S|-1)νS,δ.

The sign is compatible with the single-valued Fubini theorem discussed in [BD20, §5]: for ωΩTc1, ωΩS|S|-3, ωΩS|S|-3 we have

M¯0,S(C)νS(fcγ)(ωωω)¯=M¯0,4(C)νcω¯M¯0,S(C)νSω¯M¯0,S(C)νSω¯.

Let J={j1,,jk} be a set of k non-crossing chords. With the notation of Definition 2.1 we may define

νJ=νj1νjkΩTj11ΩTjk1νS/J=νS0νSkΩS0|S0|-3ΩSk|Sk|-3.

We have the compatibility

(fJγ)(νJνS/J)=±νS, 29

with a sign that is compatible with the single-valued Fubini theorem.

String Amplitudes in Genus 0

We give a self-contained account of open and closed string amplitudes in genus 0, recast them in terms of dihedral coordinates, and discuss their convergence. The results in this section are standard in the physics literature, which is very extensive. The seminal references are [GSW12], [KLT86]. More recent work, including [SS13], [Sti14], [ST14], [BSST14], served as the main inspiration for the results below.

Momentum conservation

Let N=n+33 and let sijC for 1i,jN satisfying sij=sji and sii=0. Let (xi:yi) denote homogeneous coordinates on P1 for 1iN. Consider the functions

fs_=1i<jN(xjyi-xiyj)sijandgs_=1i<jN|xjyi-xiyj|2sij

on ((C×C)\{0,0})N. The former is multi-valued, the latter is single-valued.

Lemma 3.1

The functions fs_, gs_ define (multi-valued, in the case of fs_) functions on the configuration space of distinct points p1,,pNP1(C) if and only if

1jNsij=0for all1iN. 30

In this case, they are automatically PGL2-invariant and define (multi-valued, in the case of fs_) functions on the moduli space M0,N(C).

Proof

The functions fs_, gs_ are invariant under scalar transformations (xi,yi)(λixi,λiyi) if and only if (30) holds. The first part of the statement follows. For the second, observe that GL2 acts by left matrix multiplication on

x1x2xNy1y2yN

Since each term xjyi-xiyj is minus the determinant of the matrix formed from the columns ij, GL2 acts via scalar multiplication. We have already established that scalar invariance is equivalent to (30), and hence proves the second part. The last part follows since the moduli space M0,N is the quotient of the configuration space of N distinct points in P1 modulo the action of PGL2.

We call (30), together with sij=sji and sii=0 the momentum conservation equations. The solutions to these equations form a vector space (scheme) VN.

When they hold, denote the above functions simply by

fs_=1i<jN(pj-pi)sijandgs_=1i<jN|pj-pi|2sij,

where pi=xi/yi. The former has a canonical branch on the locus where the points pi are located on the circle P1(R) in the natural order, which corresponds to the domain XδM0,N(R). When (30) holds, the differential 1-form

ωs_=dfs_fs_=1i<jNsijdpj-dpipj-pi 31

defines a logarithmic 1-form on (M¯0,N,M¯0,N). We therefore obtain a linear map

VN(K)Γ(M¯0,N,ΩM¯0,N/K1(logM¯0,N))HdR1(M0,N/K) 32

for any field K of characteristic zero.

Lemma 3.2

The map (32) is an isomorphism.

Proof

It is injective: if ωs_ were to vanish then its residue along pi=pj, viewed as a divisor in the configuration space of N distinct points on the projective line, would vanish. Hence all sij=0. Next observe that VN-1VN, and that VN/VN-1 is generated by siN=sNi, for 1iN-1, subject to the single relation

s1N++sN-1N=0.

Therefore dimVN=dimVN-1+N-2. By injectivity, V3=0 since M0,3 is a point. Hence dimVN=N(N-3)/2, which equals dimHdR1(M0,N) by (21), and so (32) is an isomorphism.

String amplitudes in simplicial coordinates

It is customary in the physics literature to write the open and closed string amplitudes in simplicial coordinates (10). We use the coordinate system on VN consisting of the sij for 1i<jn along with the si,n+1 and si,n+3 for 1in. We use the notation s0,i=si,n+3. Let |S|=N=n+3 and let ωΩSn be a global logarithmic form. Let sijC be a solution to the momentum conservation equations (30). The associated open string amplitude is formally written as the integral

0<t1<<tn<10i<jn+1(tj-ti)sijω 33

with the convention t0=0, tn+1=1. In the literature (see Theorem 6.10 below), ω is typically of the form

dt1dtni=0n(tσ(i+1)-tσ(i)) 34

for some permutation σ of {0,,n+1}.

Closed string amplitudes are written in the form

(2πi)-nCn0i<jn+1|zj-zi|2sijνSω¯ 35

where νS was given in Definition 2.7. For ω of the form (34), we can rewrite (35) as

π-nCn0i<jn+1|zj-zi|2siji=0n(zi+1-zi)i=0n(z¯σ(i+1)-z¯σ(i))d2z1d2zn,

with the notation d2z=dRe(z)dIm(z). Note that the apparently complicated sign in the definition of νS is such that all signs cancel in the previous formula, in agreement with the conventions in the physics literature.

Convergence of these integrals is discussed below. As we shall see, a huge amount is gained by first rewriting them in dihedral coordinates.

String amplitudes in dihedral coordinates

Let S=(S,δ) be a set of cardinality N=n+33 and fix a dihedral structure. Suppose that sij are solutions to the momentum-conservation equations. It follows from Lemma 3.2 and (21) that we can uniquely write

ωs_=cχSscducuc,

where the sc are linear combinations of the sij indexed by each chord in (S,δ). Thus the sc form a natural system of coordinates for the space VN. More precisely:

Lemma 3.3

Denoting a chord by a set of edges c={a,a+1,b,b+1}, we have

sij=s{i,i+1,j-1,j}+s{i-1,i,j,j+1}-s{i-1,i,j-1,j}-s{i,i+1,j,j+1}s{i,i+1,j,j+1}=i<a<bjsab. 36

Proof

See [Bro09, (6.14) and (6.17)].

The coordinates sc are better suited than the sij for studying (33) and (35). By (36), we have on appropriate branches (e.g., on Xδ) the equation:

1i<jN(pj-pi)sij=cχSucsc.

Definition 3.4

For a tuple of complex numbers s_=(sc)cχS and a logarithmic form ωΩS|S|-3, define the open string amplitude, when it converges, to be:

Iopen(ω,s_)=XδcχSucscω. 37

Define the closed string amplitude, when it converges, to be:

Iclosed(ω,s_)=(2πi)-nM¯0,S(C)cχS|uc|2scνSω¯. 38

These definitions are equivalent to (33) and (35), respectively, after passing to simplicial coordinates. For the closed string case, one can change its domain using the fact that M¯0,S(C)M0,S(C)Cn differ by sets of Lebesgue measure zero.

In the physics literature, one usually wants to expand string amplitudes in the Mandelstam variables s_. However, the integrals (37) and (38) generally do not converge if s_ is close to zero, as the following propositions show.

Convergence of the open and closed string amplitudes

Proposition 3.5

The integral Iopen(ω,s_) of (37) converges absolutely for scC satisfying

Re(sc)>0ifResDcω0;-1ifResDcω=0.

Proof

Let JχS be a set of non-crossing chords. The set J can be extended to a maximal set JJχS of non-crossing chords. The uj for jJ form a system of local coordinates on M0,Sδ [Bro09, §2.4]. For any ε>0, consider the set

SεJ={0uj<εforjJ,ujεforjJ\J}Xδ¯.

The sets SεJ, for varying J, cover Xδ¯ for sufficiently small ε. This follows because the latter is defined by the domain uc0 for all cχS. Since uc and uc can only vanish simultaneously if c and c do not cross by (13), it follows that

Xδ¯=ε>0JSεJ.

This implies the covering property by compactness of Xδ¯. It suffices to show that the integrand is absolutely convergent on each SεJ. In the local coordinates uj, the normal crossing property means that we can write the integrand of (37) as

cχSucscω=cJucscω0cJucscducucpc

where pc=-ordDcω is the order of the pole of ω along uc=0, and ω0 has no poles on SεJ. Since xαdx is integrable on [0,ε) for Reα>-1, the condition Re(sc-pc)>-1 for all cJ guarantees absolute convergence over SεJ.

Note that the region of convergence does not permit a Taylor expansion at sc=0.

Proposition 3.6

Let N=|S|. The integral Iclosed(ω,s_) of (38) converges absolutely for scC satisfying

1N2>Re(sc)>0ifResDcω0;-12ifResDcω=0. 39

Proof

Let Ω denote the integrand of (38). Let DM¯0,S be an irreducible boundary divisor. Supose first that D is a component of M0,Sδ and is therefore defined by uc=0 for some cχS. By Lemma 2.8, νS has a simple pole along D. In the local coordinate z=uc, Ω has at worst poles of the form:

|z|2scdzdz¯zz¯ifResDcω0,|z|2scdzdz¯zifResDcω=0.

In polar coordinates z=ρeiθ, the left-hand term is proportional to ρ2sc-1dρdθ and hence integrable for Re(sc)>0, the right-hand term to ρ2scdρdθ and hence integrable for Re(sc)>-1/2. Now consider a boundary divisor D which is a component of M¯0,S\M0,S but which is not a component of M0,Sδ (at ‘infinite distance’). It is defined by a local coordinate z=0 (which is a dihedral coordinate with respect to some other dihedral structure on S). By Lemma 2.8, νS has no pole along z=0. Since ω has logarithmic singularities, Ω is locally at worst of the form

|z|2pdzdz¯z

where p is a linear form in the sc. Since any cross-ratio uc has at most a simple zero or pole along D, it follows that p=cχSacsc where ac{0,±1} (an explicit formula for p in terms of sc is given in [Bro09, §7.3]). By passing to polar coordinates one sees that the integrability condition reads 2Re(p)>-1. Assuming (39) one gets the inequality

2Re(p)>-2|χS|N2=-N(N-3)N2>-1,

and we are done.

Put differently, for any sc satisfying the assumptions (39), the integrand of (38) is polar-smooth on (M¯0,S,M¯0,S) in the sense of Definition 3.7 of [BD20].

Example

Let |S|=4, and set ω=dxx(1-x). We have for s,tC,

Iopen(ω,(s,t))=01xs(1-x)tdxx(1-x)=01xs-1(1-x)t-1dx.

This is the classical beta function β(s,t), which converges for Re(s)>0, Re(t)>0. For the closed string amplitude we get

Iclosed(ω,(s,t))=12πiP1(C)|z|2s|1-z|2tdzz(1-z)-dz¯z¯(1-z¯)=1πC|z|2(s-1)|1-z|2(t-1)d2z,

where d2z=dRe(z)dIm(z). This is the complex beta function βC(s,t), which converges for Re(s)>0, Re(t)>0, Re(s+t)<1.

‘Renormalisation’ of String Amplitudes

Formal moduli space integrands

Let us fix S=(S,δ) as above. We shall interpret the integrands of string amplitudes as formal symbols in dihedral coordinates, with a view to either taking a Taylor expansion in the variables sc, or specialising to complex numbers in the case when the integrals are convergent. This will furthermore enable us to treat the open and closed string integrands simultaneously. To this end, consider a fixed commutative monoid (M,+) which is free with finitely many generators. The main example will be MS=cχSNsc, the monoid of non-negative integer linear combinations of the symbols sc.

Definition 4.1

Denote by FS(M) the Q-algebra generated by formal symbols ucm, for cχS and mM, modulo the relations uc0=1 and

ucm+m=ucmucmfor allm,mM.

Similarly, if c is a chord, let Fc(M) be the Q-algebra generated by ucm for mM modulo the above relations. Let us write

AS(M)=FS(M)ΩS|S|-3

and set

Ac(M)=Fc(M)Qdlog(uc).

We write the elements of AS(M) without the tensor product, as linear combinations of cχSucmcω. Similarly, an element of Ac(M) is denoted ucmcdlog(uc).

Definition 4.2

Let JχS be any subset of non-crossing chords as in Definition 2.1. Write J={j1,,jk}. Let us define

AJ(M)=Aj1(M)Ajk(M)AS/J(M)=AS0(M)ASk(M).

Remark 4.3

There is no preferred linear order on J or on the set of polygons that are cut out by J. The tensor products in Definition 4.2 are therefore to be understood in the tensor category of graded vector spaces with the Koszul sign rule, where AS(M) has degree |S|-3, and Ac(M) has degree 4-3=1.

Remark 4.4

We can think of the formal function ucmc as a horizontal section of the formal connection =d-mcdlog(uc) on the trivial rank 1 bundle on the punctured (total space of the) normal bundle to Dc.

A forgetful map fT:M0,SM0,T defines a morphism fT:FT(M)FS(M) via formula (14). By combining it with (19) we get a morphism

fT:AT(M)AS(M). 40

We can realise the formal moduli space integrands as differential forms as follows.

Definition 4.5

Given an additive map α:MC, define a Q-linear map

ραopen:AS(M)differential forms onXδcucmcωcucα(mc)ω.

