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

aainstitutetext: Bethe Center for Theoretical Physics, Universität Bonn, D-53115, Germanybbinstitutetext: Skobeltsyn Institute of Nuclear Physics of Moscow State University, Moscow 119992, Russian Federationccinstitutetext: Moscow Center for Fundamental and Applied Mathematics, Moscow 119992, Russian Federationddinstitutetext: Rudolf Peierls Centre for Theoretical Physics, University of Oxford, Clarendon Laboratory, Parks Road, Oxford OX1 3PU

Analytic results for two-loop planar master integrals for Bhabha scattering

Claude Duhr b,c    Vladimir A. Smirnov d    Lorenzo Tancredi [email protected] [email protected] [email protected]
Abstract

We analytically evaluate the master integrals for the second type of planar contributions to the massive two-loop Bhabha scattering in QED using differential equations with canonical bases. We obtain results in terms of multiple polylogarithms for all the master integrals but one, for which we derive a compact result in terms of elliptic multiple polylogarithms. As a byproduct, we also provide a compact analytic result in terms of elliptic multiple polylogarithms for an integral belonging to the first family of planar Bhabha integrals, whose computation in terms of polylogarithms was addressed previously in the literature.

Keywords:
Feynman integrals, multiple polylogarithms, elliptic polylogarithms, Bhabha scattering

BONN-TH-2021-06 , OUTP-21-19P

1 Introduction

The Bhabha scattering process, i.e., the elastic scattering of an electron-positron pair, is one of the standard candles at lepton colliders, and it will play a crucial role at future circular or linear colliders. A precise theoretical knowledge of this process, including up to next-to-next-leading order effects in the QED coupling constant, is therefore highly desirable. So far the NNLO cross section is only known in the massless limit, supplemented by the leading logarithmic finite mass effects (see, e.g., ref. Banerjee:2021mty for recent results). Complete NNLO results including the full dependence on the electron mass, however, are not yet available.

One of the main obstacles to obtain the complete NNLO results is the complexity of the two-loop integrals involved. The relevant two-loop integrals were evaluated in the small-mass limit in refs. Actis:2006dj ; Actis:2007gi ; Actis:2007pn ; Actis:2007fs ; Actis:2008br . Up to now, there are only partial results for two-loop integrals with massive fermions. Analytic results for diagrams with one-loop insertions and a closed massive fermion loop were obtained in refs. Bonciani:2003cj ; Bonciani:2004gi . First analytic results for massive two-loop double-box Bhabha diagrams were obtained in refs. Smirnov:2001cm ; Heinrich:2004iq . More attempts to evaluate two-loop Bhabha integrals can be found in refs. Czakon:2004wm ; Czakon:2005jd ; Czakon:2005gi ; Czakon:2006pa ; Czakon:2006hb .

The master integrals relevant for the calculation of the two-loop corrections to Bhabha scattering can be classified into two planar families and one non-planar one. The systematic evaluation of the integrals in the first family was started in ref. Henn:2013woa , where the master integrals for the family associated with graph (a) of fig. 1 were evaluated in terms of multiple polylogarithms Lappo:1927 ; Goncharov:1998kja ; GoncharovMixedTate (MPLs)

Refer to caption
Figure 1: Planar graphs of the first and the second type for two-loop Bhabha scattering. Solid (dashed) lines indicate massive (massless) propagators. All the external momenta are incoming.

in the framework of the method of differential equations Kotikov:1990kg ; Remiddi:1997ny ; Gehrmann:1999as with the help of the strategy based on canonical bases Henn:2013pwa . In ref. Henn:2013woa , a solution in terms of MPLs had been provided for all integrals except one, whose analytic evaluation was hindered by the presence of the a non-rationalisable square-root in the symbol alphabet. More recently, it was shown that also this last integral can be expressed in terms of multiple polylogarithms by an integration technique based on an ansatz of MPLs with suitable arguments Heller:2019gkq .

The goal of the present paper is to analytically evaluate the master integrals for the second planar family, which is associated with graph (b) of fig. 1. In a way that is reminiscent of the first planar family of integrals, we will show that also in this case we can obtain results in terms of MPLs for all master integrals but one (see fig. 2), due to the presence of the same non-rationalisable square root found in the evaluation of the master integrals for graph (a). While it can be shown by direct integration techniques that also for this master integral a representation in terms of MPLs exist, the representation we obtained is extremely cumbersome and of no practical use. Nevertheless, it turns out that a compact result for this integral can be derived in terms of the elliptic MPLs introduced in refs. Broedel:2017kkb ; Broedel:2017siw ; Broedel:2018iwv . As a byproduct of our analysis, we will also provide a very compact analytic result for the remaining master integral in the first family in fig. 1(a) in terms of the same class of functions.111This result had first been presented by one of the authors at the Loops and Legs Conference 2018 in St. Goar, but it has never been published before.

The rest of the paper is organised as follows. In section 2 we explain the notation, present the system of canonical differential equations for the problem at hand, and introduce the alphabet needed to describe the planar family in fig. 1(b). The alphabet contains multiple square roots, and it is a priori not obvious that the differential equations can be integrated in terms of multiple polylogarithms. In section 3 we review general sufficient criteria and algorithms to solve a system of canonical differential equations in terms of MPLs or their elliptic generalisations. We then continue in section 4 to apply these ideas to our computation. We explain how our differential equations can straightforwardly be solved in terms of multiple polylogarithms for all elements of the basis of the master integrals but one, which is connected with the graph shown in fig. 2.

Refer to caption
Figure 2: The graph associated with the master integral which we evaluate in terms of elliptic polylogarithms.

In section 5 we show how a convenient and compact representation for this integral can be found in terms of elliptic generalisations of MPLs. We also provide a very compact, alternative, analytic solution for one of the master integrals in the first planar family considered in refs. Henn:2013pwa ; Heller:2019gkq . Finally we draw our conclusions in section 6.

2 Canonical differential equations

The Feynman integrals for the family of fig. 1 (b) can be organised in an integral family with nine propagators, where the first seven propagators correspond to the edges of the graph and the last two are so-called irreducible numerators,

Fa1,a2,,a9(s,t,m2;D)=(eγEϵiπD/2)2dDk1dDk2[k12+m2]a1[(k1+p1+p2)2+m2]a2[k22]a3×[(k2+p1)2]a8[(k1p3)2]a9[(k2+p1+p2)2]a4[(k1+p1)2]a5[(k1k2)2+m2]a6[(k2p3)2+m2]a7.\begin{split}&F_{a_{1},a_{2},\ldots,a_{9}}(s,t,m^{2};D)=\left(\frac{e^{\gamma_{E}\epsilon}}{i\pi^{D/2}}\right)^{2}\int\frac{d^{D}k_{1}\,d^{D}k_{2}}{[-k_{1}^{2}+m^{2}]^{a_{1}}[-(k_{1}+p_{1}+p_{2})^{2}+m^{2}]^{a_{2}}[-k_{2}^{2}]^{a_{3}}}\\ &\quad\times\frac{[-(k_{2}+p_{1})^{2}]^{-a_{8}}\,[-(k_{1}-p_{3})^{2}]^{-a_{9}}}{[-(k_{2}+p_{1}+p_{2})^{2}]^{a_{4}}[-(k_{1}+p_{1})^{2}]^{a_{5}}[-(k_{1}-k_{2})^{2}+m^{2}]^{a_{6}}[-(k_{2}-p_{3})^{2}+m^{2}]^{a_{7}}}\,.\end{split} (1)

The indices aia_{i}, i=17i=1\ldots 7 can be positive or negative integers, but we restrict our computation to a8,a90a_{8},a_{9}\leq 0. We work in dimensional regularisation in D=42ϵD=4-2\epsilon dimensions in order to regulate both infrared and ultraviolet divergences. The electron mass is denoted by mm, and the external momenta pip_{i} are on shell, pi2=m2p_{i}^{2}=m^{2}. We introduce the usual Mandelstam variables

s=(p1+p2)2,t=(p1+p3)2,u=(p2+p3)2,s=(p_{1}+p_{2})^{2}\,,\qquad t=(p_{1}+p_{3})^{2}\,,\qquad u=(p_{2}+p_{3})^{2}\,, (2)

with s+t+u=4m2s+t+u=4m^{2}.

Using the public codes FIRE Smirnov:2019qkx and KIRA Maierhoefer:2017hyi ; Klappert:2020nbg , we solve the integration-by-parts (IBP) identities Chetyrkin:1981qh ; Tkachov:1981wb and reveal 43 independent master integrals, which we collect into the vector g=(g1,,g43)Tg=(g_{1},\ldots,g_{43})^{T}. This vector satisfies a system of linear differential equations of the form

vg=Avg,\partial_{v}g=A_{v}g\,, (3)

where v=s,t,m2v=s,t,m^{2}, v=v\partial_{v}=\frac{\partial}{\partial v} and the matrices As,At,Am2A_{s},A_{t},A_{m^{2}} are rational functions of s,t,m2s,t,m^{2} and ϵ\epsilon. In the following, it will be useful to collect all the partial derivatives into a total differential, and to work with the equation

dg=dAg,dA=dsAs+dtAt+dm2Am2.dg=dAg\,,\qquad dA=ds\,A_{s}+dt\,A_{t}+dm^{2}\,A_{m^{2}}\,. (4)

In order to evaluate the master integrals, it is convenient to search for a so-called canonical basis Henn:2013pwa , i.e., a basis of master integrals for which the matrix dAdA on the right-hand side of eq. (4) is proportional to ϵ\epsilon and its entries can all be expressed as linear combinations of total differentials of logarithms. While it has been suggested (at least conjecturally) that the study of the residues of the integrands can provide the full information to determine the elements of the canonical basis ArkaniHamed:2010gh , this analysis can become computational very expensive when square-roots are involved.222See ref. Henn:2020lye for an automated implementation of this approach. Therefore, we follow here a mixed approach to find a canonical basis. In order to select suitable candidates, we start by analysing only the leading singularities associated to the maximal cuts of the various integrals, and we supplement this analysis by the method of ref. Gehrmann:2014bfa . By choosing integrals with unit leading singularities at the level of the maximal cuts, one can often bring the initial differential equations into a so-called precanonical form, where the corresponding matrices depend linearly on ϵ\epsilon. Once this is achieved, the prescriptions of ref. Gehrmann:2014bfa can be successfully applied to arrive at a fully canonical form. In our case, the precanonical basis is composed of the 4343 Feynman integrals present on the right-hand side of the above expression for the ϵ\epsilon-basis. In this way we obtain the new system of differential equations

df=ϵdA¯f,df=\epsilon\,d\bar{A}\,f\,, (5)

with A¯\bar{A} independent of ϵ\epsilon. Our canonical basis fif_{i} is given in appendix A. The price to pay for casting the equations in canonical form is that the new matrix dA¯d\bar{A} is not a matrix of rational one-forms, but it involves four square roots:

rs\displaystyle r_{s} =s4m2s,rt=t4m2t,\displaystyle\,=\sqrt{-s}\sqrt{4m^{2}-s},\qquad r_{t}=\sqrt{-t}\sqrt{4m^{2}-t}, (6)
ru\displaystyle r_{u} =st4m2st,rst=s4m6s(m2t)2.\displaystyle=\sqrt{-s-t}\sqrt{4m^{2}-s-t},\quad r_{st}\,=\sqrt{-s}\sqrt{4m^{6}-s(m^{2}-t)^{2}}\,.