It is single-valued since ucα=exp(αlog(uc)) and log(uc) has a canonical branch on Xδ, which is the region 0<uc<1. In a similar manner, we can define

ραclosed:AS(M)differential forms onM0,S(C)cucmcω(2πi)-nc|uc|2α(mc)νSω¯.

Infinitesimal behaviour

We define a kind of residue of formal differential forms along boundary divisors which encodes the infinitesimal behaviour of functions in the neighbourhood of the divisor. We first define the evaluation map

evc:FS(M)Fc(M)FS(M)FS(M)

as the morphism sending a formal symbol uc to 1 if c crosses c, and all other symbols to identically named symbols.

Definition 4.6

For any cχS we define the map

Rc:AS(M)Ac(M)AS(M)AS(M)

to be the tensor product of the evaluation map evc and the map of logarithmic differential forms ωdlog(uc)ResDc(ω).

Lemma 4.7

If c1,c2χS do not cross, then Rc1Rc2=Rc2Rc1.

Proof

The commutativity for the evaluation maps is clear. Since dlog(uc1)dlog(uc2)=-dlog(uc2)dlog(uc1), the residues anticommute: ResDc1ResDc2=-ResDc2ResDc1. This sign is compensated by the Koszul sign rule (see Remark 4.3) for the tensor product Ac1(M)Ac2(M)Ac2(M)Ac1(M).

Let J={j1,,jk}χS be a subset of non-crossing chords as in Definition 2.1. By the previous lemma one can compute the iterated residue RJ=Rj1Rjk in any order, which provides a linear map

RJ:AS(M)AJ(M)AS/J(M).

Trivialisation maps

Fix a cyclic ordering γ on S which is compatible with δ. Using the morphisms (40) define for each chord cχS a trivialisation map

fc:Fc(M)FS(M)FS(M)FS(M) 41

by fc(ucmUU)=ucmfT(U)fT(U) (compare (22)). One checks that

evcfc=id. 42

Lemma 4.8

For UFS(M), the difference U-fc(evc(U)) lies in the ideal of FS(M) generated by elements ucm-1 for all chords c crossing c, and mM.

Proof

Follows from the definitions.

By tensoring (41) with the map of logarithmic forms (22) one gets a map

fc:Ac(M)AS(M)AS(M)AS(M). 43

Lemma 4.9

If c1,c2χS do not cross, then fc1fc2=fc2fc1.

Proof

The commutativity for the maps on the components of the tensor products involving formal symbols is clear. The claim is thus a consequence of Lemma 2.3, which treats the components which are logarithmic forms.

Let J={j1,,jk}χS be a subset of non-crossing chords as in Definition 2.1. By the previous lemma one can compute the iterated trivialisation map fJ=fj1fjk in any order, which provides a map

fJ:AJ(M)AS/J(M)AS(M).

Compatibilities

The maps Rc and fc satisfy the following compatibilities.

Lemma 4.10

  1. For every chord c we have Rcfc=id.

  2. For two crossing chords c,c we have Rcfc=0.

Proof

(1) follows from (23) and (42), and (2) follows from (24).

Lemma 4.11

Let c1,c2 be two chords in χS which do not cross. Cutting along the chord ci produces polygons Si,Si, for i=1,2 with the induced cyclic or dihedral structures. Without loss of generality, suppose that c1 lies in S2. Then

Rc1fc2=(idfc2id)(idRc1id) 44

as a map from Ac2AS2AS2Ac1AS1AS1.

Proof

Use the notations of lemma 2.3. We wish to show the following diagram commutes, where the horizontal maps are induced by forgetful morphisms:

Ac2AT1T12AT2ASidRc1idRc1Ac2Ac1AT1AT12AT2Ac1AT1AT12T2

The commutativity of this diagram on the level of formal symbols is clear, and the commutativity on the level of logarithmic forms is a consequence of Lemma 2.5.

Note that (44) has to be understood via the Koszul sign rule.

We can simply write it in the unambiguous form Rc2fc1=fc1Rc2 since the source and target of a map fc or Rc is uniquely determined by the data of c.

The following lemma will not be needed in our renormalisation procedure, but will play a role in the analysis of the convergence of string amplitudes.

Lemma 4.12

For ΩAS(M) and a chord cχS, the difference Ω-fcRcΩ is a linear combination of elements:

  • (i)

    Uω with UFS(M) and ωΩS|S|-3 such that ResDcω=0;

  • (ii)

    (ucm-1)Uω, with UFS(M) and ωΩS|S|-3, for some chord c crossing c and some mM.

Proof

This is a consequence of Lemma 4.8.

Integrability and residues

Proposition 4.13

Let ΩAS(M) such that RcΩ=0 for every chord c. There exists an ε>0 such that:

  1. ραopen(Ω) is absolutely integrable on Xδ¯ for any realisation α:MC such that Reα(m)>-ε for every generator m of M;

  2. ραclosed(Ω) is absolutely integrable on M¯0,S(C) for any realisation α:MC such that -ε<Reα(m)<ε for every generator m of M.

Proof

  1. It suffices to show that the integrand is absolutely convergent on each set SJ, defined in the proof of Proposition 3.5. The normal crossing property implies that we only need to treat divergences in every local coordinate t=uc for c a chord. By Lemma 4.12 it is enough to consider integrands of the form (i) and (ii). Since tαdt is integrable around 0 if Reα>-1, it suffices to check in each case that ραopen(Ω) is a linear combination of tα(m)ω0 for mM and ω0 a smooth form with no poles along t=0. The case (i) is clear. In the case (ii) we use (13) and write 1-uc=tψ0 where ψ0 has no pole along t=0, to deduce that ucα(m)-1=(1-tψ0)α(m)-1. Since forms in ΩS|S|-3 have at most logarithmic poles at t=0, the claim follows.

  2. We need to prove that ραclosed(Ω) is integrable in the neighbourhood of every irreducible component D of M¯0,S=M¯0,S\M0,S. Supose first that D=Dc is a component of M0,Sδ. By Lemma 2.8, νS has a logarithmic pole along Dc. By Lemma 4.12 it is enough to treat the case of integrands (i) and (ii). We work with a local coordinate z=uc. In case (i) we see that the singularities of ραclosed(Ω) are of the type |z|2α(m)dzdz¯z for some mM. Rewriting in polar coordinates z=ρeiθ, this is proportional to ρ2α(m)dρ, which is integrable for Reα(m)>-12. In case (ii) we use (13) to write
    |uc|2α(m)-1=|1-xz|2α(m)-1=zψ0+z¯ξ0
    where ψ0 and ξ0 have no pole along z=0. The singularities of ραclosed(Ω) are thus at worst of the type |z|2α(m)dzdz¯z or |z|2α(m)dzdz¯z¯, and the claim follows as in case (i). Now consider an irreducible component D of M¯0,S which is not a component of M0,Sδ (at ‘infinite distance’). It is defined by a local coordinate z=0. By Lemma 2.8, νS has no pole along z=0 and the singularities of ραclosed(Ω) are at worst of the type |z|2α(m~)dzdz¯z¯ for some m~ in the abelian group generated by M, by the same argument as in the proof of Proposition 3.6. This is integrable around z=0 for Reα(m~)>-12. Since there are finitely many such divisors, the latter condition is implied by the hypotheses (2) for sufficiently small ε.

Remark 4.14

In the case of closed string amplitudes, an integrand ραclosed(Ω) satisfying the assumptions of Proposition 4.13 (2) is polar-smooth on (M¯0,S,M¯0,S) in the sense of Definition 3.7 of [BD20].

Renormalisation of formal moduli space integrands

Definition 4.15

Define a renormalisation map

AS(M)AS(M)ΩΩren=JχS(-1)|J|fJRJΩ, 45

where J ranges over all sets of non-crossing chords in χS.

The reason for calling this map the renormalisation map, even though it does not agree with the notion of renormalisation in the strict physical sense, is that it is mathematically very close to the renormalisation procedure given in [BK13].

Proposition 4.16

For all cχS, RcΩren=0.

Proof

By the second part of Lemma 4.10, RcfJ=RcfcfJ\c=0 if J contains a chord c which crosses c. Let us denote by Sc the set of subsets KχS consisting of non-crossing chords cc that do not cross c. Then, in RcΩren, only the summands indexed by J=K and J=K{c}, for KSc, contribute. Therefore

RcΩren=KSc((-1)|K|RcfKRK+(-1)|K{c}|RcfK{c}RK{c})Ω.

Each summand is of the form

(-1)|K|(RcfKRK-RcfcfKRcRK)Ω.

By the first part of Lemma 4.10, (Rcfc)fKRcRK=fKRcRK. By the commutation relation (44), this is RcfKRK, and therefore the previous expression vanishes.

We extend the renormalisation map to tensor products of forms by defining it be the identity on every Ac(M). For |J|=k it acts upon

AJ(M)AS/J(M)

via idkrenk+1, and is denoted also by ren.

Proposition 4.17

Any form Ω admits a canonical decomposition (depending only on the choice of cyclic ordering of S involved in the definition of fJ):

Ω=JχSfJ(RJΩ)ren 46

where the sum is over non-crossing sets of chords in χS.

Proof

We prove formula (46) by induction on |S|. Suppose it is true for all sets S with <N elements, and let |S|=N. Then applying the formula (46) to each component of RKΩ, for K, we obtain

RKΩ=KJfJ\K(RJ\KRKΩ)ren=KJfJ\K(RJΩ)ren.

Now, substituting into the definition of Ωren, we obtain

Ωren=Ω+K(-1)|K|fK(RKΩ)=Ω+K(-1)|K|KJfKfJ\K(RJΩ)ren=Ω+J(KJ(-1)|K|)fJ(RJΩ)ren

Via the binomial formula,

KJ(-1)|K|=k1(-1)k|J|k=-1

and therefore

Ωren=Ω-JfJ(RJΩ)ren.

Rearranging gives (46) and completes the induction step. The initial case with |S|=3 is trivial, since Ω=Ωren.

Example 4.18

Let |S|=4. Let M=NsNt and consider

Ω=xs(1-x)t(dxx+dx1-x).

Then R0Ω=xsdxx and R1Ω=(1-x)tdx1-x. We have

Ωren=(xs(1-x)t-xs)dxx+(xs(1-x)t-(1-x)t)dx1-x=xs-1((1-x)t-1)dx+(1-x)t-1(xs-1)dx,

and formula (46) is the statement:

Ω=Ωren+xsdxx+(1-x)tdx1-x·

If we identify st and their images by a realisation α:MC then the renormalised open string integrand ραopen(Ωren) is integrable on [0, 1] if Re(s),Re(t)>-1.

On the other hand, the renormalised closed string integrand ραclosed(Ωren) is, up to the factor -(2πi)-1:

(|z|2s|1-z|2t-|z|2s)dzdz¯zz¯(1-z)+(|z|2s|1-z|2t-|1-z|2t)dzdz¯z(1-z)(1-z¯)=|z|2(s-1)(|1-z|2t-1)dzdz¯1-z+|1-z|2(t-1)(|z|2s-1)dzdz¯z·

It is integrable on P1(C) for Re(s),Re(t)>-12 and Re(s+t)<1.

Laurent expansion of open string integrals

Let

Ω=cχSucscω 47

be the integrand of (37), viewed inside AS(MS), where MS=cχSNsc.

Definition 4.19

Let JχS be a set of non-crossing chords. Set

ΩJ=(cχJucsc)ResDJ(ω)AS/J(MS) 48

where ResDJ denotes the iterated residue along irreducible components of DJ and χJ denotes the set of chords in χS\J which do not cross any element of J. Let

sJ=cJsc. 49

The integral of Ω over Xδ can be canonically renormalised as follows.

Theorem 4.20

For all scC satisfying the assumptions of Proposition 4.13 (1),

XδΩ=JχS1sJXJΩJren, 50

where the sum in the right-hand side is over all subsets of non-crossing chords (including the empty set). The integrals on the right-hand side converge for

Re(sc)>-εforsomeε>0.

Proof

For any subset J of non-crossing chords, we have

RJΩ=(cJucscducuc)ΩJhence(RJΩ)ren=(cJucscducuc)ΩJren,

where tensors are omitted for simplicity. By Proposition 4.17, we have

XδΩ=JXδfJ(RJΩ)ren=JfJ(Xδ)(RJΩ)ren.

By (18) we have fJ(Xδ)=(0,1)J×XJ. Each summand in the last term equals

(0,1)J×XJ(cJucscducuc)ΩJren=(cJ1sc)XJΩJren

which proves (50). Absolute convergence of every integral for Re(sc)>-ε is guaranteed by Proposition 4.13 (1) and Proposition 4.16.

The upshot is that each integral on the right-hand side of (50) now admits a Taylor expansion around sc=0 which lies in the region of convergence:

XJΩJrenC[[sc:cχS]].

Note that this integral is a linear combination of a product of integrals over Xδ, for various δ, by (17).

Corollary 4.21

The open string amplitude has a canonical Laurent expansion

Iopen(ω,s_)C[1sc:ordDcω=-1][[sc:cχS]].