More precisely, the matrix A¯\bar{A} takes the general form

A¯=i=116A¯ilogRi(s,t,m2),\displaystyle\bar{A}=\sum_{i=1}^{16}\bar{A}_{i}\,\log R_{i}(s,t,m^{2})\,, (7)

where the RiR_{i} are algebraic functions, referred to as letters (and their collection is called the alphabet):

R1(s,t,m2)\displaystyle R_{1}(s,t,m^{2}) =sm2,R2(s,t,m2)=s4m2m2,R3(s,t,m2)=srss+rs,\displaystyle\,=\frac{s}{m^{2}}\,,\qquad R_{2}(s,t,m^{2})=\frac{s-4m^{2}}{m^{2}}\,,\qquad R_{3}(s,t,m^{2})=\frac{s-r_{s}}{s+r_{s}}\,,
R4(s,t,m2)\displaystyle R_{4}(s,t,m^{2}) =tm2,R5(s,t,m2)=t4m2m2,R6(s,t,m2)=trtt+rt,\displaystyle\,=\frac{t}{m^{2}}\,,\qquad R_{5}(s,t,m^{2})=\frac{t-4m^{2}}{m^{2}}\,,\qquad R_{6}(s,t,m^{2})=\frac{t-r_{t}}{t+r_{t}}\,,
R7(s,t,m2)\displaystyle R_{7}(s,t,m^{2}) =s+tm2,R8(s,t,m2)=s+t4m2m2,R9(s,t,m2)=s+trus+t+ru,\displaystyle\,=\frac{s+t}{m^{2}}\,,\qquad R_{8}(s,t,m^{2})=\frac{s+t-4m^{2}}{m^{2}}\,,\qquad R_{9}(s,t,m^{2})=\frac{s+t-r_{u}}{s+t+r_{u}}\,,
R10(s,t,m2)\displaystyle R_{10}(s,t,m^{2}) =st+rsrtstrsrt,R11(s,t,m2)=(s+t)rssru(s+t)rs+sru,\displaystyle\,=\frac{st+r_{s}r_{t}}{st-r_{s}r_{t}}\,,\qquad R_{11}(s,t,m^{2})=\frac{(s+t)r_{s}-sr_{u}}{(s+t)r_{s}+sr_{u}}\,, (8)
R12(s,t,m2)\displaystyle R_{12}(s,t,m^{2}) =(s+t)rttru(s+t)rt+tru,R13(s,t,m2)=rstm2s+strst+m2sst,\displaystyle\,=\frac{(s+t)r_{t}-tr_{u}}{(s+t)r_{t}+tr_{u}}\,,\qquad R_{13}(s,t,m^{2})=\frac{r_{st}-m^{2}s+st}{r_{st}+m^{2}s-st}\,,
R14(s,t,m2)\displaystyle R_{14}(s,t,m^{2}) =rtrst3m2st+st2rtrst+3m2stst2,R15(s,t,m2)=rsrst+4m4sm2s2+s2trsrst4m4s+m2s2s2t,\displaystyle\,=\frac{r_{t}r_{st}-3m^{2}st+st^{2}}{r_{t}r_{st}+3m^{2}st-st^{2}}\,,\qquad R_{15}(s,t,m^{2})=\frac{r_{s}r_{st}+4m^{4}s-m^{2}s^{2}+s^{2}t}{r_{s}r_{st}-4m^{4}s+m^{2}s^{2}-s^{2}t}\,,
R16(s,t,m2)\displaystyle R_{16}(s,t,m^{2}) =m2rsrstsrsrst4m6s3m4s2+m2s3+3m2s2ts3tm2rsrstsrsrst+4m6s+3m4s2m2s33m2s2t+s3t.\displaystyle\,=\frac{m^{2}r_{s}r_{st}-sr_{s}r_{st}-4m^{6}s-3m^{4}s^{2}+m^{2}s^{3}+3m^{2}s^{2}t-s^{3}t}{m^{2}r_{s}r_{st}-sr_{s}r_{st}+4m^{6}s+3m^{4}s^{2}-m^{2}s^{3}-3m^{2}s^{2}t+s^{3}t}\,.

As we will see, the appearance of these square roots makes the solution of the differential equation (5) in terms of MPLs highly non-trivial.

It is easy to write down a formal solution to eq. (5) as

f(s,t,m2;ϵ)=exp[ϵγ𝑑A~]f0(ϵ),f(s,t,m^{2};\epsilon)=\mathbb{P}\textrm{exp}\left[\epsilon\int_{\gamma}d\tilde{A}\right]\,f_{0}(\epsilon)\,, (9)

where exp\mathbb{P}\textrm{exp} denotes the path-ordered exponential and f0(ϵ)f_{0}(\epsilon) encodes the initial condition and is related to the value of ff at a specific point (s0,t0,m02)(s_{0},t_{0},m_{0}^{2}). The path γ\gamma connects the initial point (s0,t0,m02)(s_{0},t_{0},m_{0}^{2}) to the generic point (s,t,m2)(s,t,m^{2}).

When expanding eq. (9) in ϵ\epsilon, then at each order we can write ff in terms of Chen iterated integrals ChenSymbol , defined in the following way: consider a path γ\gamma and a collection of one-forms ωi\omega_{i}. If tt denotes a local coordinate on γ\gamma, we can pull each ωi\omega_{i} back to γ\gamma and write γωi=dtFi(t)\gamma^{*}\omega_{i}=dt\,F_{i}(t). We can then define the iterated integral of a sequence ωi1ωin\omega_{i_{1}}\ldots\omega_{i_{n}} (usually referred to as a word) by

γωi1ωin=0t1tn1𝑑tnFin(tn)0tn0t2𝑑t1Fi1(t1).\int_{\gamma}\omega_{i_{1}}\ldots\omega_{i_{n}}=\int_{0\leq t_{1}\leq\ldots\leq t_{n}\leq 1}dt_{n}\,F_{i_{n}}(t_{n})\int_{0}^{t_{n}}\ldots\int_{0}^{t_{2}}dt_{1}\,F_{i_{1}}(t_{1})\,. (10)

The number nn of integrations is called the length of the iterated integral. In general, this integral will not be homotopy-invariant, i.e., it will depend on the details of the path γ\gamma. There is a necessary and sufficient condition, called the integrability condition, for a combination of iterated integrals to be homotopy-invariant ChenSymbol . The details of this criterion are not important in the following. Here it suffices to say that it is always satisfied for the solutions in eq. (9). The corresponding Chen iterated integrals are then multi-valued functions of the end point (s,t,m2)(s,t,m^{2}) of the path γ\gamma, where the multi-valuedness only comes from choosing two non-homotopic paths from (s0,t0,m02)(s_{0},t_{0},m_{0}^{2}) to (s,t,m2)(s,t,m^{2}).

The master integrals f(s,t,m2;ϵ)f(s,t,m^{2};\epsilon) can be expressed at every order in terms of Chen iterated integrals over words from the alphabet in eq. (8). The coefficient of ϵn\epsilon^{n} in the path-ordered exponential in eq. (9) only involves iterated integrals of length nn. Chen iterated integrals are a very general class of functions, and it can be useful to express them in terms of a class of functions that are well-studied in the literature. In the next section we discuss sufficient criteria for when Chen iterated integrals over dlogd\log-forms with algebraic arguments, as the ones above, can be expressed in terms of other classes of functions.

3 From dlogd\log-forms to multiple polylogarithms

Since the solution of differential equations in canonical form as Laurent series in dimensional regularisation leads naturally to linear combinations of Chen iterated integrals over words of dlogd\log-forms, it is natural to ask when it possible to evaluate such iterated integrals in terms of other classes of functions, and if it is possible to perform this rewriting algorithmically. The advantage of this rewriting lies in the fact that these classes of functions may be well studied in the literature, and there may be established methods or computer codes for their manipulation and/or their numerical evaluation.

Let us consider Chen iterated integrals of the form γw\int_{\gamma}w, where γ\gamma is a path from an initial point 𝐱0=(x0,1,,x0,p){\bf x}_{0}=(x_{0,1},\ldots,x_{0,p}) to the point 𝐱1=(x1,1,,x1,p){\bf x}_{1}=(x_{1,1},\ldots,x_{1,p}), and ww is an integrable combination of words of length nn of the form

w=i1inci1indlogRi1(𝐱)dlogRin(𝐱),ci1in.w=\sum_{{i_{1}\ldots i_{n}}}c_{i_{1}\ldots i_{n}}\,d\log R_{i_{1}}({\bf x})\ldots d\log R_{i_{n}}({\bf x})\,,\qquad c_{i_{1}\ldots i_{n}}\in\mathbb{Q}\,. (11)

We consider the initial point 𝐱0{\bf x}_{0} fixed, and we see the integral as a multi-valued function of the end-point 𝐱1{\bf x}_{1}. The arguments of the logarithms are assumed to be algebraic functions of the variables 𝐱=(x1,,xp){\bf x}=(x_{1},\ldots,x_{p}).

For some time, there was a folklore belief in the physics community that all such iterated integrals could be evaluated in terms of a rather simple class of iterated integrals, namely the so-called multiple polylogarithms (MPLs), defined by Lappo:1927 ; Goncharov:1998kja ; Remiddi:1999ew ; GoncharovMixedTate

Ga1,,an(z)=G(a1,,an;z)=0zdtta1G(a2,,an;t),G_{a_{1},\ldots,a_{n}}(z)=G(a_{1},\ldots,a_{n};z)=\,\int_{0}^{z}\,\frac{dt}{t-a_{1}}\,G(a_{2},\ldots,a_{n};t)\,, (12)

where the recursion starts with G(;z)1G(;z)\equiv 1. In the special case where all the aia_{i}’s are zero, we define

G0,,0(z)=G(0,,0n;z)=1n!lognz.G_{0,\ldots,0}(z)=G(\underbrace{0,\ldots,0}_{n};z)=\frac{1}{n!}\,\log^{n}z\,. (13)

The shared belief in physics was that, whenever the dlogd\log’s in eq. (11) have algebraic arguments RiR_{i}, then this integral can be written in terms of MPLs whose arguments aia_{i} and zz are algebraic functions of 𝐱0{\bf x}_{0} and 𝐱1{\bf x}_{1}. This belief was shown to be false in ref. Brown:2020rda , where an example of an iterated integral of length two was constructed that cannot be evaluated in terms of MPLs with algebraic arguments. The result of ref. Brown:2020rda shows that the question of whether an iterated integral of dlogd\log-forms with algebraic arguments can be expressed in terms of MPLs depends in general on the details of the integral, including the details of the integration contour.

In the remainder of this section we discuss some special cases of algebraic letters RiR_{i} where the Chen iterated integrals can be evaluated in terms of other classes of special functions. The starting point is the observation that in the context of Feynman integrals, the RiR_{i} are not generic algebraic functions, but often involve at most square roots of polynomials, i.e., they are of the form:

Ri(𝐱)=ri,0(𝐱)+a=1Nsqrtri,a(𝐱)qa(𝐱),R_{i}({\bf x})=r_{i,0}({\bf x})+\sum_{a=1}^{N_{\textrm{sqrt}}}r_{i,a}({\bf x})\,\sqrt{q_{a}({\bf x})}\,, (14)

where the ri,jr_{i,j} are rational functions, and the qaq_{a} are polynomials. Without loss of generality, we assume the ri,jr_{i,j} to be polynomials. At this point we have to make a comment: The integrand of the iterated integration, and thus the functional form of the letters RiR_{i}, is sensitive to a change of variables. In particular, one may ask if we can find a change of variables 𝐱=ψ(𝐲){\bf x}=\psi({\bf y}), with ψ\psi a rational function, such that qa(ψ(𝐲))=ua(𝐲)2q_{a}(\psi({\bf y}))=u_{a}({\bf y})^{2} is a perfect square, for some values of aa. This operation is known as rationalisation of square roots. Over the last few years, several algorithmic criteria have been developed to find such a parametrisation for specific Feynman integral computations, or to prove that rationalisation is instead not possible Festi1 ; Festi:2018qip ; Besier:2018jen ; Besier:2019kco ; Besier:2020klg ; Besier:2020hjf ; Festi:2021tyq . Note that rationalisability can easily be decided for a single square root of a one-variable polynomial (p=Nsqrt=1p=N_{\textrm{sqrt}}=1) based on the degree degq1\deg q_{1} of the polynomial: the square root q1(x1)\sqrt{q_{1}(x_{1})} can be rationalised (and the corresponding function ψ\psi can be constructed explicitly) if and only if degq12\deg q_{1}\leq 2. The question of the rationalisability for p>1p>1 is much more involved, even in the presence of a single square root, see refs. Besier:2018jen ; Besier:2019kco ; Besier:2020klg ; Besier:2020hjf ; Festi:2021tyq .