By proposition 3.5, it only has simple poles in the sc corresponding to chords c such that ResDcω0. More precisely,

Corollary 4.22

The residue at sc=0 of Iopen(ω,s_) is

ResscXδΩ=XδDcΩc. 51

Proof

It follows from the formula (50) that:

1scResscXδΩ=1scResscJ1sJXJΩJren=cJ1sJXJΩJren.

By similar arguments to those in the proof of theorem 4.20, we have

1scXδDcΩc=fc(Xδ)ucscducucΩc=XδfcRcΩ.

Proposition 4.17 is stated for a form ΩAS(M) but holds more generally for a tensor products of forms in AS(M)AS(M), where S,S are the polygons obtained by cutting S along c. This is because the maps f,R and ren are all compatible with tensor products. Therefore writing RcΩ=ωω (Sweedler’s notation) with ωAS(M), ωAS(M), we deduce that RcΩ equals

JχSfJ(RJω)renJχSfJ(RJω)ren=cJfJ(RJRcΩ)ren

We therefore deduce that

XδfcRcΩ=cJXδfcfJ(RJRcΩ)ren=cJXδfJ(RJΩ)ren=cJ1sJXJΩJren,

where the last equality follows from the same arguments as in the proof of theorem 4.20. We have therefore shown that both sides of (51) coincide for all values of sc such that the integrals converge. Note that since the left-hand side admits a Laurent expansion, the same is true of the right-hand side.

Laurent expansion of closed string amplitudes

The following lemma is the single-valued version of the formula 1s=01xs-1dx.

Lemma 4.23

For all 0<Re(s)<12 the following Lebesgue integral equals

12πiP1(C)|z|2s-dzz(1-z)dz¯z¯=1s.

Proof

Since d|z|2s=s|z|2sdzz+dz¯z¯, the integrand equals

1sd|z|2sdzz(1-z).

For ε>0 small enough, let Uε be the open subset of P1(C) given by the complement of three open discs of radius ε around 0,1, (in the local coordinates z,1-z,z-1). By Stokes’ theorem,

12πiUε|z|2s-dzz(1-z)dz¯z¯=12πi1sUε|z|2sdzz(1-z)

where the boundary Uϵ is a union of three negatively oriented circles around 0,1,. By using Cauchy’s theorem, we see that all integrals in the right-hand side are bounded as ε0, and that the only one which is non-vanishing in the limit as ε0 is around the point 1, giving

limε012πi1sUε|z|2sdzz(1-z)=1s.

Let Ω be as in (47). We set

Ωclosed=(2πi)-ncχS|uc|2scνSω¯.

The closed string amplitudes can be canonically renormalised as follows.

Theorem 4.24

For all scC satisfying the assumptions of proposition 4.13 (2):

M¯0,S(C)Ωclosed=JχS1sJM¯0,S/J(C)(ΩJren)closed, 52

where the sum in the right-hand side is over all subsets of non-crossing chords (including the empty set). The integrals on the right-hand side converge if

-ε<Resc<εfor someε>0.

Proof

As in the proof of Theorem 4.20 we have

M¯0,S(C)Ωclosed=JM¯0,S(C)(fJ(RJΩ)ren)closed

We note that

fJ:M¯0,S(C)(M¯0,4(C))J×M¯0,S/J(C)

is an isomorphism outside a set of Lebesgue measure zero. Using the compatibility (29) between fJ and the forms νS, and making a change of variables via fJ, we can write each summand in the above expression as

cJ12πiM¯0,4(C)|uc|2scνcduc¯uc¯×M¯0,S/J(C)(ΩJren)closed.

By applying Lemma 4.23 we deduce (52).

Corollary 4.25

The closed string amplitude has a canonical Laurent expansion

Iclosed(ω,s_)C[1sc:ordDcω=-1][[sc:cχS]].

By Proposition 3.6, it only has simple poles in the sc corresponding to chords c such that ResDcω0. More precisely,

Corollary 4.26

The residue at sc=0 of Iclosed(ω,s_) is

ResscM¯0,S(C)Ωclosed=M¯0,S(C)×M¯0,S(C)(Ωc)closed, 53

where cutting along c decomposes S into S,S.

The proof is similar to the proof of (51).

Motivic String Perturbation Amplitudes

Having performed a Laurent expansion of string amplitudes, we now turn to their interpretation as periods of mixed Tate motives.

Decomposition of convergent forms

Let VcFS(M) denote the ideal generated by ucm-1 for any mM, where c is a chord which crosses c. More generally, for any set of chords IχS set V=FS(M) and

VI=cIVc.

Let ΩS|S|-3(I)ΩS|S|-3 denote the subspace of forms whose residue vanishes along Dc for all cI. We have the following refinement of Lemma 4.12.

Lemma 5.1

Let χχS be a subset of chords, and let ΩAS(M) such that RcΩ=0 for all chords cχ. Then Ω has a canonical decomposition

Ω=IχΩχ(I) 54

where I ranges over all subsets of χ, and

Ωχ(I)VIΩS|S|-3(χ\I).

Proof

First observe that the case where χ={c} is a single chord follows from Lemma 4.12, since RcΩ=0 implies that Ω=Ω-fcRcΩ, and therefore

ΩVcΩS|S|-3+FS(M)ΩS|S|-3({c}).

Although the sum is not direct, the decomposition into two parts can be made canonical. For this, consider the natural inclusion

ic:Fc(M)FS(M)FS(M)FS(M)

corresponding to the inclusions χS,χSχS. This map satisfies evcic=id. Observe that VcFS(M) is the kernel of the map evc. The map icevc simply sends uc to 1 for all c crossing c, and preserves uc for all other chords. Now write

Ω=(1-icevcid)Ω+(icevcid)Ω

The first term is annihilated by evcid, and so lies in VcΩS|S|-3. The second term satisfies (iddlog(uc)ResDc)(icevcid)Ω=(icid)RcΩ=0 by definition of Rc, and hence lies in FS(M)ΩS|S|-3({c}). This gives a canonical decomposition of the form (54) when χ={c}.

In the general case, proceed by induction on the size of χ by setting:

Ωχc(I)=(icevcid)Ωχ(I)andΩχc(Ic)=Ωχ(I)-Ωχc(I).

Since the maps evc commute for different c, the definition does not depend on the order in which the chords in χc are taken, and the decomposition is canonical. By an identical argument to the one above, we check by induction that indeed

Ωχc(Ic)VIcΩS|S|-3(χ\I)andΩχc(I)VIΩS|S|-3(χc\I)

since (evcid)Ωχc(Ic)=0 and (idResDc)Ωχc(I)=0.

Note that the sum of the spaces VIΩS|S|-3(χ\I) is not direct.

Logarithmic expansions

For each chord c, let c be a formal symbol which we think of as corresponding to a logarithm of uc.

There is a continuous homomorphism of algebras defined on generators by

FS(M)Q[{c}cχS][[M]]ucmn0mnn!cn. 55

This extends to a map

AS(M)ΩS|S|-3[{c}cχS][[M]]. 56

For an additive map α:MC we have the realisation maps ραopen and ραclosed from Definition 4.5. A form ραopen(Ω) (resp. ραclosed(Ω)) has a series expansion given by composing (56) with α and by interpreting the formal symbols c as:

clog(uc)(resp.clog|uc|2).

Definition 5.2

A convergent monomial is one of the form

(cχSckc)ωΩS|S|-3[{c}cχS] 57

where for every cχS such that ResDcω0, there exists another chord cχS which crosses c such that kc1.

Lemma 5.3

For a convergent monomial (57), the corresponding integrals

Xδ(cχSlogkc(uc))ωandM¯0,S(C)(cχSlogkc|uc|2)νSω¯

are convergent.

Proof

For any chord c which crosses cχS, equation (13) implies that

log(uc)=αucandlog|uc|2=βuc+γuc¯

for some α,β,γ. By applying this to every chord c for which ResDcω0, we can rewrite the above integrals as linear combinations of

XδF(uc)ωrespectivelyM¯0,S(C)G(uc)νScχucuc¯ω¯

where FG have at most logarithmic singularties near boundary divisors, χχS is a subset of chords, and ω has no poles along Dc, for all cχS. Convergence in both cases follows from a very small modification of propositions 3.5, 3.6 to allow for possible logarithmic divergences. The latter do not affect the convergence since |log(z)|kzs tends to zero as z0 for any Res>0.

Proposition 5.4

Let ΩAS(M) such that RcΩ=0 for all chords c. Then Ω admits a canonical expansion which only involves convergent monomials (57).

Proof

Apply (55) to each term in Ωχ(I) in the decomposition (54).

Corollary 5.5

Renormalised amplitudes, where they converge, can be canonically written as infinite sums of integrals of convergent monomials in logarithms:

XδραopenΩren=K=(kc)aKXδ(cχSlogkc(uc))ωKM¯0,S(C)ραclosedΩren=(2πi)-nK=(kc)aKM¯0,S(C)(cχSlogkc(|uc|2))νSωK¯, 58

where aK,aK lie in the Q-subalgebra of C generated by α(M). Each integral on the right-hand side converges.

Equivalently, if we treat the elements of M as formal variables then the open and closed string amplitudes admit expansions in C[[M]] whose coefficients are canonically expressible as Q-linear combinations of integrals of convergent monomials as above.

Example 5.6

We apply the above recipe to Example 4.18. By abuse of notation we identify s, t with their images under a realisation α:NsNtC, and write (Ωren)open=ραopen(Ωren), (Ωren)closed=ραclosed(Ωren). In the open case we get

Xδ(Ωren)open=m0,n1smm!tnn!01logm(x)logn(1-x)dxx+m1,n0smm!tnn!01logm(x)logn(1-x)dx1-x· 59

Note that log(x) vanishes at x=1, and log(1-x) at x=0, so the integrals on the right-hand side are convergent. In the closed case we get

P1(C)(Ωren)closed=1πm0,n1smm!tnn!P1(C)logm|z|2logn|1-z|2d2z|z|2(1-z)+1πm1,n0smm!tnn!P1(C)logm|z|2logn|1-z|2d2zz|1-z|2· 60

Again, the integrals on the right-hand side are convergent.

Removing a logarithm

We can now replace each logarithm with an integral one by one. It suffices to do this once and for all for |S|=5.

Example 5.7

(The logarithm) Consider the forgetful map x:M0,5M0,4 of example 2.9. It is a fibration whose fibers are isomorphic to the projective line minus 4 points. More precisely, the forgetful maps u24=x and u25=y embed M0,5 as the complement of 1-xy=0 in the product M0,4×M0,4. The projection onto the first factor gives a commutative diagram

M0,5M0,4×M0,4xM0,4=M0,4

The projection restricts to the real domains X5X4 whose fibers are identified with (0, 1) with respect to the coordinate y. Then

0<y<1du13u13=0<y<1dlog(1-xy)=log(1-x).

The function 1-x is the dihedral coordinate v13 on M0,4 and so

logv13=0<y<1du13u13. 61

In this manner we shall inductively replace all logarithms of dihedral coordinates with algebraic integrals. Note that it is not possible to express the logarithmic dihedral coordinate logv24=logx as an integral of another logarithmic dihedral coordinate over the fiber in y with respect to the same dihedral structure. It is precisely this subtlety that complicates the following arguments.

From now on, we fix a dihedral structure (S,δ), and consider a differential form of degree |S|-3 of the following type:

Ω=(cχS,δlognc(uc))ω0 62

where ω0Ω|S|-3(M0,S). Suppose that it is convergent, i.e., for every chord c such that ResDcω00, there exists a cI which crosses c with nc1. Define

weight(Ω)=|S|-3+cχS,δnc.

We can remove one logarithm at a time as follows.

Lemma 5.8

Pick any chord c such that nc1, and write

Ω=log(uc)Ω.

Then there exists an enlargement (Sc,δc) of (S,δ), i.e., SSc and |Sc\S|=1, where the restriction of δc to S induces δ, and a differential form Ω of degree |Sc|-3 which is a sum of convergent forms (62) such that

Xδlog(uc)Ω=XδcΩ. 63

Furthermore, each monomial in Ω has weight equal to that of Ω.

Proof

The chord c on S is determined by four edges T={t1,,t4}S, where t1,t2 and t3,t4 are consecutive with respect to δ. This identifies M0,TM0,4 and uc with the dihedral coordinate v13 in M0,4. Consider the set Sc=S{t5}, equipped with the dihedral structure δc obtained by inserting a new edge t5 next to t1 and in between t1 and t4. Let T be the set of five edges T{t5} with the dihedral structure inherited from δc. Then M0,TM0,5 (see example 5.7).

Consider the diagram

M0,ScδcfTM0,5δc|TfSxM0,SδucM0,4δ.

It commutes since forgetful maps are functorial. Let βΩ1(M0,5) denote the form dlogu13 of example 5.7 whose integral in the fiber yields log(v13) and set

Ω=fS(Ω)fT(β).

The product f=fS×fT induces a morphism

f:M0,ScM0,S×M0,4M0,5

and an isomorphism f:XδcXδ×(0,1). Since Ω=f(Ωβ), we find by changing variables along the map f that

XδcΩ=Xδ×(0,1)Ωβ=Xδlog(uc)Ω.