In the following we discuss two special cases of eq. (14), in which we can express the Chen iterated integrals in terms of other classes of iterated integrals:

  • If Nsqrt=0N_{\textrm{sqrt}}=0, the Chen iterated integrals can be expressed in terms of MPLs evaluated at algebraic arguments.

  • If Nsqrt=1N_{\textrm{sqrt}}=1 and degq1=3\deg q_{1}=3 or 44, the Chen iterated integrals can be expressed in terms of elliptic generalisations of MPLs evaluated at algebraic arguments.

In particular, we explain how we can algorithmically rewrite all Chen iterated integrals that meet these criteria in terms of MPLs and their elliptic analogues. This algorithm is well known, but we document it here because it will be an important tool to obtain compact analytic expressions for some master integrals that contribute to Bhabha scattering at two loops. Two comments are in order: First, in the presence of (multiple) square roots, it may be possible to change variables and rationalise (some of) the roots. In this way it may possible to reduce the problem to a situation covered by the criteria above, even though the original problem did not satisfy these conditions. Second, we stress that the aforementioned conditions are only sufficient, but by no means necessary, to rewrite Chen iterated integrals in terms of (elliptic) MPLs. In particular, there are several examples of Feynman integrals whose alphabets involve non-rationalisable square roots, but nevertheless, it was possible to express them in terms of ordinary MPLs, cf.,  e.g., refs. Heller:2019gkq ; Heller:2021gun ; Kreer:2021sdt .

3.1 Rational alphabets without square roots

If the alphabet does not contain any square roots, or if all square roots can be rationalised, we can always express the Chen iterated integral in terms of MPLs evaluated at algebraic arguments. First, we can use the additivity of the logarithm to assume that all letters RiR_{i} are irreducible polynomials. Next, we can use the homotopy-invariance of the integral to deform the contour γ\gamma into a new contour γ0\gamma_{0} with the same end-points 𝐱0{\bf x}_{0} and 𝐱1{\bf x}_{1}, without changing the value of the integral (except for picking up residues when we cross a pole). We choose the new contour as follows: We order the integration variables in some way, which for simplicity we assume to be the natural order x1,,xpx_{1},\ldots,x_{p}. The contour γ0\gamma_{0} is then obtained as the concatenation of the straight-line segments γr\gamma_{r}, 1rp1\leq r\leq p, defined by 𝐱=φ(r)(t)=(φ1(r)(t),φp(r)(t)){\bf x}=\varphi^{(r)}(t)=(\varphi^{(r)}_{1}(t)\,\ldots,\varphi^{(r)}_{p}(t)) with 0t10\leq t\leq 1 and

φi(r)(t)={x1,i,i<r,(x1,rx0,r)t+x0,r,i=r,x0,i,i>r.\varphi^{(r)}_{i}(t)=\left\{\begin{array}[]{ll}x_{1,i}\,,&\quad i<r\,,\\ \left(x_{1,r}-x_{0,r}\right)t+x_{0,r}\,,&\quad i=r\,,\\ x_{0,i}\,,&\quad i>r\,.\end{array}\right. (15)

Next we can iteratively apply the path-composition formula for iterated integrals,

γ1γ2ωi1ωin=k=0nγ1ωi1ωikγ2ωik+1ωin,ωi=dlogRi(𝐱).\int_{\gamma_{1}\gamma_{2}}\omega_{i_{1}}\ldots\omega_{i_{n}}=\sum_{k=0}^{n}\int_{\gamma_{1}}\omega_{i_{1}}\ldots\omega_{i_{k}}\int_{\gamma_{2}}\omega_{i_{k+1}}\ldots\omega_{i_{n}}\,,\qquad\omega_{i}=d\log R_{i}({\bf x})\,. (16)

The previous equation allows us to reduce the integral to a linear combination of products of Chen iterated integrals over the straight-line segments γr\gamma_{r}. On the line segment γr\gamma_{r} the dlogd\log-forms take a particularly simple form. Indeed, since all the letters RiR_{i} are polynomials in 𝐱{\bf x}, it is easy to see that Ri(φr(t))R_{i}(\varphi_{r}(t)) is a polynomial in tt. In the following we write

Ri(φr(t))=c0(tc1)(tcd),R_{i}(\varphi_{r}(t))=c_{0}\,(t-c_{1})\ldots(t-c_{d})\,, (17)

where dd denotes the degree of the polynomial Ri(φr(t))R_{i}(\varphi_{r}(t)) and the quantities cjc_{j}, 0jd0\leq j\leq d are algebraic functions of 𝐱0{\bf x}_{0} and 𝐱1{\bf x}_{1}. The pull-back of the one-form dlogRi(𝐱)d\log R_{i}({\bf x}) to the straight-line segment γr\gamma_{r} then reads

γrdlogRi(𝐱)=dtFr,i(t),\gamma^{\ast}_{r}d\log R_{i}({\bf x})=dt\,F_{r,i}(t)\,, (18)

with

Fr,i(t)={0, if xrRi(𝐱)=0,j=1d1tcj, if xrRi(𝐱)0.F_{r,i}(t)=\left\{\begin{array}[]{ll}0\,,&\textrm{ if }\partial_{x_{r}}R_{i}({\bf x})=0\,,\\ \sum_{j=1}^{d}\frac{1}{t-c_{j}}\,,&\textrm{ if }\partial_{x_{r}}R_{i}({\bf x})\neq 0\,.\end{array}\right. (19)

The previous equation makes it manifest that on each straight-line segment the Chen iterated integral evaluates to ordinary MPLs with algebraic arguments cjc_{j}, cf. eq. (12).

3.2 Alphabets with a single elliptic square root

We have already seen that a square root of a polynomial of degree three or four cannot be rationalised. More precisely, consider the set of points (x,y)(x,y) constrained by

y2=Pn(x),y^{2}=P_{n}(x)\,, (20)

where Pn(x)P_{n}(x) is a polynomial of degree nn in xx. If n=3n=3 or 44, this equation defines an algebraic variety called an elliptic curve. It is thus not surprising that the case Nsqrt=1N_{\textrm{sqrt}}=1 and q1=Pnq_{1}=P_{n}, with n=3n=3 or 44, leads to generalisations of MPLs related to elliptic curves. We start by giving a lightning review of elliptic multiple polylogarithms (eMPLs), before we comment on the generalisation of the algorithm from section 3.1. More details can be found in appendix B.

We focus here on the case n=4n=4, and we assume that P4P_{4} has the form P4(x)=(xa1)(xa2)(xa3)(xa4)P_{4}(x)=(x-a_{1})(x-a_{2})(x-a_{3})(x-a_{4}). Elliptic multiple polylogarithms (eMPLs) can then be defined as the iterated integrals BrownLevin ; Broedel:2014vla ; Broedel:2017kkb ; Broedel:2018qkq

4(n1nkc1ck;x,a)=0x𝑑tΨn1(c1,t,a)4(n2nkc2ck;t,a),{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}n_{1}&\ldots&n_{k}\\ c_{1}&\ldots&c_{k}\end{smallmatrix};x,\vec{a}\right)=\int_{0}^{x}dt\,\Psi_{n_{1}}(c_{1},t,\vec{a})\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}n_{2}&\ldots&n_{k}\\ c_{2}&\ldots&c_{k}\end{smallmatrix};t,\vec{a}\right)\,, (21)

with nin_{i}\in\mathbb{Z} and ci{}c_{i}\in{\mathbb{C}}\cup\{\infty\}, and a=(a1,a2,a3,a4)\vec{a}=(a_{1},a_{2},a_{3},a_{4}) is the vector of the four branch points aia_{i}. Just like ordinary MPLs, eMPLs have at most logarithmic singularities, but no poles. The number n=i=1k|ni|n=\sum_{i=1}^{k}|n_{i}| is called the weight of the eMPL, and the number of integrations kk is its length. In the case where all the indices Ai=(nici)A_{i}=\left(\begin{smallmatrix}n_{i}\\ c_{i}\end{smallmatrix}\right) are equal to (±10)\left(\begin{smallmatrix}\pm 1\\ 0\end{smallmatrix}\right), the integral is divergent and requires a special treatment similar to the case an=0a_{n}=0 for ordinary MPLs, cf. eq. (13). We refer to appendix B for details. Also note that we need to be careful about how we choose the branches of the square root y=P4(x)y=\sqrt{P_{4}(x)}. We will come back to this point in section 5.

There are infinitely many integration kernels Ψn(c,x,a)\Psi_{n}(c,x,\vec{a}) for given (c,x,a)(c,x,\vec{a}) in eq. (21). In concrete applications only a finite number of these kernels appear. Here we only list the kernels with |n|1|n|\leq 1, which are of direct relevance to this paper. For n=0n=0, we have

Ψ0(0,x,a)=c4ω1y,\Psi_{0}(0,x,\vec{a})=\frac{c_{4}}{\omega_{1}\,y}\,, (22)

with c4=12a13a24c_{4}=\frac{1}{2}\sqrt{a_{13}a_{24}}, aij=aiaja_{ij}=a_{i}-a_{j}. ω1\omega_{1} is one of the two periods associated to the elliptic curve defined by the equation y2=P4(x)y^{2}=P_{4}(x),

ω1=2K(λ),ω2=2iK(1λ),λ=a14a23a13a24,\omega_{1}=2\,\textrm{K}(\lambda)\,,\qquad\omega_{2}=2i\,\textrm{K}(1-\lambda)\,,\qquad\lambda=\frac{a_{14}\,a_{23}}{a_{13}\,a_{24}}\,, (23)

where K is the complete elliptic integral of the first kind

K(λ)=01dt(1t2)(1λt2).\textrm{K}(\lambda)=\int_{0}^{1}\frac{dt}{\sqrt{(1-t^{2})(1-\lambda t^{2})}}\,. (24)

For |n|=1|n|=1, we have (with cc\neq\infty)

Ψ1(c,x,a)\displaystyle\Psi_{1}(c,x,\vec{a}) =1xc,\displaystyle\,=\frac{1}{x-c}\,,
Ψ1(c,x,a)\displaystyle\Psi_{-1}(c,x,\vec{a}) =ycy(xc)+Z4(c,a)c4y,\displaystyle\,=\frac{y_{c}}{y(x-c)}+Z_{4}(c,\vec{a})\,\frac{c_{4}}{y}\,, (25)
Ψ1(,x,a)\displaystyle\Psi_{1}(\infty,x,\vec{a}) =Z4(x,a)c4y,\displaystyle\,=-Z_{4}(x,\vec{a})\,\frac{c_{4}}{y}\,,
Ψ1(,x,a)\displaystyle\Psi_{-1}(\infty,x,\vec{a}) =xy1y[a1+2c4G(a)],\displaystyle\,=\frac{x}{y}-\frac{1}{y}\left[{a_{1}}+2c_{4}\,G_{\ast}(\vec{a})\right]\,,

where we introduced the shorthand yc=P4(c)y_{c}=\sqrt{P_{4}(c)}. These functions have the property that the differential form dxΨ±1(c,x,a)dx\,\Psi_{\pm 1}(c,x,\vec{a}) has a simple pole at x=cx=c, and no other poles (except for dxΨ1(c,x,a)dx\,\Psi_{1}(c,x,\vec{a}), which always has a pole at x=x=\infty). Note that the kernel Ψ1\Psi_{1} is identical to the kernel that defines ordinary MPLs (cf. eq. (12)), and so ordinary MPLs are a subset of eMPLs,

4(11c1ck;x,a)=G(c1,,ck;x).{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}1&\ldots&1\\ c_{1}&\ldots&c_{k}\end{smallmatrix};x,\vec{a}\right)=G(c_{1},\ldots,c_{k};x)\,. (26)

The functions Z4(c,a)Z_{4}(c,\vec{a}) and G(a)G_{\ast}(\vec{a}) in eq. (25) are in general transcendental functions. They can be expressed in terms of incomplete elliptic integrals of the first and second kind (see appendix B). Depending on the value of the argument cc, in many applications it is possible to evaluate Z4(c,a)Z_{4}(c,\vec{a}) and G(a)G_{\ast}(\vec{a}) in the form A+Biπω1A+B\frac{i\pi}{\omega_{1}}, where AA and BB are algebraic functions of a\vec{a} and cc Broedel:2018qkq . In particular, in all cases relevant for this paper, Z4(c,a)Z_{4}(c,\vec{a}) and G(a)G_{\ast}(\vec{a}) will always be algebraic functions (see section 5.1).