The second integral takes place on the fiber product M0,S×M0,4M0,5 and is computed using (61). This proves equation (63).

We now check that Ω is of the required shape (62) with respect to M0,Scδc and convergent. First of all, observe that for any forgetful morphism f:SS and any chords ab in S which cross, we have by (14)

flog(ua)dubub=a,blog(ua)dubub

where a,b range over chords in S in the preimage of a and b respectively. Every pair a,b crosses. It remains to check the convergence condition along the poles of the 1-form β. For this, denote the two chords in Sc lying above the chord c by c1,c2. The chord c1 corresponds to edges {t5,t1;t3,t4} and c2 to {t1,t2;t3,t4}. By (14), we have fS(uc)=uc1uc2, and by example 5.7fTβ=duc2uc2, (β corresponds to dlogu13 in example 2.9). We therefore check that

fS(duc)fT(β)=d(uc1uc2)duc2uc2=duc1duc2fSlog(uc)ducucfT(β)=clogucduc1uc1duc2uc2

where c crosses c. In the sum, c ranges over the preimages of c under fS, and necessarily crosses both c1 and c2. It follows that Ω is a sum of convergent monomials in logarithms. The statement about the weights is clear.

Remark 5.9

Note that Sc depends on the choice of where to insert the new edge we called t5. Similarly, the computation in example 5.7 also involves a choice: we could instead have used

-log(1-x)=0<y<1du35u35=0<y<1dlog(1-x1-xy).

Thus there are two different ways in which we can remove each logarithm. One can presumably make these choices in a canonical way.

Corollary 5.10

Let Ω be of the form (62) and convergent. Then the integral I of Ω over Xδ is an absolutely convergent integral

Xδω 64

where SS is a set with dihedral structure δ compatible with δ, and ωΩS a logarithmic algebraic differential form with no poles along the boundary of M0,Sδ. Furthermore, |S|=weight(Ω)+3.

Proof

Apply the previous lemma inductively to remove the logarithms log(uc) one at a time. At each stage the total degree of the logarithms decreases by one. One obtains a Z-linear combination of convergent integrals of the form (64). Add the integrands together to obtain a single integral of the required form.

Motivic versions of open string amplitude coefficients

Let MT(Z) denote the tannakian category of mixed Tate motives over Z with rational coefficients [DG05]. An object HMT(Z) has two underlying Q-vector spaces HdR (the de Rham realisation) and HB (the Betti realisation) together with a comparison isomorphism comp:HdRQCHBQC. For every integer N3 we have an object H=HN-3(M0,Nδ,M0,Nδ) in MT(Z) whose de Rham and Betti realisations are the usual relative de Rham and Betti cohomology groups of the pair (M0,Nδ,M0,Nδ) [GM04], [Bro09].

Let us recall [Bro14b], [Bro17] the algebra PMT(Z)m of motivic periods of the category MT(Z). Its elements can be represented as equivalence classes of triples [H,[σ],[ω]]m with HMT(Z), [σ]HB and [ω]HdR. It is equipped with a period map

per:PMT(Z)mC

defined by per[H,[σ],[ω]]m=[σ],comp[ω]. Let us also recall the subalgebra PMT(Z)m,+ of effective motivic periods of MT(Z).

Corollary 5.11

Let Ω be of the form (62) and convergent. Then the integral

I=XδΩ

is a period of a universal moduli space motive HN-3(M0,Sδ,M0,Sδ), where SS is a set with dihedral structure δ compatible with δ, and |S|=N=3+weight(Ω). More precisely, we can write I=perIm where

Im=[HN-3(M0,Sδ,M0,Sδ),Xδ,[ω]]mPMT(Z)m,+

is an effective motivic period of weight N-3 and ωΓ(M¯0,N,ΩM¯0,SN-3(logM¯0,S)) is a logarithmic differential form.

Proof

Let (S,δ) and ω be as in Corollary 5.10, set H=HN-3(M0,Sδ,M0,Sδ) and define Im as in the statement. We have

I=Xδω=perIm

by definition of the comparison isomorphism for H. The statement about the weight follows from the fact that ω is logarithmic in the following way. As for every mixed Tate motive, the weight filtration on HdR is canonically split by the Hodge filtration [DG05, 2.9] and we have a weight grading on H; this splitting implies in particular that H=W2(N-4)HFN-3H. The statement about the weight says that the class of ω is a homogeneous element of weight 2(N-3) with respect to this grading, i.e., that [ω]FN-3H. This can be checked after extending the scalars to C and thus follows from Corollary 4.13 in [BD20] (compare with [Dup18, Proposition 3.12]).

Remark 5.12

Note that the motivic lift Im of I depends on some choices which go into Lemma 5.8. One expects, from the period conjecture, that it is independent of these choices. One can possibly make the lift canonical by fixing choices in the application of Lemma 5.8.

We deduce a number of consequences:

Theorem 5.13

The coefficients in the Laurent expansion of open string amplitudes with N particles are multiple zeta values. More precisely,

Iopen(ω,s_)=n_=(nc)cχSζn_s_n_wheres_n_=cχSscnc

and each nc-1. Here, ζn_ is a Q-linear combination of multiple zeta values of weight N+|n_|-3, where |n_|=cχSnc.

Proof

Use the fact that the periods of universal moduli space motives are multiple zeta values [Bro09]. One can obtain the statement about the weights either by modifying the argument of loc. cit or as a corollary of the next theorem (using the fact that a real motivic period of weight n of an effective mixed Tate motive over Z is a Q-linear combination of motivic multiple zeta values of weight n).

Theorem 5.13 is well-known in this field using results scattered throughout the literature, but until now lacked a completely rigorous proof from start to finish.

Theorem 5.14

The above expansion admits a (non-canonical) motivic lift

Im(ω,s_)=n_=(nc)cχSζn_ms_n_ 65

where ζn_m is a Q-linear combination of motivic multiple zeta values of weight N+|n_|-3, whose period is ζn_.

Proof

Apply the Laurent expansions (58) to each renormalised integrand (50) and invoke Corollary 5.11. This expresses the terms in the Laurent expansion as linear combinations of products of motivic periods of the required type.

Remark 5.15

The existence of a motivic lift is a pre-requisite for the computations of Schlotterer and Stieberger [SS13], in which the motivic periods are decomposed into an ‘f-alphabet’ (rephrased in a different language, that paper and related literature studies the action of the motivic Galois group on Im(ω,s_)). In [SS19], this is achieved by assuming the period conjecture. The computation of universal moduli space periods in terms of multiple zeta values can be carried out algorithmically [Bro09], [Pan15], [Bog16]. This type of analytic argument (or [Ter02], [BSST14]) is used in the literature to deduce a theorem of the form 5.13, but it is not capable of proving the much stronger statement 5.14.

Example 5.16

(‘Motivic’ beta function). We now treat the case of the beta function (Example 4.18) by using the expansion (59) of the renormalised part. We can remove all logarithms at once and write, for ω{dxx,dx1-x}:

01logm(x)m!logn(1-x)n!ω=(-1)m+nΔm+n+1du11-u1dun1-unωdv1v1dvmvm

where Δm+n+1={0<u1<<un<x<v1<<vm<1} is the standard simplex. We thus get the following expansion:

β(s,t)=1s+1t+m0,n1(-s)m(-t)nζ({1}n-1,m+2)+m1,n0(-s)m(-t)nζ({1}n,m+1),

which can be rewritten as

β(s,t)=1s+1t1-m1,n1(-s)m(-t)nζ({1}n-1,m+1). 66

The above argument yields a ‘motivic’ beta function

βm(s,t)=1s+1t1-m1,n1(-s)m(-t)nζm({1}n-1,m+1). 67

whose period, applied termwise, gives back (66).

Note that Ohno and Zagier observed in [OZ01] that (66) agrees with the more classical expansion of the beta function (3). Likewise, one can verify using motivic-Galois theoretic techniques that (67) indeed coincides with the definition (5).

Single valued projection and single-valued periods

We let PMT(Z)m,dR denote the algebra of motivic de Rham periods of the category MT(Z) [Bro17] (see also [BD20, §2.3]). It is equipped with a single-valued period map

s:PMT(Z)m,dRR

defined in [Bro14a], [Bro17] (see also [BD20, §2]). The de Rham projection

πm,dR:PMT(Z)m,+PMT(Z)m,dR

on effective mixed Tate motivic periods was defined in [Bro14a] and [Bro17, 4.3] (see also [BD20, Definition 4.3]).

Definition 5.17

Given a choice of motivic lift (65), define its de Rham projection to be its image after applying πm,dR term-by-term:

Im,dR(ω,s_)=n_=(nc)cχSζn_m,dRs_n_,whereζn_m,dR=πm,dRζn_m.

This makes sense since ζn_m is effective. Likewise, define its single-valued version

Isv(ω,s_)=n_=(nc)cχSζn_svs_n_,whereζsv=sζn_m,dR

It is a Laurent series whose coefficients are Q-linear combinations of single-valued multiple zeta values.

Since svπm,dR=sπm,dR, we could equivalently have applied the map sv, which is specific to the mixed Tate situation (see [BD20, §2.6]). We now compute Isv(ω,s_).

Lemma 5.18

For any xC\{1},

log|1-x|2=12πiP1(C)-dyy(1-y)dlog(1-x¯y¯)

Proof

This follows from the computations of [BD20, §6.3] after a change of coordinates.

Theorem 5.19

Consider an integral of the form

I=XδcχS(lognc(uc))ω0

where the integrand is convergent of the form (62). Let Im denote a choice of motivic lift (Corollary 5.11). Its single-valued period Isv=sπm,dR(Im) is

Isv=(2πi)3-|S|M¯0,S(C)cχS(lognc|uc|2)νSω0¯, 68

and in particular does not depend on the choice of motivic lift.

Proof

Repeated application of Lemma 5.8 (which may involve a choice at each stage), gives rise to a dihedral structure (S,δ), a morphism

f:M0,SM0,S×Ak(M0,5)kM0,S×(P1\{0,1,})k,

and a differential form ωΩS with no poles along M0,Sδ such that

I=Xδωis the period ofIm=[H|S|-3(M0,Sδ,M0,Sδ),Xδ,ω]m.

The form ω satisfies ω=f(ω0β) where

β=dlog(1-x1y1)dlog(1-xkyk)

and x1,,xk denote the coordinates on Ak and correspond to the 1-uc, with multiplicity nc, taken in some order. By [BD20, Theorem 3.16] and Corollary 2.9,

sπm,dRIm=(2πi)3-|S|M¯0,S(C)νSω¯. 69

Since ω,νS are logarithmic with singularities along distinct divisors, the integral converges. By repeated application of Lemma 2.10, we obtain that νS is, up to a sign, the pullback by f of the form

νS-dy1y1(1-y1)-dykyk(1-yk),

the sign being such that after changing coordinates via f we obtain:

Isv=(2πi)3-|S|M¯0,S(C)νSω0¯×j=1kP1(C)-dyjyj(1-yj)dlog(1-xj¯yj¯).

Formula (68) follows on applying Lemma 5.18.

Theorem 5.20

We have

Isv(ω,s_)=Iclosed(ω,s_). 70

In other words, the coefficients in the canonical Laurent expansion of the closed string amplitudes (52) are the images of the single-valued projection of the coefficients in any motivic lift of the expansion coefficients of open string amplitudes.

Proof

By (50), we write the open string amplitude as

XδΩ=JχS1sJXJΩJren, 71

By Corollary 5.5, each integrand on the right-hand side admits a Taylor expansion, whose coefficients are products of integrals over moduli spaces, each of which can be lifted to motivic periods by Corollary 5.11. Thus

XJΩJren=per(IJm(s_))

for some formal power series IJm(s_) in the sc, cχS whose coefficients are effective motivic periods coming from tensor products of universal moduli space motives. Since s is an algebra homomorphism, (68) yields

sπm,dRIJm=M¯0,S/J(C)(ΩJren)closed.

On the other hand, by (52) these integrals can be repackaged into

JχS1sJM¯0,S/J(C)(ΩJren)closed=M¯0,S(C)Ωclosed.

By abuse of notation, we may express the previous theorem as the formula

sXδcχSucscω=(2πi)3-|S|M¯0,S(C)cχS|uc|2scνSω¯

which is equivalent to the form conjectured in [Sti14].

Remark 5.21

Define a motivic version of closed string amplitudes by setting

Isv,m(ω,s_)=smIm,dR(ω,s_)

where sm was defined in [Bro17, (4.3)] (see also [BD20, Remark 2.10]). Its period is per(Isv,m(ω,s_))=Isv(ω,s_), which coincides with the closed string amplitude, by the previous theorem. This object is of interest because it immediately implies a compatibility between the actions of the motivic Galois group on the open and closed string amplitudes.

Background on (Co)homology of M0,S with Coefficients

Most, if not all, of the results reviewed below are taken from the literature. A proof that the Parke–Taylor forms are a basis for cohomology with coefficients can be found in the appendix. See [KY94a], [KY94b], [CM95], [Mat98], [MY03] and references therein for more details.