Let us now assume that the alphabet is given by dlogd\log-forms with arguments

Ri(𝐱)=ri,0(𝐱)+ri,1(𝐱)q1(𝐱),R_{i}({\bf x})=r_{i,0}({\bf x})+r_{i,1}({\bf x})\,\sqrt{q_{1}({\bf x})}\,, (27)

where ri,0(𝐱)r_{i,0}({\bf x}) are polynomials in 𝐱{\bf x} and q1(𝐱)q_{1}({\bf x}) is a non-constant polynomial of degree at most four. We assume that the ri,j(𝐱)r_{i,j}({\bf x}) do not have any common zero (otherwise we could factor out a term from this letter in the alphabet), and q1(𝐱)q_{1}({\bf x}) does not have a double zero (otherwise we could factor it out of the square root). We allow ri,1(𝐱)r_{i,1}({\bf x}) to be zero (in which case this letter does not depend on the square root), but ri,0(𝐱)r_{i,0}({\bf x}) is assumed not to vanish. We can separate all the letters into even and odd parts:

dlogRi(𝐱)=12dlogRi+(𝐱)12dlogRi(𝐱),d\log R_{i}({\bf x})=\frac{1}{2}d\log R_{i}^{+}({\bf x})-\frac{1}{2}d\log R_{i}^{-}({\bf x})\,, (28)

with

Ri+(𝐱)=ri,0(𝐱)2ri,1(𝐱)2q1(𝐱),Ri(𝐱)=ri,0(𝐱)ri,1(𝐱)q1(𝐱)ri,0(𝐱)+ri,1(𝐱)q1(𝐱).\begin{split}R_{i}^{+}({\bf x})&\,=r_{i,0}({\bf x})^{2}-r_{i,1}({\bf x})^{2}{q_{1}({\bf x})}\,,\\ R_{i}^{-}({\bf x})&\,=\frac{r_{i,0}({\bf x})-r_{i,1}({\bf x})\,\sqrt{q_{1}({\bf x})}}{r_{i,0}({\bf x})+r_{i,1}({\bf x})\,\sqrt{q_{1}({\bf x})}}\,.\end{split} (29)

Next we deform the integration path γ\gamma to the sequence of line segments γr\gamma_{r} defined in eq. (15). Even letters give rise to integration kernels of the form dttcj\frac{dt}{t-c_{j}} and can be dealt with exactly as in section 3.1. We will not discuss them any further. For the odd letters, note that on the straight-line segment γr\gamma_{r}, q1(φr(t))q_{1}(\varphi_{r}(t)) is a polynomial of degree nn in tt, with n4n\leq 4. We write:

q1(φr(t))=αrP(r)(t)=αr(tar,1)(tar,n),ar=(ar,1,,ar,n).q_{1}(\varphi_{r}(t))=\alpha_{r}\,P^{(r)}(t)=\alpha_{r}\,(t-a_{r,1})\cdots(t-a_{r,n})\,,\qquad\vec{a}_{r}=(a_{r,1},\ldots,a_{r,n})\,. (30)

If n2n\leq 2, then the square root P(r)(t)\sqrt{P^{(r)}(t)} can be rationalised, and all letters are rational on the line segment γr\gamma_{r}, and we do not need to discuss this case anymore. If the degree of P(r)(t)P^{(r)}(t) is three or four, then we cannot rationalise the square root on γr\gamma_{r}. Instead, we obtain

γrdlogRi(φr(t))=dtFr,i(t),\gamma_{r}^{*}d\log R_{i}^{-}(\varphi_{r}(t))=dt\,F^{-}_{r,i}(t)\,, (31)

where the parity of RiR_{i}^{-} implies that Fr,i(t)F^{-}_{r,i}(t) is of the form

Fr,i(t)=1P(r)(t)Sr,i(t),F^{-}_{r,i}(t)=\frac{1}{\sqrt{P^{(r)}(t)}}\,S_{r,i}(t)\,, (32)

where Sr,i(t)S_{r,i}(t) is a rational function in tt. At this point we make an important observation: By construction, the differential form dlogRi(φr(t))d\log R_{i}^{-}(\varphi_{r}(t)) has only logarithmic singularities. This implies that Sr,i(t)S_{r,i}(t) has poles of order at most one (possibly including a simple pole at infinity). As a consequence, Sr,i(t)S_{r,i}(t) may be written as a linear combination of the functions Ψ1(cj,t,ar)\Psi_{-1}(c_{j},t,\vec{a}_{r}) and Ψ0(cj,t,ar)\Psi_{0}(c_{j},t,\vec{a}_{r}). Hence, we can evaluate the integral on γr\gamma_{r} in terms of eMPLs for the elliptic curve associated to P(r)(t)P^{(r)}(t). Note that, a priori, we may obtain eMPLs with different values of ar\vec{a}_{r} for each segment γr\gamma_{r}.

4 Integration of the differential equations in terms of MPLs

After the general remarks of the previous section, we go back to the explicit solution for the system in eq. (5). The standard way to rationalize the first two square roots in eq. (6) is to turn to the dimensionless Landau variables xx and yy related to the Mandelstam invariants by

sm2=(1x)2x and tm2=(1y)2y.\frac{-s}{m^{2}}=\frac{(1-x)^{2}}{x}\textrm{~{}~{}and~{}~{}}\frac{-t}{m^{2}}=\frac{(1-y)^{2}}{y}\,. (33)

In terms of these variables the Euclidean region s,t<0s,t<0 corresponds to 0x,y10\leq x,y\leq 1 (using the symmetry of eq. (33) under (x,y)(1/x,1/y)(x,y)\to(1/x,1/y)).

There exist different strategies to solve the differential equation in eq. (5). The first method consists in evaluating numerically the ϵ\epsilon-expansion of the path-ordered exponential in eq. (9). This can be achieved by application of the Frobenius method to look for power-series solutions of ordinary differential equations in the vicinity of regular singular points. The Frobenius method has been successfully used in various one-dimensional problems in the past Pozzorini:2005ff ; Aglietti:2007as ; Caffo:2008aw ; Bonciani:2018uvv , and more recently a generalisation of this method has been proposed in ref. Francesco:2019yqt to deal with complicated multidimensional problems, see for example Bonciani:2019jyb ; Abreu:2020jxa . This strategy has also been implemented into the public code code DiffExp Hidding:2020ytt .333A Mathematica notebook is provided as ancillary material with the arXiv submission. The input data for this code are the matrices that define the differential equations, and the boundary conditions in some limit, e.g., at some point (x0,y0)(x_{0},y_{0}). The code then uses this input to evolve the solution numerically from the point (x0,y0)(x_{0},y_{0}) to some generic point (x,y)(x,y). We fix the boundary conditions in the limit s,t0s,t\to 0, which corresponds to (x0,y0)=(1,1)(x_{0},y_{0})=(1,1) in terms of the Landau variables in eq. (33). Using the expansion by regions strategy implemented in the public code asy.m Pak:2010pt ; Jantzen:2012mw (which is now included in the code FIESTA Smirnov:2015mct ) we obtain the following leading order asymptotic behaviour in this limit

f1\displaystyle f_{1} 1+π2ϵ262ζ3ϵ33+7π4ϵ4360+𝒪(ϵ5),\displaystyle\,\sim 1+\frac{\pi^{2}\epsilon^{2}}{6}-\frac{2\zeta_{3}\epsilon^{3}}{3}+\frac{7\pi^{4}\epsilon^{4}}{360}+\cal O(\epsilon^{5})\,,
f6\displaystyle f_{6} 145π2ϵ22411ζ3ϵ36101480π4ϵ4+𝒪(ϵ5),\displaystyle\,\sim-\frac{1}{4}-\frac{5\pi^{2}\epsilon^{2}}{24}-\frac{11\zeta_{3}\epsilon^{3}}{6}-\frac{101}{480}\pi^{4}\epsilon^{4}+\cal O(\epsilon^{5})\,,
f9\displaystyle f_{9} π2ϵ212+14ϵ3(2π2log27ζ3)\displaystyle\,\sim-\frac{\pi^{2}\epsilon^{2}}{12}+\frac{1}{4}\epsilon^{3}\left(2\pi^{2}\log 2-7\zeta_{3}\right) (34)
+1180ϵ4(13π490log42180π2log222160Li4(12))+𝒪(ϵ5),\displaystyle\,+\frac{1}{180}\epsilon^{4}\left(13\pi^{4}-90\log^{4}2-180\pi^{2}\log^{2}2-2160\text{Li}_{4}\left(\frac{1}{2}\right)\right)+\cal O(\epsilon^{5})\,,
f18\displaystyle f_{18} 12ϵ3(2π2log23ζ3)+120ϵ4(7π420log4240π2log22480Li4(12))+𝒪(ϵ5),\displaystyle\,\sim\frac{1}{2}\epsilon^{3}\left(2\pi^{2}\log 2-3\zeta_{3}\right)+\frac{1}{20}\epsilon^{4}\left(7\pi^{4}-20\log^{4}2-40\pi^{2}\log^{2}2-480\text{Li}_{4}\left(\frac{1}{2}\right)\right)+\cal O(\epsilon^{5})\,,
f19\displaystyle f_{19} (s)ϵ(1+8ζ3ϵ33+π4ϵ430)+𝒪(ϵ5),\displaystyle\,\sim(-s)^{-\epsilon}\left(-1+\frac{8\zeta_{3}\epsilon^{3}}{3}+\frac{\pi^{4}\epsilon^{4}}{30}\right)+\cal O(\epsilon^{5})\,,
f22\displaystyle f_{22} (s)ϵ(12+4ζ3ϵ33+π4ϵ460)+(s)2ϵ(14π2ϵ22414ζ3ϵ3367480π4ϵ4)+𝒪(ϵ5),\displaystyle\,\sim(-s)^{-\epsilon}\left(-\frac{1}{2}+\frac{4\zeta_{3}\epsilon^{3}}{3}+\frac{\pi^{4}\epsilon^{4}}{60}\right)+(-s)^{-2\epsilon}\left(\frac{1}{4}-\frac{\pi^{2}\epsilon^{2}}{24}-\frac{14\zeta_{3}\epsilon^{3}}{3}-\frac{67}{480}\pi^{4}\epsilon^{4}\right)+\cal O(\epsilon^{5})\,,
f23\displaystyle f_{23} (s)2ϵ(π2ϵ2+2π2ϵ3log2+2π2ϵ4(π2+log22))+𝒪(ϵ5),\displaystyle\,\sim(-s)^{-2\epsilon}\left(\pi^{2}\epsilon^{2}+2\pi^{2}\epsilon^{3}\log 2+2\pi^{2}\epsilon^{4}\left(\pi^{2}+\log^{2}2\right)\right)+\cal O(\epsilon^{5})\,,
f25\displaystyle f_{25} (s)ϵ(π2ϵ22π2ϵ3log212π2ϵ4(π2+4log22))+𝒪(ϵ5),\displaystyle\,\sim(-s)^{-\epsilon}\left(-\pi^{2}\epsilon^{2}-2\pi^{2}\epsilon^{3}\log 2-\frac{1}{2}\pi^{2}\epsilon^{4}\left(\pi^{2}+4\log^{2}2\right)\right)+\cal O(\epsilon^{5})\,,

and fi0f_{i}\sim 0, i.e., fi=𝒪(s,t)f_{i}=\cal O(s,t), for all the other elements.

Let us emphasise that, in order to profit maximally from the automated code DiffExp, it is crucial that the input data are in an optimal form. This includes providing differential equations in canonical form and fixing the boundary conditions in a simple point (s=t=0s=t=0, i.e., x=y=1x=y=1). With our input, the code works very well and provides the possibility to obtain high-precision numerical results (100 digits accuracy and more), both in the Euclidean and the physical regions. In other words, DiffExp allows us to obtain high-precision numerical results for all master integrals and for all values of the input parameters. The code runs fast enough to allow its usage for practical applications. This is then of course sufficient for all phenomenological applications one has in mind. However, both from a formal and from a practical point of view, it may still be desirable in some situations to have full-fledged analytic representations of the master integrals in terms of functions that are well studied in the literature, e.g., MPLs. In the remainder of this section we explain how we can obtain analytic results for all master integrals but f14f_{14} in terms of MPLs evaluated at algebraic arguments. The reason why f14f_{14} is different will be discussed in section 5. Here it suffices to say that the square root rur_{u} only enters the differential equation for f14f_{14}. Since our strategy of solving the differential equations in terms of MPLs will be closely based on the ideas from section 3.1, it is not suprising that an additional square root implies that the sufficient condition of section 3.1 is not satisfied. We will discuss the case of f14f_{14} in detail in section 5, and we focus for the rest of this section only on the other master integrals.

In order to solve the master integrals in terms of MPLs, we start by observing that the square root rstr_{st} does not appear when solving the differential equations up to weight 3 for all elements but f37f_{37}, and at weight 4 for all elements but fi,i{35,36,37,38,39,41,43}f_{i},i\in\{35,36,37,38,39,41,43\}. The equations can be solved by integrating first in xx, resulting in MPLs of the form G(c;x)G(\vec{c};x) with ci{0,1,1,y,1/y}c_{i}\in\{0,-1,1,-y,-1/y\}. This allows us to fix the solution up to a function of yy, which can be determined by substituting the results of the xx-integration into the differential equations in yy, checking that the variable xx disappears in them, and finally solving them in terms of MPLs of the form G(c;y)G(\vec{c};y) with ci{0,1,1}c_{i}\in\{0,-1,1\}, i.e., harmonic polylogarithms Remiddi:1999ew . At this point the solution is fixed up to a set of undetermined integration constants, which we fix using the boundary conditions in eq. (34).

To evaluate f37f_{37} at weights 3 and 4 and fif_{i} with i{35,36,38,39,41,43}i\in\{35,36,38,39,41,43\} at weight 4, we have to deal with the square root rstr_{st}. It can be rationalized by the following further change of variables

x=2(1w)(y2y+1)22y2(1w2)(y2y+1)2.x=2\,\frac{(1-w)\left(y^{2}-y+1\right)^{2}-2y^{2}}{\left(1-w^{2}\right)\left(y^{2}-y+1\right)^{2}}\,. (35)

The resulting system of differential equations for 42 elements can be solved, first in ww and then in yy. Solving in ww gives a linear combination of MPLs G(;w)G(\ldots;w) with the letters the alphabet {l1w,,l15w}\{l^{w}_{1},\ldots,l^{w}_{15}\}, where liw,i=1,,11l^{w}_{i},i=1,\ldots,11, are taken from the set

{1,1,y42y3+y22y+1(y2y+1)2,y2+y+1y2y+1,y23y+1y2y+1,y2+2y2+1y2y2+1y+1y2y+1,y22y2+1y+2y2+1y+1y2y+1,y2y1y2y+1,y2+y1y2y+1,2y3y2+y1y2y+1,y3y2+y2y(y2y+1)},\begin{split}&\,\left\{1,-1,\frac{y^{4}-2y^{3}+y^{2}-2y+1}{\left(y^{2}-y+1\right)^{2}},\frac{y^{2}+y+1}{y^{2}-y+1},\frac{y^{2}-3y+1}{y^{2}-y+1},\right.\\ &\,-\frac{y^{2}+2\sqrt{y^{2}+1}y-2\sqrt{y^{2}+1}-y+1}{y^{2}-y+1},-\frac{y^{2}-2\sqrt{y^{2}+1}y+2\sqrt{y^{2}+1}-y+1}{y^{2}-y+1},\\ &\,\left.\frac{y^{2}-y-1}{y^{2}-y+1},-\frac{y^{2}+y-1}{y^{2}-y+1},-\frac{2y^{3}-y^{2}+y-1}{y^{2}-y+1},\frac{y^{3}-y^{2}+y-2}{y\left(y^{2}-y+1\right)}\right\}\,,\end{split} (36)

and liw,i=12,,15l^{w}_{i},i=12,\ldots,15 are the roots of the polynomial 𝒬(y,w)\cal Q(y,w) defined by

𝒬(y,w)\displaystyle\cal Q(y,w) =w4(y2y+1)4y+2w3(y24y+1)(y2y+1)4\displaystyle=w^{4}\left(y^{2}-y+1\right)^{4}y+2w^{3}\left(y^{2}-4y+1\right)\left(y^{2}-y+1\right)^{4} (37)
2w2(y2y+1)2(y67y5+12y411y3+12y27y+1)\displaystyle-2w^{2}\left(y^{2}-y+1\right)^{2}\left(y^{6}-7y^{5}+12y^{4}-11y^{3}+12y^{2}-7y+1\right)
2w(y2y+1)2(y62y5+4y412y3+4y22y+1)\displaystyle-2w\left(y^{2}-y+1\right)^{2}\left(y^{6}-2y^{5}+4y^{4}-12y^{3}+4y^{2}-2y+1\right)
+2y1011y9+30y862y7+90y689y5+90y462y3+30y211y+2.\displaystyle+2y^{10}-11y^{9}+30y^{8}-62y^{7}+90y^{6}-89y^{5}+90y^{4}-62y^{3}+30y^{2}-11y+2\,.

Solving the differential equation in yy (where the dependence on ww drops out) gives a linear combination of MPLs G(;y)G(\ldots;y) with the letters from the set {l1y,,23}\{l^{y}_{1},\ldots,23\}, with

l1y=0,l6y=12(35),l7y=12(3+5),l8y=eiπ3,l9y=eiπ3,l10y=eiπ3,l11y=eiπ3,l12y=1,l13y=1,l24y=12(15),l25y=12(1+5),l26y=12(15),l27y=12(51).\begin{split}&\,l^{y}_{1}=0\,,\,l^{y}_{6}=\frac{1}{2}\left(3-\sqrt{5}\right)\,,\,l^{y}_{7}=\frac{1}{2}\left(3+\sqrt{5}\right)\,,\,l^{y}_{8}=e^{\frac{i\pi}{3}}\,,\,l^{y}_{9}=e^{-\frac{i\pi}{3}}\,,\,l^{y}_{10}=-e^{\frac{i\pi}{3}}\,,\\ &\,l^{y}_{11}=-e^{-\frac{i\pi}{3}}\,,\,l^{y}_{12}=1\,,\,l^{y}_{13}=-1\,,\,l^{y}_{24}=\frac{1}{2}\left(1-\sqrt{5}\right)\,,\,l^{y}_{25}=\frac{1}{2}\left(1+\sqrt{5}\right)\,,\\ &\,l^{y}_{26}=\frac{1}{2}\left(-1-\sqrt{5}\right)\,,\,l^{y}_{27}=\frac{1}{2}\left(\sqrt{5}-1\right)\,.\end{split} (38)

Moreover, liy,i=2,,5l^{y}_{i},i=2,\ldots,5 are the roots of 12y+y22y3+y41-2y+y^{2}-2y^{3}+y^{4}; liy,i=14,,17l^{y}_{i},i=14,\ldots,17 are the roots of 36y+5y26y3+3y43-6y+5y^{2}-6y^{3}+3y^{4}; liy,i=18,,20l^{y}_{i},i=18,\ldots,20 are the roots of 2+yy2+y3-2+y-y^{2}+y^{3} and liy,i=21,,23l^{y}_{i},i=21,\ldots,23 are the roots of 1+yy2+2y3-1+y-y^{2}+2y^{3}.

Following this procedure, we obtain complete analytical results for f37f_{37} at weights 3 and 4, and for the elements fif_{i} with i{35,36,38,39,41,43}i\in\{35,36,38,39,41,43\} at weight 4, in terms of MPLs in ww or yy with the letters defined above. We have checked our analytic results for the master integrals against FIESTA Smirnov:2015mct using also the forthcoming new release Smirnov:fiesta2021 . It turns out, however, (as it is easy to imagine from the dimension of the alphabet) that our results at weight 4 for these elements are rather complicated due to a very intricate branch-cut structure. In particular, evaluating them in this form at a phase-space point with GiNaC Bauer:2000cp ; Vollinga:2004sn meets problems connected both with timing and stability, so that our results for the complicated elements become impractical, and for phenomenological applications the numerical solution obtained using DiffExp might be preferable. For the same reason, we also recommend to turn to DiffExp also for the contribution of element 37 of weight 3. Of course, this does not imply that a better analytical representation in terms of MPLs does not exist, whose numerical evaluation could be much faster than automated numerical codes. Our analytic results obtained by direct integration of the differential equations could be used as a starting point to look for such a better analytic representation, if required.