Koba–Nielsen connection and local system

Let S be a finite set with |S|=N=n+33. Let s_=(sij) be a solution to the momentum conservation equations (30). Let

Qs_B=Q(e2πiskl)andQs_dR=Q(skl) 72

be the subfields of C generated by the exp(2πiskl) and skl respectively.

Definition 6.1

Let OS denote the structure sheaf on M0,S×QQs_dR. The Koba–Nielsen connection [KN69] is the logarithmic connection on OS defined by

s_:OSΩS1wheres_=d+ωs_

and ωs_ was defined in (31). The Koba–Nielsen local system is the QsB-local system of rank one on M0,S(C) defined by

Ls_=Qs_B1i<jN(pj-pi)-sij.

Since ωs_ is a closed one-form, the connection s_ is integrable. The horizontal sections of the analytification (OSan,s_an) define a rank one local system over the complex numbers that is naturally isomorphic to the complexification of Ls_:

OSans_anLs_Qs_BC. 73

We will also consider the dual of the Koba–Nielsen local system

Ls_=Qs_B1i<jN(pj-pi)sijL-s_. 74

Let (S,δ) be a dihedral structure. In dihedral coordinates,

s_=d+cχS,δscducucandLs_=Qs_BcχS,δuc-sc.

Definition 6.2

A solution to the momentum conservation equations (30) is generic if

i,jIsijZ 75

for every subset IS with |I|2, and |S\I|2.

Remark 6.3

Write H=HdR1(M0,S/Q). By formality (20) and (21), and Lemma 3.2, the form ωs_ is the specialisation of the universal abelian one-form

ωHΩS1HH

which represents the identity in HHEnd(H).

Remark 6.4

The formal one-form ω defines a logarithmic connection on the universal enveloping algebra of the braid Lie algebra. It is the universal connection on the affine ring of the unipotent de Rham fundamental group π1dR(M0,S):

KZ:O(π1dR(M0,S))ΩS1O(π1dR(M0,S))

The Koba–Nielsen connection (viewed as a connection over the field Qs_dR, i.e., for the universal solution of the momentum-conservation equations) is its abelianisation. Given any particular complex solution to the moment conservation equations, the latter specialises to a connection over C.

Singular (co)homology

Denote the (singular) homology, locally finite (Borel-Moore) homology, cohomology, and cohomology with compact supports of M0,S with coefficients in Ls_ by

Hk(M0,S,Ls_),Hklf(M0,S,Ls_),Hk(M0,S,Ls_),Hck(M0,S,Ls_).

They are finite-dimensional Qs_B-vector spaces. The second is the cohomology of the complex of formal infinite sums of cochains with coefficients in Ls_ whose restriction to any compact subset have only finitely many non-zero terms.

Because of (74), duality between homology and cohomology gives rises to canonical isomorphisms of Qs_B-vector spaces for all k:

Hk(M0,S,L-s_)Hk(M0,S,Ls_),Hklf(M0,S,L-s_)Hck(M0,S,Ls_).

Proposition 6.5

If the sij are generic in the sense of (75), then the natural maps induce the following isomorphisms:

Hck(M0,S,Ls_)Hk(M0,S,Ls_) 76
Hk(M0,S,Ls_)Hklf(M0,S,Ls_). 77

Proof

Let j:M0,SM¯0,S denote the open immersion. We claim that the natural map j!Ls_RjLs_ is an isomorphism in the derived category of the category of sheaves on M¯0,S, which amounts to the fact that RjLs_ has zero stalk at any point of M¯0,S. Since M¯0,S is a normal crossing divisor we are reduced, by Künneth, to proving that Ls_ has non-trivial monodromy around each boundary divisor. For a divisor defined by the vanishing of a dihedral coordinate uc=0, monodromy acts by multiplication by exp(-2πisc), since Ls_ is generated by cχS,δuc-sc. Every other boundary divisor on M0,S is obtained from such a divisor by permuting the elements of S. From formula (36) and the action of the symmetric group, it follows that the monodromy of Ls_ around a divisor D defined by a partition S=S1S2 is given by exp(-2πisD) where sD=i,jS1sij=i,jS2sij. It is non-trivial if and only if sDZ, which is equation (75). The first statement follows by applying RΓcRΓ to the isomorphism j!Ls_RjLs_. The second statement is dual to the first.

Remark 6.6

Under the assumptions (75), Artin vanishing and duality imply that all homology and cohomology groups in Proposition 6.5 vanish if in.

The inverse to the isomorphism (77) is sometimes called regularisation. For any dihedral structure δ on S, the function fs_=cχS,δucsc is well-defined on the domain XδM0,S(R) and defines a class [Xδfs_] in Hnlf(M0,S,L-s_). (Strictly speaking, this class is represented by the infinite sum of the simplices of a fixed locally finite triangulation of Xδ, and does not depend on the choice of triangulation). Its image under the regularisation map, assuming (75), defines a class we abusively also denote by [Xδfs_]Hn(M0,S,L-s_). The following result is classical.

Proposition 6.7

Assume that the sij are generic in the sense of (75). Choose three distinct elements a,b,cS. A basis of Hn(M0,S,L-s_) is provided by the classes [Xδfs_], where δ ranges over the set of dihedral structures on S with respect to which abc appear consecutively, and in that order (or the reverse order).

Proof

We may assume that S={1,,n+3}, (a,b,c)=(n+1,n+2,n+3), and work in simplicial coordinates (t1,,tn) by fixing pn+1=1, pn+2=, pn+3=0. These coordinates give an isomorphism of M0,S with the complement of the hyperplane arrangement in An consisting of the hyperplanes {ti=0} and {ti=1} for 1in, and {ti=tj} for 1i<jn. This arrangement is defined over R and the real points of its complement is the disjoint union of the domains Xδ for δ a dihedral structure on S. Such a domain is bounded in Rn if and only if it is of the form {0<tσ(1)<<tσ(n)<1} for some permutation σΣn, i.e., if and only if no simplicial coordinate is adjacent to in the dihedral ordering. Equivalently the points n+1,n+2,n+3 are consecutive and in that order (or its reverse) in the dihedral ordering δ. The proposition is thus a special case of [DT97, Proposition 3.1.4].

Betti pairing

Under the assumptions (75), Poincaré–Verdier duality combined with (76) defines a perfect pairing of Qs_B-vector spaces in cohomology

,B:Hn(M0,S,Ls_)Qs_BHn(M0,S,L-s_)Qs_B,

which is dual to a perfect pairing in homology

,B:Hn(M0,S,L-s_)Qs_BHn(M0,S,Ls_)Qs_B.

If σf-s_ and τfs_ are locally finite representatives for homology classes in Hn(M0,S,Ls_) and Hn(M0,S,L-s_) respectively, then the corresponding pairing

[τfs_],[σf-s_]B

is the number of intersection points (with signs) of σ and τ~ where [τ~fs_] is a regularisation of [τfs_] and τ~ is in general position with respect to σ [KY94a], [KY94b]. The matrix of the cohomological Betti pairing is the inverse transpose of the matrix of the homological Betti pairing.

Algebraic de Rham cohomology

Let Hk(M0,S,s_) denote the algebraic de Rham cohomology of M0,S with coefficients in the algebraic vector bundle (OS,s_) with integrable connection over M0,S×QQs_dR. It is a finite-dimensional Qs_dR-vector space. Let (OSan,s_an) denote the analytic rank one vector bundle with connection on M0,S(C) obtained from (OS,s_). We have an isomorphism

Hk(M0,S,s_)Qs_dRCHk(M0,S(C),s_an), 78

where the right-hand side denotes the cohomology of the complex of global smooth differential forms on M0,S(C) with differential s_. Recall from [BD20, §3] the notation AM¯0,S(logM¯0,S) for the complex of sheaves of smooth forms on M¯0,S with logarithmic singularities along M¯0,S.

Proposition 6.8

Under the assumptions (75) we have a natural isomorphism

Hk(M0,S(C),s_an)Hk(Γ(M¯0,S,AM¯0,S(logM¯0,S)),s_an).

Proof

This is a smooth version of [Del70, Proposition 3.13]. The assumptions of [loc. cit.] are implied by (75) and one can check that its proof can be copied in the smooth setting.

By a classical argument due to Esnault–Schechtman–Viehweg [ESV92], we can replace global logarithmic smooth forms with global algebraic smooth forms and the cohomology group Hk(M0,S,s_) is given, under the assumptions (75), by the cohomology of the complex (ΩSQs_dR,ωs_-). In particular, Hn(M0,S,s_) is simply the quotient of ΩSnQs_dR by the subspace spanned by the elements ωs_φ for φΩSn-1. The following theorem gives a basis of that quotient.

Theorem 6.9

Assume that the sij are generic in the sense of (75). A basis of Hn(M0,S,s_) is provided by the classes of the differential forms

dt1dtnk=1n(tk-tik)

for the tuples (i1,,in) with 0ikk-1 and where we set t0=0.

Proof

This is a special case of [Aom87, Theorem 1].

The following more symmetric basis is more prevalent in the string theory literature. Its elements are called Parke–Taylor factors [PT86]. Therefore we shall refer to it as the Parke–Taylor basis, as opposed to the Aomoto basis of Theorem 6.9. Although the following theorem is frequently referred to in the literature, we could not find a complete proof for it and therefore provide one in “Appendix 7.4”.

Theorem 6.10

Assume that the sij are generic in the sense of (75). A basis of Hn(M0,S,s_) is provided by the classes of the differential forms

dt1dtnk=1n+1(tσ(k)-tσ(k-1))

for permutations σΣn, where we set tσ(0)=0 and tσ(n+1)=1.

Other bases can be found in the literature, e.g., the βnbc bases of Falk–Terao [FT97].

de Rham pairing

Under the assumptions (75) there is an algebraic de Rham version of the intersection pairing, which is a perfect pairing

,dR:Hn(M0,S,s_)Qs_dRHn(M0,S,-s_)Qs_dR

of Qs_dR-vector spaces. This is easily checked after extending the scalars to C by working with smooth de Rham complexes. The only part to check is that this pairing is algebraic, i.e., defined over Qs_dR. Indeed, it can be defined algebraically (see [CM95] for the general case of curves), and computed explicitly for hyperplane arrangements [Mat98], which contains the present situation as a special case.

If ω,νQs_dRΩSn are logarithmic n-forms, let ν~ be a smooth s_-closed n-form on M0,S(C) which represents [ν] and has compact support. Then

[ν],[ω]dR=(2πi)-nM0,S(C)ν~ω.

Our normalisation differs from the one in the literature by the factor of (2πi)-n.

Periods

Since algebraic de Rham cohomology is defined over Qs_dR, we can meaningfully speak of periods. Using (73) we see that integration induces a perfect pairing of complex vector spaces:

Hn(M0,S,L-s_Qs_BC)CHn(M0,S(C),s_an)C[γfs_][ω]γfs_ω

which is well-defined by Stokes’ theorem. By (78) it induces an isomorphism:

compB,dR:Hn(M0,S,s_)Qs_dRCHn(M0,S,Ls_)Qs_BC. 79

We will use the notation compB,dR(s_) when we want to make the dependence on s_ explicit. If we choose a Qs_dR-basis of the left-hand vector space, and a Qs_B-basis of the right-hand vector space, the isomorphism compB,dR(s_) can be expressed as a matrix Ps_, and we will sometimes abusively use the notation Ps_ instead of compB,dR(s_).

Theorem 6.11

(Twisted period relations [KY94a], [CM95]) Assume that the sij are generic in the sense of (75). Let ω,νQs_dRΩSn be logarithmic n-forms giving rise to classes in Hn(M0,S,-s_) and Hn(M0,S,s_) respectively. We have the equality:

(2πi)n[ν],[ω]dR=Ps_[ν],P-s_[ω]B, 80

where the cohomological Betti pairing is naturally extended by C-linearity.

Proof

This follows from the fact that the (iso)morphisms (76) and (77) and Poincaré–Verdier duality are compatible with the comparison isomorphisms.

The reason for the factor (2πi)n in the formula (80) is because of our insistence that the de Rham intersection pairing IdR be algebraic and have entries in Qs_dR.

Proposition 6.12

Assume that the sij are generic in the sense of (75). Let δ be a dihedral structure on S and let ωQs_dRΩSn be a regular logarithmic form on M0,S of top degree. If the inequalities of Proposition 3.5 hold then we have

[Xδfs_],compB,dR[ω]=Xδfs_ω.