5 Compact analytic results in terms of eMPLs

5.1 An analytic result for f14f_{14} in terms of eMPLs

As mentioned in the previous section, we could solve the differential equations for all master integrals analytically in terms of MPLs with algebraic arguments, except for f14f_{14}. The reason why we cannot easily apply the techniques from the previous section to f14f_{14} can be seen from the form of the differential equation itself. If we write f14(x,y)=ϵ4f¯(x,y)f_{14}(x,y)=\epsilon^{4}\,\bar{f}(x,y), then the differential equation in xx takes the form:

xf¯(x,y)=1(x1)xΔ(x,y){2(x+1)x(y21)G0(y)G0,0(x)\displaystyle\,\frac{\partial}{\partial x}\bar{f}(x,y)=\frac{1}{(x-1)x\sqrt{\Delta(x,y)}}\Bigg{\{}-2(x+1)x\left(y^{2}-1\right)G_{0}(y)G_{0,0}(x)
(x1)G0(x)[2(3x2y+x(y1)2+y)G0,0(y)+π2(x21)y]\displaystyle\,\phantom{+}(x-1)G_{0}(x)\Big{[}2\left(3x^{2}y+x(y-1)^{2}+y\right)G_{0,0}(y)+\pi^{2}\left(x^{2}-1\right)y\Big{]} (39)
(x+1)(x1)2y[2G0(y)(G1/y,0(x)Gy,0(x))+2G1/y,0,0(x)+2Gy,0,0(x)\displaystyle\,-(x+1)(x-1)^{2}y\Big{[}2G_{0}(y)\left(G_{-1/y,0}(x)-G_{-y,0}(x)\right)+2G_{-1/y,0,0}(x)+2G_{-y,0,0}(x)
+2G0,0,0(x)+4G1,0,0(x)4G0,0,0(y)+4G1,0,0(y)+2ζ3+(2G0,0(y)+π2)G1/y(x)\displaystyle\,+2G_{0,0,0}(x)+4G_{1,0,0}(x)-4G_{0,0,0}(y)+4G_{1,0,0}(y)+2\zeta_{3}+\left(2G_{0,0}(y)+\pi^{2}\right)G_{-1/y}(x)
+(2G0,0(y)+π2)Gy(x)]},\displaystyle\,+\left(2G_{0,0}(y)+\pi^{2}\right)G_{-y}(x)\Big{]}\Bigg{\}}\,,

with

Δ(x,y)=(x+y)(xy+1)(x2y+xy24xy+x+y).\Delta(x,y)=(x+y)(xy+1)\left(x^{2}y+xy^{2}-4xy+x+y\right)\,. (40)

Due to the symmetry f14(x,y)=f14(y,x)f_{14}(x,y)=f_{14}(y,x), the differential equation with respect to yy can easily be obtained by exchanging the roles of xx and yy in eq. (39). The square root in eq. (39) is identical to the square root that has appeared in ref. Henn:2013woa in the computation of the first planar two-loop family for Bhabha scattering. The polynomial under the square root has degree four in both xx and yy, so it cannot be rationalised in one variable only. However, this is not sufficient to exclude that one cannot rationalise it by performing a change of variable involving simultaneously xx and yy. In ref. Festi:2018qip it was shown that the algebraic variety defined by ξ2=Δ(x,y)\xi^{2}=\Delta(x,y) is a K3 surface. As a consequence, it cannot be rationalised by any rational change of variables, and the strategy of rationalising all square roots and evaluating all iterated integrals in terms of MPLs using ideas from section 3.1 cannot be applied. We can of course also obtain numerical results for f14f_{14} by solving the canonical system with DiffExp (as described in the previous section), but it would of course be interesting to obtain analytic results also for f14f_{14}.

We observe that the structure of the square root matches precisely the criteria of section 3.2. Indeed, the polynomial under the square root in eq. (39) has degree four in both xx and yy. So, if we keep one variable fixed, the square root defines an elliptic curve in the other. This reflects the fact that the K3 surface defined by the square root is elliptically fibered Festi:2018qip . As a consequence, we can choose the integration path in eq. (15) and solve the differential equation in terms of eMPLs. Before we do this, we have to make an important comment. While the strategy of section 3.1 to solve eq. (39) does not apply, it does not exclude that a solution in terms of MPLs evaluated at algebraic arguments exists. On the contrary, we have found that it is possible to evaluate f14f_{14} from its Feynman parameter representation using direct integration techniques (cf., e.g., refs. Brown:2009qja ; Brown:2008um ; Anastasiou:2013srw ; Ablinger:2014yaa ; Bogner:2014mha ; Panzer:2014caa ; Duhr:2019tlz ).444We are grateful to Erik Panzer for providing us with a change of variables that renders the Feynman parameter integral linearly reducible. The resulting expression, however, is extremely lengthy and involves MPLs evaluated at complicated algebraic arguments. We have not been able to confirm the final expression numerically, as its size and the complexity of the branch cuts renders the evaluation of the MPLs extremely challenging. Our representation obtained from direct integration is therefore not useful for practical purposes. As we will show now, the representation in terms of eMPLs is very compact.

Let us now explain how we can solve the system of differential equations for f¯\bar{f} in terms of MPLs and eMPLs. We first of all introduce new variables (x¯,y¯)=(1x,1y)(\bar{x},\bar{y})=(1-x,1-y), and use PolyLogTools Duhr:2019tlz to express all MPLs of the form G(;x)G(\ldots;{x}) or G(;y)G(\ldots;{y}) in eq. (39) in terms of G(;x¯)G(\ldots;\bar{x}) and G(;y¯)G(\ldots;\bar{y}) respectively. For example, we find:

Gy,0,0(x)=G2y¯,1,1(x¯)π212G1(y¯)G1,1,1(y¯)+G1,1,2(y¯)+log2G1,1(y¯)+34ζ3.\begin{split}G_{-y,0,0}(x)&\,=G_{2-\bar{y},1,1}(\bar{x})-\frac{\pi^{2}}{12}\,G_{1}(\bar{y})-G_{1,1,1}(\bar{y})+G_{1,1,2}(\bar{y})+\log 2\,G_{1,1}(\bar{y})+\frac{3}{4}\,\zeta_{3}\,.\end{split} (41)

We know from eq. (34) that f14f_{14} must vanish for s=t=0s=t=0, and so f¯(x¯=0,y¯=0)=0\bar{f}(\bar{x}=0,\bar{y}=0)=0. Our strategy is then as follows. We first use the differential equation in y¯\bar{y} to evolve f¯\bar{f} from (x¯,y¯)=(0,0)(\bar{x},\bar{y})=(0,0) to (x¯,y¯)=(0,y¯0)(\bar{x},\bar{y})=(0,\bar{y}_{0}). We then use the differential equation in x¯\bar{x} to evolve from (x¯,y¯)=(0,y¯0)(\bar{x},\bar{y})=(0,\bar{y}_{0}) to the generic point (x¯,y¯)=(x¯0,y¯0)(\bar{x},\bar{y})=(\bar{x}_{0},\bar{y}_{0}). In the following we assume 0<x¯0,y¯0<10<\bar{x}_{0},\bar{y}_{0}<1 for concreteness (which corresponds to he Euclidean region s,t<0s,t<0). On the line x¯=0\bar{x}=0 we find

Δ(x=1,y)=(1y2)2=y¯2(2y¯)2.\Delta(x=1,y)=(1-y^{2})^{2}=\bar{y}^{2}(2-\bar{y})^{2}\,. (42)

Hence, the square root disappears in the limit x¯0\bar{x}\to 0, and we can solve the differential equation on the line x¯{\bar{x}} in terms of MPLs. We find (relabelling y¯0\bar{y}_{0} as y¯\bar{y}):

f¯(x¯=0,y¯)=(2π2log23ζ3)G1(y¯)π2G1,1(y¯)+2π2G1,2(y¯)+4G1,0,1,1(y¯)4G1,1,1,1(y¯)+4G1,2,1,1(y¯).\begin{split}\bar{f}(\bar{x}=0,\bar{y})&\,=\left(2\pi^{2}\log 2-3\zeta_{3}\right)\,G_{1}(\bar{y})-\pi^{2}\,G_{1,1}(\bar{y})+2\pi^{2}\,G_{1,2}(\bar{y})\\ &\,+4\,G_{1,0,1,1}(\bar{y})-4\,G_{1,1,1,1}(\bar{y})+4\,G_{1,2,1,1}(\bar{y})\,.\end{split} (43)

Next, let us discuss the solution on the line y¯=y¯0\bar{y}=\bar{y}_{0}. Clearly, Δ(x,y)\Delta(x,y) is not a perfect square for x0x\neq 0. Looking at the constraint ξ2=Δ(x,y0)\xi^{2}=\Delta(x,y_{0}) as a function of xx with y0y_{0} fixed, we see that it defines an elliptic curve. The branch points, i.e., the zeroes in x¯\bar{x} of Δ(1x¯,1y¯0)\Delta(1-\bar{x},1-\bar{y}_{0}), are

a(y¯0)=(2y¯0,y¯0(y¯0y¯02+4y¯04)2(1y¯0),y¯0(y¯0+y¯02+4y¯04)2(1y¯0),2y¯01y¯0).\vec{a}(\bar{y}_{0})=\left(2-\bar{y}_{0},\frac{\bar{y}_{0}\left(\bar{y}_{0}-\sqrt{\bar{y}_{0}^{2}+4\bar{y}_{0}-4}\right)}{2\left(1-\bar{y}_{0}\right)},\frac{\bar{y}_{0}\left(\bar{y}_{0}+\sqrt{\bar{y}_{0}^{2}+4\bar{y}_{0}-4}\right)}{2\left(1-\bar{y}_{0}\right)},\frac{2-\bar{y}_{0}}{1-\bar{y}_{0}}\right)\,. (44)

We can use eq. (26) to interpret every MPL of the form G(b(y¯);x¯)G(\vec{b}(\bar{y});\bar{x}) as an eMPL of the form 4(nc(y¯);x¯,a(y¯)){\mathcal{E}_{4}}\!\left(\begin{smallmatrix}\vec{n}\\ \vec{c}(\bar{y})\end{smallmatrix};\bar{x},\vec{a}(\bar{y})\right). Moreover, we can express all the algebraic coefficients multiplying the MPLs in the differential equation in terms of the functions Ψ±1(c,x¯,a(y¯))\Psi_{\pm 1}(c,\bar{x},\vec{a}(\bar{y})). For example, we find

Ψ1(,x¯,a(y¯))=x¯ξ(y+1)24yξ,Ψ1(0,x¯,a(y¯))=y21x¯yξy212yξ,Ψ1(1,x¯,a(y¯))=(y1)24yξ1(x¯1)ξ,\begin{split}\Psi_{-1}(\infty,\bar{x},\vec{a}(\bar{y}))&\,=\frac{\bar{x}}{\xi}-\frac{(y+1)^{2}}{4y\xi}\,,\\ \Psi_{-1}(0,\bar{x},\vec{a}(\bar{y}))&\,=\frac{y^{2}-1}{\bar{x}y\xi}-\frac{y^{2}-1}{2y\xi}\,,\\ \Psi_{-1}(1,\bar{x},\vec{a}(\bar{y}))&\,=\frac{(y-1)^{2}}{4y\xi}-\frac{1}{(\bar{x}-1)\xi}\,,\end{split} (45)

with ξ=Δ(1x¯,1y¯)\xi=\sqrt{\Delta(1-\bar{x},1-\bar{y})}. In the process, we discover the relations:

Z4(0,a(y¯))=(2y¯)y¯2c4(1y¯)=1y22yc4,Z4(1,a(y¯))=y¯24c4(1y¯)=(1y)24yc4,G(a(y¯))=(2y¯)(3y¯2)8c4(1y¯)=(1y)(13y)8yc4.\begin{split}Z_{4}(0,\vec{a}(\bar{y}))&\,=\frac{\left(2-\bar{y}\right)\bar{y}}{2c_{4}\left(1-\bar{y}\right)}=\frac{1-y^{2}}{2yc_{4}}\,,\\ Z_{4}(1,\vec{a}(\bar{y}))&\,=\frac{\bar{y}^{2}}{4c_{4}\left(1-\bar{y}\right)}=\frac{(1-y)^{2}}{4yc_{4}}\,,\\ G_{\ast}(\vec{a}(\bar{y}))&\,=\frac{\left(2-\bar{y}\right)\left(3\bar{y}-2\right)}{8c_{4}\ \left(1-\bar{y}\right)}=\frac{(1-y)(1-3y)}{8yc_{4}}\,.\end{split} (46)

At this point we need to make an important comment about how we choose the branches of the square root ξ=1x¯,1y¯)\xi=\sqrt{1-\bar{x},1-\bar{y})}. From eq. (44) we can see that for 0<y<3220<y<3-2\sqrt{2}, the four branch points are real and ordered according to a1(y)<a2(y)<a3(y)<a4(y)a_{1}(y)<a_{2}(y)<a_{3}(y)<a_{4}(y). The branches of the square root are chosen according to

ξ=Δ(x,y)=|Δ(x,y)|×{1,x<a1(y) or xa4(y),i,a1(y)x<a2(y),1,a2(y)x<a3(y),i,a3(y)x<a4(y).\xi=\sqrt{\Delta(x,y)}=\sqrt{|\Delta(x,y)|}\times\left\{\begin{array}[]{ll}-1\,,&x<a_{1}(y)\textrm{ or }x\geq a_{4}(y)\,,\\ -i\,,&a_{1}(y)\leq x<a_{2}(y)\,,\\ \phantom{-}1\,,&a_{2}(y)\leq x<a_{3}(y)\,,\\ \phantom{-}i\,,&a_{3}(y)\leq x<a_{4}(y)\,.\end{array}\right. (47)

For 323<y<13-2\sqrt{3}<y<1, we have a1(y¯0),a4(y¯0)>0a_{1}(\bar{y}_{0}),a_{4}(\bar{y}_{0})>0 and a2(y¯0)=a3(y¯0)a_{2}(\bar{y}_{0})^{*}=a_{3}(\bar{y}_{0}), with Im a3(y¯0)>0\textrm{Im }a_{3}(\bar{y}_{0})>0. The branches of the square are then chosen as:

ξ=Δ(x,y)=|Δ(x,y)|×{1,x<a1(y) or xa4(y),i,a1(y)x<a4(y).\xi=\sqrt{\Delta(x,y)}=\sqrt{|\Delta(x,y)|}\times\left\{\begin{array}[]{ll}-1\,,&x<a_{1}(y)\textrm{ or }x\geq a_{4}(y)\,,\\ -i\,,&a_{1}(y)\leq x<a_{4}(y)\,.\end{array}\right. (48)

We stress that part of this choice is purely conventional, and compensated by the leading singularity 1/Δ(x,y)1/\sqrt{\Delta(x,y)} when passing from the master integral g14g_{14} to its canonical analogue f14f_{14}.

The resulting differential equation can be solved by quadrature, and all integrals over x¯\bar{x} can be performed in terms of eMPLs using eq. (21). The final result reads (we relabel again (x0,y0)(x_{0},y_{0}) as (x,y)(x,y)):

f¯(x,y)\displaystyle\bar{f}(x,y) =24(11111y+111;x¯,a)+24(1111y+111;x¯,a)+24(111111y+111;x¯,a)\displaystyle\,=2{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&\frac{1}{y}+1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+2{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&y+1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+2{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&\frac{1}{y}+1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)
+24(11111y+111;x¯,a)+44(1111011;x¯,a)44(1111111;x¯,a)\displaystyle\,+2{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&y+1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+4{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&0&1&1\end{smallmatrix};\bar{x},\vec{a}\right)-4{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)
+44(11111011;x¯,a)44(11111111;x¯,a)(3log2y+π2)4(111;x¯,a)\displaystyle\,+4{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&0&1&1\end{smallmatrix};\bar{x},\vec{a}\right)-4{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)-\left(3\log^{2}y+\pi^{2}\right){\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ \infty&1\end{smallmatrix};\bar{x},\vec{a}\right)
+(log2y+π2)[4(111y+1;x¯,a)+4(11y+1;x¯,a)+4(1111y+1;x¯,a)\displaystyle\,+\left(\log^{2}y+\pi^{2}\right)\Big{[}{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ \infty&\frac{1}{y}+1\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ \infty&y+1\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 1&\frac{1}{y}+1\end{smallmatrix};\bar{x},\vec{a}\right) (49)
+4(111y+1;x¯,a)]+(log2yπ2)4(1111;x¯,a)+2logy[4(1111y+11;x¯,a)\displaystyle\,\quad+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 1&y+1\end{smallmatrix};\bar{x},\vec{a}\right)\Big{]}+\left(\log^{2}y-\pi^{2}\right){\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 1&1\end{smallmatrix};\bar{x},\vec{a}\right)+2\log y\Big{[}{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ \infty&\frac{1}{y}+1&1\end{smallmatrix};\bar{x},\vec{a}\right)
4(111y+11;x¯,a)+24(111011;x¯,a)+4(11111y+11;x¯,a)4(1111y+11;x¯,a)]\displaystyle\,\quad-{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ \infty&y+1&1\end{smallmatrix};\bar{x},\vec{a}\right)+2{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ 0&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ 1&\frac{1}{y}+1&1\end{smallmatrix};\bar{x},\vec{a}\right)-{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ 1&y+1&1\end{smallmatrix};\bar{x},\vec{a}\right)\Big{]}
+[4Li3(y)4Li3(y)+4Li2(y)logy+4Li2(y)logy23log3y\displaystyle\,+\Big{[}-4\text{Li}_{3}(-y)-4\text{Li}_{3}(y)+4\text{Li}_{2}(-y)\log y+4\text{Li}_{2}(y)\log y-\frac{2}{3}\log^{3}y
+2log(1y)log2y+2log(y+1)log2yπ2logy+2π2log(y+1)\displaystyle\,\quad+2\log(1-y)\log^{2}y+2\log(y+1)\log^{2}y-\pi^{2}\log y+2\pi^{2}\log(y+1)
2ζ3][4(1;x¯,a)+4(11;x¯,a)]\displaystyle\,\quad-2\zeta_{3}\Big{]}\Big{[}{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1\\ \infty\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1\\ 1\end{smallmatrix};\bar{x},\vec{a}\right)\Big{]}
12Li4(y)12Li4(y)2Li2(y)log2y2Li2(y)(log2y+π2)\displaystyle\,-12\text{Li}_{4}(-y)-12\text{Li}_{4}(y)-2\text{Li}_{2}(y)\log^{2}y-2\text{Li}_{2}(-y)\left(\log^{2}y+\pi^{2}\right)
+8Li3(y)logy+8Li3(y)logy2ζ3logy16log4y12π2log2y3π420,\displaystyle\,+8\text{Li}_{3}(-y)\log y+8\text{Li}_{3}(y)\log y-2\zeta_{3}\log y-\frac{1}{6}\log^{4}y-\frac{1}{2}\pi^{2}\log^{2}y-\frac{3\pi^{4}}{20}\,,

where we introduced the shorthand aa(y¯)\vec{a}\equiv\vec{a}(\bar{y}), and we have replaced all MPLs by classical polylogarithms:

Lin(z)=G(0,,0n1,1;z).\textrm{Li}_{n}(z)=-G(\underbrace{0,\ldots,0}_{n-1},1;z)\,. (50)

We have checked numerically that our final result for f¯(x,y)\bar{f}(x,y) is correct by comparing against Fiesta. The eMPLs in eq. (49) were evaluated numerically using an in-house code (a public numerical implementation of eMPLs into GiNaC exists Walden:2020odh ). Note that f¯(x,y)\bar{f}(x,y) is a pure function of uniform weight four Broedel:2018qkq . We find it interesting that such a compact analytic expression in terms of eMPLs exists, while our MPL expression obtained by naive direct integration is prohibitively large.

5.2 A compact analytic result for the first planar family

Refer to caption
Figure 3: The graph associated with the integral B(s,t,m2)B(s,t,m^{2}) in eq. (51).

Let us conclude this paper by applying the techniques from section 5.1 to the planar family in fig. 1 (a). In ref. Henn:2013woa it was shown that all integrals in this family can be expressed in terms of 23 master integrals, and all but one master integral were evaluated in terms of MPLs. The remaining master integral that could not be expressed in terms of MPLs in ref. Henn:2013woa is (see fig. 3):

B(s,t,m2)\displaystyle B(s,t,m^{2}) =e2γEϵπDdDk1dDk2[(k1+p1+p2)2m2][k22m2](k1+p1)2(k1k2)2(k2p3)2.\displaystyle\,=\frac{e^{2\gamma_{E}\epsilon}}{\pi^{D}}\int\frac{d^{D}k_{1}\,d^{D}k_{2}}{[(k_{1}+p_{1}+p_{2})^{2}-m^{2}][k_{2}^{2}-m^{2}](k_{1}+p_{1})^{2}(k_{1}-k_{2})^{2}(k_{2}-p_{3})^{2}}\,. (51)

This integral was evaluated in terms of a rather lengthy combination of MPLs with algebraic arguments in ref. Heller:2019gkq . Here we show that, by using the same strategy as in section 5.1, we can obtain a very compact representation for B(s,t,m2)B(s,t,m^{2}) in terms of the same type of eMPLs as in eq. (49).

The integral in eq. (51) is finite in four dimensions. Let us define B~(x,y)\widetilde{B}(x,y) by

B(s,t,m2)=14(s+t)(s+t4m2)B~(x,y)+𝒪(ϵ),B(s,t,m^{2})=\frac{1}{4\sqrt{(s+t)(s+t-4m^{2})}}\,\widetilde{B}(x,y)+\cal O(\epsilon)\,, (52)

where the Landau variables (x,y)(x,y) have been defined in eq. (33). Our starting point is the differential equation satisfied by B~(x,y)\widetilde{B}(x,y) Henn:2013woa :

dB~(x,y)\displaystyle d\widetilde{B}(x,y) =g1dlog1Q(x,y)1+Q(x,y)+g2dlog(1+x)+(1x)Q(x,y)(1+x)(1x)Q(x,y)\displaystyle\,=g_{1}\,d\log\frac{1-Q(x,y)}{1+Q(x,y)}+g_{2}\,d\log\frac{(1+x)+(1-x)Q(x,y)}{(1+x)-(1-x)Q(x,y)} (53)
+g3dlog(1+y)+(1y)Q(x,y)(1+y)(1y)Q(x,y).\displaystyle\,+g_{3}\,d\log\frac{(1+y)+(1-y)Q(x,y)}{(1+y)-(1-y)Q(x,y)}\,. (54)

The functions gig_{i} appearing in the right-hand side of eq. (53) are pure combinations of MPLs of weight three,

g1=16G0(y)G0,0(x)+32G1(y)G0,0(x)+8G0,0,0(y)+16G0,0,1(y)32G0,1,0(x)+8G0,0,0(x)+16G1,0,0(x)+83π2G0(y)163π2G1(y)83π2G0(x)32ζ3,g2=16G0,0,0(x)83π2G0(x),g3=8G0(x)G1x,0(y)+16G0(x)G1x,1(y)8G0(x)Gx,0(y)43π2G0(x)16G0(x)Gx,1(y)+16G1,0(x)G1x(y)16G1,0(x)Gx(y)24ζ38G0(y)G0,0(x)8G0,0(x)G1x(y)+8G0,0(x)Gx(y)+8G1x,0,0(y)+16G1x,0,1(y)+8Gx,0,0(y)+16Gx,0,1(y)16G1,0,0(y)32G1,0,1(y)16G0,1,0(x)+8G0,0,0(x)+203π2G1x(y)+4π2Gx(y)323π2G1(y).\begin{split}g_{1}&\,=-16\,G_{0}(y)\,G_{0,0}(x)+32\,G_{1}(y)\,G_{0,0}(x)+8\,G_{0,0,0}(y)+16\,G_{0,0,1}(y)\\ &\,-32\,G_{0,-1,0}(x)+8\,G_{0,0,0}(x)+16\,G_{1,0,0}(x)+\frac{8}{3}\pi^{2}\,G_{0}(y)-\frac{16}{3}\pi^{2}\,G_{1}(y)\\ &\,-\frac{8}{3}\pi^{2}\,G_{0}(x)-32\,\zeta_{3}\,,\\ g_{2}&\,=-16\,G_{0,0,0}(x)-\frac{8}{3}\pi^{2}\,G_{0}(x)\,,\\ g_{3}&\,=8\,G_{0}(x)\,G_{-\frac{1}{x},0}(y)+16\,G_{0}(x)\,G_{-\frac{1}{x},1}(y)-8\,G_{0}(x)\,G_{-x,0}(y)-\frac{4}{3}\pi^{2}\,G_{0}(x)\\ &\,-16\,G_{0}(x)\,G_{-x,1}(y)+16\,G_{-1,0}(x)\,G_{-\frac{1}{x}}(y)-16\,G_{-1,0}(x)\,G_{-{x}}(y)-24\,\zeta_{3}\\ &\,-8\,G_{0}(y)\,G_{0,0}(x)-8\,G_{0,0}(x)\,G_{-\frac{1}{x}}(y)+8\,G_{0,0}(x)\,G_{-x}(y)+8\,G_{-\frac{1}{x},0,0}(y)\\ &\,+16\,G_{-\frac{1}{x},0,1}(y)+8\,G_{-x,0,0}(y)+16\,G_{-x,0,1}(y)-16\,G_{-1,0,0}(y)-32\,G_{-1,0,1}(y)\\ &\,-16\,G_{0,-1,0}(x)+8\,G_{0,0,0}(x)+\frac{20}{3}\pi^{2}\,G_{-\frac{1}{x}}(y)+4\pi^{2}\,G_{-x}(y)-\frac{32}{3}\pi^{2}\,G_{-1}(y)\,.\end{split} (55)