Proof

  1. We first prove that the formula holds for any algebraic n-form ω on M0,Sδ with logarithmic singularities along M0,Sδ, provided Re(sc)>0 for every chord cχS,δ. By definition we have for every s_ the formula:
    [Xδfs_],compB,dR[ω]=Xδfs_ω~,
    where ω~ is a global section of AM¯0,Sn(logM¯0,S) with compact support which is cohomologous to ω, i.e., such that ω-ω~=s_ϕ, with ϕ a global section of AM¯0,Sn-1(logM¯0,S). Thus, we need to prove that the integral of fs_s_ϕ=d(fs_ϕ) on Xδ vanishes if Re(sc)>0 for all chords cχS,δ. We note that in general fs_ϕ has singularities along the boundary of Xδ unless Re(sc)>1 for all chords cχS,δ, so that we cannot apply Stokes’ theorem directly. We can write
    ϕ=JϕJcJducuc
    where the sum is over subsets of chords JχS,δ and ϕJ extends to a smooth form on M0,nδ (i.e., has no poles along the boundary of Xδ). By properties of dihedral coordinates, we can furthermore assume that ϕJ=0 if J contains two crossing chords. Indeed, a form cJducuc extends to a regular form on M0,nδ if every chord in J is crossed by another chord in J by (13). It is therefore sufficient to consider a single term given by a set JχS,δ consisting of chords that do not cross. We write ϕ=ϕJ and set fs_=cJucsc so that we have
    d(fs_ϕ)=d(fs_ϕ)cJucscducuc.
    The forgetful maps (18) give rise to a diffeomorphism Xδ(0,1)k×XJ with k=|J| and XJ=Xδ0××Xδk. We can thus write
    Xδd(fs_ϕ)=±(0,1)ki=1kxiscidxixiXJd(fs_ϕ),
    where the xi are the coordinates on (0,1)k, corresponding to the dihedral coordinates uci for J={c1,,ck}. Now the boundary of XJ has components {uc=0} for c a chord that does not cross any chord in J. Thus, if Re(sc)>0 for every chord c we have that fs_, and hence fs_ϕ, vanishes on the boundary of XJ, and the inner integral is zero by Stokes’ theorem for manifolds with corners.
  2. Now, for ω as in the statement of the proposition, let Dc1,,Dcr be the divisors along which ω does not have a pole. Applying the first step of the proof to the product i=1ruci-1ω yields the result.

Example 6.13

Let |S|=4. Then H1(M0,S,L-s,-t)H1lf(M0,S,L-s,-t) is one-dimensional. The locally finite homology is spanned by the class of σxs(1-x)t where σ is the open interval (0, 1). The algebraic de Rham cohomology group H1(M0,S,s_)Qs_dR[ν] is one-dimensional spanned by the class of ν=-νS where

ν=dxx(1-x)ΩS1.

The period matrix Ps_ is the 1×1 matrix whose entry is the beta function:

Ps_=01xs(1-x)tdxx(1-x)=β(s,t)=Γ(s)Γ(t)Γ(s+t). 81

For any small ε>0, a representative for the regularisation of [σxs(1-x)t] is

(S0(ε)e2πis-1+[ε,1-ε]-S1(ε)e2πit-1)xs(1-x)t

where Si(ε) denotes the small circle of radius ε winding positively around i. From this one easily deduces the intersection product with the class of σx-s(1-x)-t. It is

[σxs(1-x)t],[σx-s(1-x)-t]B=1-e2πi(s+t)(1-e2πis)(1-e2πit)=i2sin(π(s+t))sin(πs)sin(πt)· 82

See, e.g., [CM95], [KY94a, §2], or [MY03, §2]. Dually:

[σxs(1-x)t],[σx-s(1-x)-t]B=2isin(πs)sin(πt)sinπ(s+t)·

On the other hand, the de Rham intersection pairing [Mat98] is

[ν],[ν]dR=1s+1t· 83

In this case, equation (80) reads

2πi(1s+1t)=β(-s,-t)β(s,t)2isin(πs)sin(πt)sinπ(s+t) 84

using (82) and (83), as observed in [CM95], or in terms of the gamma function:

Γ(s)Γ(t)Γ(s+t)Γ(-s)Γ(-t)Γ(-s-t)=-π(s+t)sin(π(s+t))ssin(πs)tsin(πt)· 85

This can easily be deduced from the well-known functional equation for the gamma function Γ(s)Γ(-s)=-πssin(πs), and is in fact equivalent to it (set t=-s/2).

Self-duality

It is convenient to reformulate the above relations as a statement about self-duality. Consider the object

MdR=Hn(M0,S,s_)Hn(M0,S,-s_)MB=Hn(M0,S,Ls_)Hn(M0,S,L-s_)

and denote the comparison

P=Ps_P-s_:MdRQs_dRCMBQs_BC

The results in the previous section can be summarised by saying that the triple of objects (MdR,MB,P) is self-dual. In other words, the Betti and de Rham pairings induce isomorphisms

IdR:MdRMdRandIB:MBMB

which are compatible with the comparison isomorphism P. With these notations, equation (80) can be written in the simpler form:

(2πi)nIdR=PIBP. 86

Single-Valued Periods for Cohomology with Coefficients

We fix a solution (sij) of the momentum conservation equations over the complex numbers.

Complex conjugation and the single-valued period map

We can define and compute a period pairing on de Rham cohomology classes by transporting complex conjugation which is the anti-holomorphic diffeomorphism:

conj:M0,S(C)M0,S(C). 87

Since it reverses the orientation of simple closed loops, and since a rank one local system on M0,S(C) is determined by a representation of the abelian group H1(M0,S(C)) we see that we have an isomorphism of local systems:

conjLs_L-s_. 88

We thus get a morphism of local systems on M0,S(C):

Ls_conjconjLs_conjL-s_,

which at the level of cohomology induces a morphism of Qs_B-vector spaces

F:Hn(M0,S,Ls_)Hn(M0,S,L-s_).

We call F the real Frobenius or Frobenius at the infinite prime. We will use the notation F(s_) when we want to make dependence on s_ explicit. One checks that the Frobenius is involutive: F(-s_)F(s_)=id.

Remark 7.1

The isomorphism (88) is induced by the trivialisation of the tensor product conjLs_Ls_ given by the section

gs_=i<j|pj-pi|-2sij=i<j(pj¯-pi¯)-sij·i<j(pj-pi)-sij.

Thus, the action of real Frobenius on homology

F:Hn(M0,S,Ls_)Hn(M0,S,L-s_)

is given by the formula

σi<j(pj-pi)-sijσ¯i<j(pj¯-pi¯)-sijgs_-1=σ¯i<j(pj-pi)sij.

Remark 7.2

A morphism similar to F was considered in [HY99] and leads to similar formulae but has a different definition. Our definition only uses the action of complex conjugation on the complex points of M0,S, whereas the definition in [loc. cit.] conjugates the field of coefficients of the local systems. Note that our definition does not require the sij to be real.

Definition 7.3

The single-valued period map is the C-linear isomorphism

s:Hn(M0,S,s_)Qs_dRCHn(M0,S,-s_)Qs_dRC

defined as the composite

s=compB,dR-1(-s_)(Fid)compB,dR(s_).

In other words, it is defined by the following commutative diagram:

graphic file with name 220_2021_3969_Equ333_HTML.gif

The single-valued period map can be computed explicitly by choosing a Qs_dR-bases {[ω]} and {[ν]} for Hn(M0,S,-s_) and Hn(M0,S,s_) respectively and a Qs_B-basis {[σfs_]} for Hn(M0,S,L-s_). In these bases, the isomorphism (79) is represented by a matrix Ps_ with entries

Ps_([σfs_],[ν])=σfs_ν.

By Remark 7.1 the entries of FP-s_ are

(FP-s_)([σfs_],[ω])=σ¯f-s_ω.

The single-valued period matrix (the matrix of s) is then the product

P-s_-1(FPs_)=(FP-s_)-1Ps_.

This formula is often impractical because one needs to compute all the entries of the period matrix in order to compute any single entry of the single-valued period matrix.

Example 7.4

With the notation of Example 6.13 we have σ¯=σ since (0, 1) is real, and the single-valued period matrix is

(FP-s_)-1Ps_=β(-s,-t)-1β(s,t)=Γ(s)Γ(t)Γ(-s-t)Γ(s+t)Γ(-s)Γ(-t).

The single-valued period pairing via the Betti pairing

If the sij are generic in the sense of (75), the single-valued period map and the de Rham pairing induce a single-valued period pairing

HdRn(M0,S,s_)Qs_dRHdRn(M0,S,s_)C

given for ω,νQs_dRΩSn by the formula

[ν][ω][ν],s[ω]dR=[ν],P-s_-1FPs_[ω]dR.

One can use the compatibility between the de Rham and the Betti pairings to express the single-valued pairing in terms of the latter.

Proposition 7.5

Assume that the sij are generic in the sense of (75). Let ω,νQs_dRΩSn and denote by [ω], [ν] their classes in Hn(M0,S,s_). The corresponding single-valued period is given by the formula

[ν],s[ω]dR=(2πi)-nPs_[ν],FPs_[ω]B

and can be computed explicitly by a sum

(2πi)-n[σf-s_][τfs_][τfs_],[σf-s_]Bτfs_νσ¯fs_ω,

where [σf-s_] and [τfs_] range over a basis of Hn(M0,S,Ls_) and Hn(M0,S,L-s_) respectively, and [σf-s_],[τfs_] are the dual bases.

Proof

We have

[ν],s[ω]dR=[ν],P-s_-1FPs_[ω]dR=(2πi)-nPs_[ν],FPs_[ω]B,

where the first equality is the definition of the single-valued period map, and the second equality follows from Theorem 6.11. The second formula follows from the definition of Ps_ and Remark 7.1.

Example 7.6

Following up on Example 7.4 and using (83) we see that we have

[ν],s[ν]dR=β(-s,-t)-1β(s,t)1s+1t=-Γ(s)Γ(t)Γ(1-s-t)Γ(s+t)Γ(1-s)Γ(1-t), 89

where we have used Γ(-x)=-xΓ(1-x). Now using (82), Proposition 7.5 reads

[ν],s[ν]dR=12πi2isin(πs)sin(πt)sin(π(s+t))β(s,t)2=-1πsin(πs)sin(πt)sin(π(s+t))Γ(s)Γ(t)Γ(s+t)2. 90

One deduces (90) from (89), and vice versa, by applying the functional equation (84).

An integral formula for single-valued periods

The single-valued period map is a transcendental comparison isomorphism that is naturally interpreted at the level of analytic de Rham cohomology via the isomorphism (78).

Lemma 7.7

In analytic de Rham cohomology, the single-valued period map is induced by the morphism of smooth de Rham complexes

san:(AM0,S(C),s_an)conj(AM0,S(C),-s_an)

given on the level of sections by

AM0,S(C)(U)ωi<j|pj-pi|2sijconj(ω)AM0,S(C)(U¯).

Proof

Recall the notation gs_=i<j|pj-pi|2sij. We first check that san is a morphism of complexes:

-s_an(san(ω))=-s_an(gs_conj(ω))=gs_i<jsijdlog(pj-pi)+i<jsijdlog(pj¯-pi¯)conj(ω)+d(conj(ω))-i<jsijdlog(pj-pi)(gs_conj(ω))=gs_i<jsijdlog(pj¯-pi¯)conj(ω)+d(conj(ω))=gs_conji<jsijdlog(pj-pi)ω+dω=san(s_an(ω)).

On the level of horizontal sections, we compute:

sani<j(pj-pi)-sij=gs_i<j(pj¯-pi¯)-sij=i<j(pj-pi)sij.

Thus, san induces the isomorphism Ls_conjL-s_ and the result follows.

We can now give an explicit formula for single-valued periods in the case of forms with logarithmic singularities.

Theorem 7.8

Assume the sij are generic in the sense of (75). Let ωQs_dRΩSn and let νS be as in Definition 2.7. Then νS, ω define de Rham cohomology classes

[νS],[ω]HdRn(M0,S,s_).

If the inequalities stated in Proposition 3.6 hold, then the single-valued period of [νS][ω] is given by the absolutely convergent integral

[νS],s[ω]dR=(2πi)-nM¯0,S(C)i<j|pj-pi|2sijνSω¯.

Proof

Recall the notation gs_=i<j|pj-pi|2sij. By definition and Lemma 7.7 we have the formula, valid for every generic (sij):

[νS],s[ω]dR=(2πi)-nM¯0,S(C)gs_νSω~¯,

where ω~ is a global section of AM¯0,Sn(logM¯0,S) with compact support which is cohomologous to ω, i.e., such that ω-ω~=s_ϕ, with ϕ a global section of AM¯0,Sn-1(logM¯0,S). We need to prove that the integral of gs_νSs_ϕ¯=±d(gs_νSϕ¯) on M¯0,S(C) vanishes under the stated assumptions on s_. By a partition of unity argument, we can assume that ϕ has support in a local chart on M¯0,S. For this, let (z1,,zn) denote local coordinates M¯0,S taking values in a polydisk Δn={|zi|<1}, with respect to which M¯0,S is a union of coordinate hyperplanes {zi=0}. We can assume the support of ϕ is contained within this polydisk. By renumbering the coordinates if necessary, we let z1,,zr denote the equations of components of M¯0,S at finite distance and zr+1,,zr+s denote the coordinates corresponding to components at infinite distance (relative to the fixed dihedral structure on S). In these coordinates we have

gs_=ai=1r|zi|2sii=r+1r+s|zi|2ti

where a is a smooth function on Δn, si is one of the Mandelstam variables sc, and ti is a linear combination of Mandelstam variables sc with coefficients in {0,1,-1}. Since νS only has simple poles located along the divisors at finite distance, we can write

νS=bdz1z1dzrzrdzr+1dzn

where b is a smooth function on Δn. Since ϕ has degree n-1, by linearity in ϕ, we can assume that there is a single coordinate zp such that dzp does not appear in ϕ. There are three cases to consider, depending on whether this coordinate is away from M¯0,S, or corresponds to a component at finite or infinite distance. In each case we use the Leibniz rule to compute d(gs_νSϕ¯).

  1. We have p>r+s. Without loss of generality, assume that p=n. Then ϕ has the form
    ϕ=cdz1z1dzr+szr+sdzr+s+1dzn-1
    where c is a smooth function on Cn. Since c has support in Δn, it is enough to show that the following integral vanishes:
    Δn-1i=1r|zi|2sidzidz¯izizi¯i=r+1r+s|zi|2tidzidz¯izi¯i=r+s+1n-1dzidzi¯|zn|1d(fdzn),
    where f=abc¯ is a smooth function on Cn with support in Δn. The inner integral vanishes by Stokes’ theorem and we are done.
  2. We have pr. Without loss of generality, let p=r. Then ϕ has the form
    ϕ=cdz1z1dzr-1zr-1dzr+1zr+1dzr+szr+sdzr+s+1dzn
    where c is a smooth function on Cn with support in Δn. It is enough to show that the following integral vanishes:
    Δn-1i=1r-1|zi|2sidzidz¯izizi¯i=r+1r+s|zi|2tidzidz¯izi¯i=r+s+1ndzidzi¯|zr|1df|zr|2srdzrzr,
    where f=abc¯ is a smooth function on Cn with support in Δn. The inner integral is the limit as ε goes to zero of the same integral over {ε|zr|1}, which by Stokes and changing to polar coordinates evaluates to
    |zr|=εf|zr|2srdzrzr=ε2sr02πifdθ.
    This tends to zero when ε goes to zero if Re(sr)>0.
  3. We have r<pr+s. We can assume that p=r+s and so ϕ has the form
    ϕ=cdz1z1dzr+s-1zr+s-1dzr+s+1dzn
    where c is a smooth function on Cn with support in Δn. We want to prove that the following integral vanishes:
    Δn-1i=1r|zi|2sidzidz¯izizi¯i=r+1r+s-1|zi|2tidzidz¯izi¯i=r+s+1ndzidzi¯|zr+s|1df|zr+s|2tr+sdzr+s,
    where f=abc¯ is a smooth function on Cn with support in Δn. The inner integral is the limit as ε goes to zero of the same integral over {ε|zr+s|1}, which by Stokes and changing to polar coordinates evaluates to
    |zr+s|=εf|zr+s|2tr+sdzr+s=ε2tr+s+102πifeiθdθ.
    This goes to zero when ε goes to zero if 2Re(tr+s)>-1, which is a consequence of the inequalities stated in Proposition 3.6 as in the proof of that Proposition.

Therefore, the integral of gs_νSs_ϕ¯ vanishes if the inequalities stated in Proposition 3.6 hold: namely -12<Re(sc)<1N2 for all c and Re(sc)>0 for every divisor Dc along which ω has a pole.

Example 7.9

In the case of the beta function, with ν=-νS=-dxx(1-x), Theorem 7.8 reads:

[ν],s[ν]dR=12πiP1(C)|z|2s|1-z|2tνν¯=12πiP1(C)|z|2s-2|1-z|2t-2dzdz¯,

which equals -βC(s,t).

Double copy formula

By equating the two expressions for the single-valued period given in Proposition 7.5 and Theorem 7.8 we obtain an equality that expresses a volume integral as a quadratic expression in ordinary period integrals.

Corollary 7.10

Under the assumptions of Theorem 7.8 we have the equality:

M¯0,S(C)i<j|pj-pi|2sijνSω¯=[σf-s_][τfs_][τfs_],[σf-s_]Bτfs_νSσ¯fs_ω,

where [σf-s_] and [τfs_] range over a basis of Hn(M0,S,Ls) and Hn(M0,S,L-s_) respectively, and [σf-s_],[τfs_] are the dual bases.

This formula bears very close similarity to the KLT formula [KLT86], and makes it apparent that the ‘KLT kernel’ should coincide with the Betti intersection pairing on twisted cohomology, which is the inverse transpose of the intersection pairing on twisted homology. Indeed, Mizera has shown in [Miz17] that the KLT kernel indeed coincides with the inverse transpose matrix of the intersection pairing.

Example 7.11

Examples 7.6 and 7.9 give rise to the equality

βC(s,t)=-12πi2isin(πs)sin(πt)sin(π(s+t))β(s,t)2,

which is an instance of Corollary 7.10.

Acknowledgements

This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement No. 724638). Both authors thank the IHES for hospitality. The second author was partially supported by ANR Grant ANR-18-CE40-0017. This paper was initiated during the trimester “Periods in number theory, algebraic geometry and physics” which took place at the HIM Bonn in 2018, to which both authors offer their thanks. Many thanks to Andrey Levin, whose talk during this programme on the dilogarithm inspired this project, and also to Federico Zerbini for discussions. The second author also thanks Mike Falk for discussions on cohomology with coefficients and hyperplane arrangements.

Appendix A: The Parke–Taylor basis

In this appendix we prove the following theorem, which is Theorem 6.10 from the main body of the paper.

Theorem A.1

Assume that the si,j are generic in the sense of (75). A basis of Hn(M0,n+3,s_) is provided by the classes of the differential forms

dt1dtnk=1n+1(tσ(k)-tσ(k-1)) 91

for permutations σΣn, where we set tσ(0)=0 and tσ(n+1)=1.

Working with the Configuration Space of Points in C

We consider the configuration space

Conf(n+2,C)={(z0,z1,,zn+1)Cn+2,zizj}

and its (rational algebraic de Rham) cohomology algebra

A=HdR(Conf(n+2,C)).

By a classical result of Arnol’d [Arn69], it is generated by the (classes of the) forms

ωi,j=dlog(zj-zi).

The natural (diagonal) C-action on Conf(n+2,C) induces a linear map in cohomology:

:AA-1HdR1(C)A-1.

Compatibility with the cup-product and the Koszul sign rule implies that it is a graded derivation, i.e., that it satisfies the Leibniz rule

(ab)=(a)b+(-1)|b|a(b)

for b homogeneous of degree |b|. It is uniquely determined by (ωi,j)=1 for all ij and is given more generally by the formula:

(ωi1,j1ωir,jr)=k=1r(-1)k-1ωi1,j1ωik,jk^ωir,jr.

We will make use of the following relations.

Lemma A.2

We have

ωi1,j1ωir,jr=0and(ωi1,j1ωir,jr)=0

if there is a cycle in the graph with vertices {0,,n+1} and edges {i1,j1},,{ir,jr}.

Proof

The second relation follows from the first, which is already satisfied at the level of differential forms.

A special case is the classical Arnol’d relations (ωi,jωi,kωj,k)=0, which generate all the relations among the generators ωi,j in the algebra A [Arn69].

The (homological) complex (A,) is contractible. A contracting homotopy is given, for instance, by multiplication by the generator ω0,1.

There is a natural quotient morphism (by the affine group CC):

Conf(n+2,C)M0,n+3,(z0,,zn+1)(z1,,zn+1,,z0),

defined so that the simplicial coordinates t0=0,t1,,tn,tn+1=1 on M0,n+3 are related to the coordinates zi by the formula

ti=zi-z0zn+1-z0·

The pullback by this quotient identifies Ω=HdR(M0,n+3) with the subalgebra ker()A. Since (A,) is contractible we have ker()=Im() and we get a short exact sequence:

0ΩAΩ-10. 92

We now move to cohomology with coefficients. A solution s_ to the momentum conservation equations is completely determined by the complex numbers si,j for 0i<jn+1, where we set s0,j:=sj,n+3 as in Sect. 3.2, subject to the single relation:

0i<jn+1sij=0. 93

Denote the pullback of the Koba–Nielsen form to Conf(n+2,C) by the same symbol:

ωs_=0i<jn+1si,jωi,j.

In the remainder of this appendix we extend the scalars from Q to the field Qs_dR=Q(si,j) generated by the si,j inside C. In order to keep the notations simple, we continue to write A and Ω for AQQs_dR and ΩQQs_dR, respectively.

The short exact sequence (92) induces a short exact sequence of complexes :

0(Ω,ωs_)(A,ωs_)(Ω-1,ωs_)0.

This is because (ωs_)=0, which follows from equation (93). Assume that s_ is generic (75). Then Hk(Ω,ωs_)=0 for kn (Remark 6.6) and the long exact sequence in cohomology shows that the morphism induced by in top degree cohomology

Hn+1(A,ωs_)Hn(Ω,ωs_)

is an isomorphism. An easy computation (see, e.g., [Miz17, Claim 3.1]) shows that the Parke–Taylor form (91) from Theorem A.1 is, up to the sign sgn(σ), the image by of the form

ωσ=ω0,σ(1)ωσ(1),σ(2)ωσ(n-1),σ(n)ωσ(n),n+1An+1,

for σΣn. Thus, a restatement of Theorem A.1 is as follows:

Theorem A.3

Assume that the si,j are generic in the sense of (75). Then a basis of

Hn+1(A,ωs_)=An+1/(Anωs_)

is provided by the classes of the forms ωσ, for permutations σΣn.

The rest of this appendix is devoted to the proof of this theorem.

A Basis of An+1

We slightly extend the notation ωσ to permutations σΣn+1:

ωσ:=ω0,σ(1)ωσ(1),σ(2)ωσ(n-1),σ(n)ωσ(n),σ(n+1)An+1.

Proposition A.4

The cohomology classes ωσ, for σΣn+1, are a basis of An+1.

Proof

The dimension of An+1 is (n+1)! by [Arn69], so it is enough to prove that the ωσ are linearly independent in An+1. For 1in+1 we have a residue morphism along {z0=zi}:

Resi:HdRn+1(Conf(n+2,C))HdRn(Conf(n+1,C)),

where in the target Conf(n+1,C) consists of tuples (z0,,zi^,,zn+1). It satisfies:

Resi(ωσ)=0ifσ(1)i;ωσifσ(1)=i,

where in the second case we implicitly use the natural bijection

ΣnBij({2,,n+1},{1,,i^,,n+1}).

The result follows by induction on n, since the case n=0 is trivial.

We now let FkAn+1 denote the subspace of An+1 spanned by the basis elements ωσ, for σΣn+1 such that σ-1(n+1)k. It forms a decreasing filtration, where F1An+1=An+1, Fn+2An+1=0, and Fn+1An+1 is spanned by the ωσ for σΣn.

Lemma A.5

Any element of An+1 which is the exterior product of a form with an element of the following type

ω0,i1ωi1,i2ωik-2,ik-1ωik-1,n+1

lies in FkAn+1.

Proof

To a graph Γ with set of vertices {0,,n+1} and (n+1) edges we associate a monomial ωΓ (well-defined up to a sign) obtained by multiplying the generators ωi,j together for {i,j} an edge of Γ. What we need to prove is that ωΓFkAn+1 if Γ contains a path of length k between the vertices 0 and n+1. If Γ contains a cycle then ωΓ=0 by Lemma A.2 and there is nothing to prove. Since Γ has n+2 vertices and n+1 edges we can thus assume that it is a tree. If we choose 0 to be the root of Γ then all edges inherit a preferred orientation (from the root to the leaves). We let δΓ(i) denote the distance between the vertices 0 and i in Γ and set

δΓ=i=1n+1δΓ(i).

      We have δΓ1+2++(n+1) and the equality holds exactly when Γ is linear, i.e., has only one leaf. We prove by decreasing induction on δΓ the statement: ωΓFkAn+1 if δΓ(n+1)k. If Γ is linear then ωΓ is one of the basis elements ωσ for some σΣn+1 such that σ-1(n+1)k and thus ωΓFkAn+1 by definition. Now assume that Γ is not linear and let a be a vertex of Γ with two children b1 and b2 (which means that there is an edge from a to b1 and an edge from a to b2 in Γ). Then we can use the Arnol’d relation

ωa,b1ωa,b2=ωa,b1ωb1,b2-ωa,b2ωb1,b2

and deduce a relation

ωΓ=±ωΓ2±ωΓ1

where Γs is obtained from Γ by deleting the edge {a,bs} and adding the edge {b1,b2}, for s{1,2}. One easily sees that we have, for all i{1,,n+1}, δΓs(i)δΓ(i), and δΓs>δΓ, for s{1,2}. One can thus apply the induction hypothesis to Γ1 and Γ2, which completes the induction step and the proof.

The Proof

The filtration Fk on An+1 induces a decreasing filtration denoted by the same symbol on the quotient Hn+1(A,ωs_)=An+1/(Anωs_).

Proposition A.6

For every k{1,,n} we have

FkHn+1(A,ωs_)Fk+1Hn+1(A,ωs_).

Proof

Let us fix a permutation σΣn such that σ-1(n+1)=k{1,,n}. We set is=σ(s) for every s{1,,n+1} and i0=0, so that we have

ωσ=XY

where we set:

X=ωi0,i1ωi1,i2ωik-1,n+1andY=ωn+1,ik+1ωin-1,inωin,in+1.

      We have the relation, in Hn+1(A,ωs_):

Xωs_(Y)=0.

      We note that if {i,j}={ia,ib} for 0a<bk then we have Xωi,j=0 by Lemma A.2. Let P denote the set of remaining pairs of indices. We then get the relation:

{i,j}Psi,jXωi,j(Y)=0.

      The Leibniz rule gives

ωi,j(Y)=Y-(ωi,jY)

and we can rewrite the relation as:

{i,j}Psi,jωσ={i,j}Psi,jX(ωi,jY). 94

      We now claim that the term X(ωi,jY) is in Fk+1 for all {i,j}P. There is one easy case: if {i,j}={ia,ib} for ka<bn+1 then we have (ωijY)=0 by Lemma A.2. Thus, we only have to treat the case where {i,j}={ia,ib} for 0ak and k+1bn+1.

We proceed by decreasing induction on a. If a=k then the first case applies and we are done. If a<k then the Leibniz rule implies:

X(ωia,ibY)-X(ωia+1,ibY)=X(ωia,ibωia+1,ib)(Y).

We now set

X=ωi0,i1ωia-1,iaωia+1,ia+2ωik-1,n+1

so that we have X=±ωia,ia+1X. We thus get

X(ωia,ibY)-X(ωia+1,ibY)=±Xωia,ibωia+1,ib(Y),

where we have used the Arnol’d relation (ωia,ia+1ωia,ibωia+1,ib)=0 in the form: ωia,ia+1(ωia,ibωia+1,ib)=±ωia,ibωia+1,ib. Now Lemma A.5 applied to Xωia,ibωia+1,ib(Y) and the induction hypothesis respectively imply that

Xωia,ibωia+1,ib(Y)andX(ωia+1,ibY)

are in Fk+1Hn+1(A,ωs_), which implies that it is also the case for X(ωia,ibY). This concludes the induction step and the proof by induction. Returning to (94) we see that we have SωσFk+1 for

S={i,j}Psi,j=-0a<bksia,ib,

where we have used equation (93). Thus S0 by the genericity assumption, and ωσFk+1, which finishes the proof of the proposition.

We can now conclude with the proof of Theorem A.3. By Proposition A.6 we have

Hn+1(A,ωs_)=F1Hn+1(A,ωs_)=Fn+1Hn+1(A,ωs_)

which is spanned by the ωσ for σΣn. Since the dimension of Hn+1(A,ωs_) is n! if the si,j are generic, this implies that the ωσ are a basis, and Theorem A.3 is proved. Theorem A.1 follows as explained in Sect. A.1.

Footnotes

1

Strictly speaking, this is not renormalisation in the physical sense since there are no ultraviolet divergences in the perturbative superstring amplitudes.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Francis Brown, Email: francis.brown@all-souls.ox.ac.uk.

Clément Dupont, Email: clement.dupont@umontpellier.fr.

References

  • [Aom87].Aomoto K. Gauss–Manin connection of integral of difference products. J. Math. Soc. Jpn. 1987;39(2):191–208. doi: 10.2969/jmsj/03920191. [DOI] [Google Scholar]
  • [Arn69].Arnol’d VI. The cohomology ring of the group of dyed braids. Mat. Zametki. 1969;5:227–231. [Google Scholar]
  • [BCS10].Brown F, Carr S, Schneps L. The algebra of cell-zeta values. Compos. Math. 2010;146(3):731–771. doi: 10.1112/S0010437X09004540. [DOI] [Google Scholar]
  • [BD19].Brown, F., Dupont, C.: Lauricella hypergeometric functions, unipotent fundamental groups of the punctured Riemann sphere, and their motivic coactions. arXiv:1907.06603 (2019)
  • [BD20].Brown, F., Dupont, C.: Single-valued integration and double copy. J für die reine und angewandte Mathematik (Crelles Journal) 2020. 10.1515/crelle-2020-0042 (2020)
  • [BK13].Brown F, Kreimer D. Angles, scales and parametric renormalization. Lett. Math. Phys. 2013;103(9):933–1007. doi: 10.1007/s11005-013-0625-6. [DOI] [Google Scholar]
  • [Bog16].Bogner C. MPL—a program for computations with iterated integrals on moduli spaces of curves of genus zero. Comput. Phys. Commun. 2016;203:339–353. doi: 10.1016/j.cpc.2016.02.033. [DOI] [Google Scholar]
  • [Bro09].Brown, F.: Multiple zeta values and periods of moduli spaces M¯0,n. Ann. Sci. Éc. Norm. Supér. (4) 42(3), 371–489 (2009)
  • [Bro14a].Brown F. Single-valued motivic periods and multiple zeta values. Forum Math. Sigma. 2014;2:e25–e37. doi: 10.1017/fms.2014.18. [DOI] [Google Scholar]
  • [Bro14b].Brown, F.: Motivic periods and P1\{0,1,}. Proceedings of the ICM Seoul 2014, vol. II, Kyung Moon Sa, Seoul, 295–318 (2014)
  • [Bro17].Brown F. Notes on motivic periods. Commun. Number Theory Phys. 2017;11(3):557–655. doi: 10.4310/CNTP.2017.v11.n3.a2. [DOI] [Google Scholar]
  • [BSST14].Broedel J, Schlotterer O, Stieberger S, Terasoma T. All order α-expansion of superstring trees from the Drinfeld associator. Phys. Rev. D. 2014;89:066014. doi: 10.1103/PhysRevD.89.066014. [DOI] [Google Scholar]
  • [CM95].Cho K, Matsumoto K. Intersection theory for twisted cohomologies and twisted Riemann’s period relations. I. Nagoya Math. J. 1995;139:67–86. doi: 10.1017/S0027763000005304. [DOI] [Google Scholar]
  • [Del70].Deligne P. Équations Différentielles à Points Singuliers Réguliers. Berlin: Springer; 1970. [Google Scholar]
  • [Del71].Deligne P. Théorie de Hodge. II. Inst. Hautes Études Sci. Publ. Math. 1971;40:5–57. doi: 10.1007/BF02684692. [DOI] [Google Scholar]
  • [DG05].Deligne P, Goncharov AB. Groupes fondamentaux motiviques de Tate mixte. Ann. Sci. École Norm. Sup. (4) 2005;38(1):1–56. doi: 10.1016/j.ansens.2004.11.001. [DOI] [Google Scholar]
  • [DT97].Douai A, Terao H. The determinant of a hypergeometric period matrix. Invent. Math. 1997;128(3):417–436. doi: 10.1007/s002220050146. [DOI] [Google Scholar]
  • [Dup18].Dupont C. Odd zeta motive and linear forms in odd zeta values. Compos. Math. 2018;154(2):342–379. doi: 10.1112/S0010437X17007588. [DOI] [Google Scholar]
  • [DV17].Dupont C, Vallette B. Brown’s moduli spaces of curves and the gravity operad. Geom. Topol. 2017;21(5):2811–2850. doi: 10.2140/gt.2017.21.2811. [DOI] [Google Scholar]
  • [ESV92].Esnault H, Schechtman V, Viehweg E. Cohomology of local systems on the complement of hyperplanes. Invent. Math. 1992;109:557–561. doi: 10.1007/BF01232040. [DOI] [Google Scholar]
  • [FT97].Falk M, Terao H. βnbc-bases for cohomology of local systems on hyperplane complements. Trans. Am. Math. Soc. 1997;349(1):189–202. doi: 10.1090/S0002-9947-97-01844-8. [DOI] [Google Scholar]
  • [GM04].Goncharov AB, Manin YuI. Multiple ζ-motives and moduli spaces M¯0,n. Compos. Math. 2004;140(1):1–14. doi: 10.1112/S0010437X03000125. [DOI] [Google Scholar]
  • [GSW12].Green, M.B., Schwarz, J.H., Witten, E.: Superstring Theory. Introduction, vol. 1, Anniversary Cambridge University Press, Cambridge (2012)
  • [HY99].Hanamura M, Yoshida M. Hodge structure on twisted cohomologies and twisted Riemann inequalities I. Nagoya Math. J. 1999;154:123–139. doi: 10.1017/S0027763000025344. [DOI] [Google Scholar]
  • [KLT86].Kawai H, Lewellen DC, Tye S-HH. A relation between tree amplitudes of closed and open strings. Nucl. Phys. B. 1986;269(1):1–23. doi: 10.1016/0550-3213(86)90362-7. [DOI] [Google Scholar]
  • [KN69].Koba Z, Nielsen HB. Generalized Veneziano model from the point of view of manifestly crossing-invariant parametrization. Z. Physik. 1969;229:243–263. doi: 10.1007/BF01396251. [DOI] [Google Scholar]
  • [KY94a].Kita M, Yoshida M. Intersection theory for twisted cycles. Mathematische Nachrichten. 1994;166(1):287–304. doi: 10.1002/mana.19941660122. [DOI] [Google Scholar]
  • [KY94b].Kita M, Yoshida M. Intersection theory for twisted cycles II—degenerate arrangements. Mathematische Nachrichten. 1994;168(1):171–190. doi: 10.1002/mana.19941680111. [DOI] [Google Scholar]
  • [Mat98].Matsumoto K. Intersection numbers for logarithmic k-forms. Osaka J. Math. 1998;35(4):873–893. [Google Scholar]
  • [Miz17].Mizera S. Combinatorics and topology of Kawai–Lewellen–Tye relations. J.High Energy Phys. 2017;2017(8):97. doi: 10.1007/JHEP08(2017)097. [DOI] [Google Scholar]
  • [MY03].Mimachi K, Yoshida M. Intersection numbers of twisted cycles and the correlation functions of the conformal field theory. Commun. Math. Phys. 2003;234:339–358. doi: 10.1007/s00220-002-0766-4. [DOI] [Google Scholar]
  • [OZ01].Ohno Y, Zagier D. Multiple zeta values of fixed weight, depth, and height. Indagationes Mathematicae. 2001;12(4):483–487. doi: 10.1016/S0019-3577(01)80037-9. [DOI] [Google Scholar]
  • [Pan15].Panzer E. Algorithms for the symbolic integration of hyperlogarithms with applications to Feynman integrals. Comput. Phys. Commun. 2015;188:148–166. doi: 10.1016/j.cpc.2014.10.019. [DOI] [Google Scholar]
  • [PT86].Parke SJ, Taylor TR. Amplitude for n-Gluon Scattering. Phys. Rev. Lett. 1986;56:2459. doi: 10.1103/PhysRevLett.56.2459. [DOI] [PubMed] [Google Scholar]
  • [Sha70].Shapiro JA. Electrostatic analogue for the Virasoro model. Phys. Lett. B. 1970;33(5):361–362. doi: 10.1016/0370-2693(70)90255-8. [DOI] [Google Scholar]
  • [SS13].Schlotterer O, Stieberger S. Motivic multiple zeta values and superstring amplitudes. J. Phys. A Math. Theor. 2013;46(47):475401. doi: 10.1088/1751-8113/46/47/475401. [DOI] [Google Scholar]
  • [SS19].Schlotterer O, Schnetz O. Closed strings as single-valued open strings: a genus-zero derivation. J. Phys. A Math. Theor. 2019;52(4):045401. doi: 10.1088/1751-8121/aaea14. [DOI] [Google Scholar]
  • [ST14].Stieberger S, Taylor TR. Closed string amplitudes as single-valued open string amplitudes. Nucl. Phys. B. 2014;881:269–287. doi: 10.1016/j.nuclphysb.2014.02.005. [DOI] [Google Scholar]
  • [Sti14].Stieberger S. Closed superstring amplitudes, single-valued multiple zeta values and the Deligne associator. J. Phys. A Math. Theor. 2014;47(15):155401. doi: 10.1088/1751-8113/47/15/155401. [DOI] [Google Scholar]
  • [Ter02].Terasoma T. Selberg integrals and multiple zeta values. Compos. Math. 2002;133(1):1–24. doi: 10.1023/A:1016377828316. [DOI] [Google Scholar]
  • [Ven68].Veneziano G. Construction of a crossing-symmetric, Regge-behaved amplitude for linearly rising trajectories. Il Nuovo Cimento A (1965-1970) 1968;57(1):190–197. doi: 10.1007/BF02824451. [DOI] [Google Scholar]
  • [Vir69].Virasoro MA. Alternative constructions of crossing-symmetric amplitudes with Regge behavior. Phys. Rev. 1969;177:2309–2311. doi: 10.1103/PhysRev.177.2309. [DOI] [Google Scholar]
  • [VZ18].Vanhove, P., Zerbini, F.: Closed string amplitudes from single-valued correlation functions. arXiv:1812.03018 (2018)
  • [Wit12].Witten, E.: Superstring Perturbation Theory Revisited. arXiv:1209.5461 (2012)

Articles from Communications in Mathematical Physics are provided here courtesy of Springer

RESOURCES