The function Q(x,y)Q(x,y) is related to exactly the same square root defined in eqs. (39) and (40):

Q(x,y)=(x+y)(1+xy)x2y+xy24xy+x+y=Δ(x,y)x2y+xy24xy+x+y,Q(x,y)=\sqrt{\frac{(x+y)(1+xy)}{x^{2}y+xy^{2}-4xy+x+y}}=\frac{\sqrt{\Delta(x,y)}}{x^{2}y+xy^{2}-4xy+x+y}\,, (56)

and the initial condition is B~(x=1,y=1)=0\widetilde{B}(x=1,y=1)=0. Following exactly the same steps as in section 5.1, we find the following very compact expression:

B~(x,y)\displaystyle\widetilde{B}(x,y) =16logtm2[4(11101+1/y1;x¯,a)4(11101+y1;x¯,a)\displaystyle\,=16\log\frac{-t}{m^{2}}\,\Big{[}{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ 0&1+{1}/{y}&1\end{smallmatrix};\bar{x},\vec{a}\right)-{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ 0&1+y&1\end{smallmatrix};\bar{x},\vec{a}\right) (57)
+4(11111;x¯,a)+4(111111;x¯,a)+ζ24(1;x¯,a)+ζ24(11;x¯,a)]\displaystyle\,+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ \infty&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1\\ 1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+\zeta_{2}\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1\\ \infty\end{smallmatrix};\bar{x},\vec{a}\right)+\zeta_{2}\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1\\ 1\end{smallmatrix};\bar{x},\vec{a}\right)\Big{]}
8(8ζ2+4Li2(y)+log2y)[4(1101+1/y;x¯,a)+4(1101+y;x¯,a)4(1101;x¯,a)]\displaystyle\,-8\left(8\zeta_{2}+4\text{Li}_{2}(y)+\log^{2}y\right)\!\left[{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 0&1+{1}/{y}\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 0&1+y\end{smallmatrix};\bar{x},\vec{a}\right)-{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 0&1\end{smallmatrix};\bar{x},\vec{a}\right)\right]
32ζ2[4(111;x¯,a)4(1111;x¯,a)]+164(111101+1/y11;x¯,a)\displaystyle\,-32\,\zeta_{2}\,\left[{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ \infty&1\end{smallmatrix};\bar{x},\vec{a}\right)-{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1\\ 1&1\end{smallmatrix};\bar{x},\vec{a}\right)\right]+16\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 0&1+{1}/{y}&1&1\end{smallmatrix};\bar{x},\vec{a}\right)
324(111101+1/y21;x¯,a)164(111101+y11;x¯,a)+324(111101+y21;x¯,a)\displaystyle\,-32\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 0&1+{1}/{y}&2&1\end{smallmatrix};\bar{x},\vec{a}\right)-16\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 0&1+y&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+32\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 0&1+y&2&1\end{smallmatrix};\bar{x},\vec{a}\right)
+164(1111011;x¯,a)244(1111111;x¯,a)324(1111121;x¯,a)\displaystyle\,+16\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&0&1&1\end{smallmatrix};\bar{x},\vec{a}\right)-24\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)-32\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ \infty&1&2&1\end{smallmatrix};\bar{x},\vec{a}\right)
+164(11111011;x¯,a)+404(11111111;x¯,a)324(11111121;x¯,a)\displaystyle\,+16\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&0&1&1\end{smallmatrix};\bar{x},\vec{a}\right)+40\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&1&1&1\end{smallmatrix};\bar{x},\vec{a}\right)-32\,{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1&1&1&1\\ 1&1&2&1\end{smallmatrix};\bar{x},\vec{a}\right)
+43(12Li3(y)+24ζ2logy+log3y)[4(1;x¯,a)+4(11;x¯,a)]\displaystyle\,+\frac{4}{3}\big{(}12\text{Li}_{3}(y)+24\zeta_{2}\log y+\log^{3}y\big{)}\left[{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1\\ \infty\end{smallmatrix};\bar{x},\vec{a}\right)+{\mathcal{E}_{4}}\!\left(\begin{smallmatrix}-1\\ 1\end{smallmatrix};\bar{x},\vec{a}\right)\right]
+64ζ432ζ2Li2(y)+16Li4(y)+8ζ2log2y+13log4y.\displaystyle\,+64\zeta_{4}-32\zeta_{2}\text{Li}_{2}(y)+16\text{Li}_{4}(y)+8\zeta_{2}\log^{2}y+\frac{1}{3}\log^{4}y\,.

We have again checked our result numerically. Note that again we find a pure function of uniform weight four.

6 Conclusions

In this paper we considered the computation of the second family of planar master integrals relevant for Bhabha scattering at two loops in QED including the full dependence on the electron mass. Our primary tool for their analytic study was the method of differential equations, augmented by the choice of a canonical basis. We described how we obtained our canonical basis and showed that four different square roots appear in the calculation of the 43 master integrals that comprise it. We also provided boundary conditions at s=t=0s=t=0 for all master integrals. Together with the matrices defining the differential equations, this input is sufficient to produce high-precision numerical results for all master integrals and for all kinematic regions using the public code DiffExp. We provide a Mathematica notebook that allows one to evaluate all master integrals numerically with DiffExp as ancillary material with the arXiv submission.

We also considered the analytic solution of the differential equations. Interestingly, the four square roots never appear all at the same time, and the differential equations for all master integrals but one can be solved in terms of MPLs by rationalising three of the four square roots by suitable changes of variables. For the contributions up to weight three, this procedure leads to analytic results which we present with the arXiv submission. While conceptually straightforward, we find that this procedure generates rather involved analytical expressions for the weight four part of the master integrals. The analytic results are available in electronic form from the authors upon request.

In the last part of the paper, we focused on the analytic computation of the remaining master integral, whose canonical differential equations contain three different square roots which cannot be rationalised at the same time. An analytic calculation in terms of MPLs cannot be easily obtained by direct integration of the differential equations due to the non-rationalisable square root. Instead, we show that compact analytic expressions can be obtained algorithmically in terms of elliptic multiple polylogarithms. We applied this idea in detail to our problem and obtained in this way a very compact analytic expression for this integral. We also showed that a similar compact expression can be obtained for another planar integral relevant for two-loop Bhabha scattering, whose calculation in terms of (a lengthy combination of) MPLs had been considered some years ago.

This paper concludes the analytic calculation of the planar master integrals for Bhabha scattering at two loops. For the future, it would be interesting to complete also the computation of the non-planar two-loop family. Once this last step is achieved, we will have all the ingredient to obtain for the first time complete two-loop results in QED for one of the standard candle processes at an electron-positron collider.

Acknowledgments

We are grateful to Martijn Hidding for help in applying DiffExp and to Erik Panzer for suggesting the change of variable that renders the integral f14f_{14} linearly reducible. We thank Marco Besier, Jake Bourjaily, Johannes Brödel, Falko Dulat and Brenda Penante for discussions. The work of V.S. was supported by the Russian Science Foundation, agreement no. 21-71-30003 (checking results with updated version of FIESTA) and by the Ministry of Education and Science of the Russian Federation as part of the program of the Moscow Center for Fundamental and Applied Mathematics under agreement no. 075-15-2019-1621 (solving differential equations).

Appendix A Canonical basis

In this appendix we show our choice of canonical basis. Note that we prefer to choose the square roots in such a form that that they are manifestly real in the Euclidean region s,t<0s,t<0. This holds also for the boundary conditions which we presented in eq (34).

f1\displaystyle f_{1} =ϵ2F2,0,0,0,0,2,0,0,0,\displaystyle\,=\epsilon^{2}F_{2,0,0,0,0,2,0,0,0}\,, (58)
f2\displaystyle f_{2} =ϵ212s4m2sF0,2,1,0,0,2,0,0,0ϵ2s4m2sF0,2,2,0,0,1,0,0,0,\displaystyle\,=-\epsilon^{2}\frac{1}{2}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,2,1,0,0,2,0,0,0}-\epsilon^{2}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,2,2,0,0,1,0,0,0}\,,
f3\displaystyle f_{3} =ϵ2sF0,2,1,0,0,2,0,0,0,\displaystyle\,=-\epsilon^{2}sF_{0,2,1,0,0,2,0,0,0}\,,
f4\displaystyle f_{4} =12ϵ2t4m2tF0,0,0,0,1,2,2,0,0ϵ2t4m2tF0,0,0,0,2,1,2,0,0,\displaystyle\,=-\frac{1}{2}\epsilon^{2}\sqrt{-t}\sqrt{4m^{2}-t}F_{0,0,0,0,1,2,2,0,0}-\epsilon^{2}\sqrt{-t}\sqrt{4m^{2}-t}F_{0,0,0,0,2,1,2,0,0}\,,
f5\displaystyle f_{5} =ϵ2tF0,0,0,0,1,2,2,0,0,\displaystyle\,=-\epsilon^{2}tF_{0,0,0,0,1,2,2,0,0}\,,
f6\displaystyle f_{6} =ϵ2m2F0,0,1,0,2,2,0,0,0\displaystyle\,=-\epsilon^{2}m^{2}F_{0,0,1,0,2,2,0,0,0}\,
f7\displaystyle f_{7} =ϵ3s4m2sF0,1,1,0,1,2,0,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,1,1,0,1,2,0,0,0}\,,
f8\displaystyle f_{8} =ϵ3t4m2tF0,0,1,0,1,2,1,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-t}\sqrt{4m^{2}-t}F_{0,0,1,0,1,2,1,0,0}\,,
f9\displaystyle f_{9} =ϵ2m2F2,0,0,0,0,2,1,0,0,\displaystyle\,=-\epsilon^{2}m^{2}F_{2,0,0,0,0,2,1,0,0}\,,
f10\displaystyle f_{10} =ϵ2m2F2,0,0,0,0,2,1,0,0,\displaystyle\,=-\epsilon^{2}m^{2}F_{2,0,0,0,0,2,1,0,0}\,,
f11\displaystyle f_{11} =ϵ3s4m2sF0,1,1,0,0,2,1,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,1,1,0,0,2,1,0,0}\,,
f12\displaystyle f_{12} =2ϵ3tF0,1,0,0,1,1,2,0,0ϵ2tF0,1,0,1,1,2,2,0,0,\displaystyle\,=2\epsilon^{3}tF_{0,1,0,0,1,1,2,0,0}-\epsilon^{2}tF_{0,1,0,-1,1,2,2,0,0}\,,
f13\displaystyle f_{13} =ϵ3t4m2tF0,1,0,0,1,1,2,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-t}\sqrt{4m^{2}-t}F_{0,1,0,0,1,1,2,0,0}\,,
f14\displaystyle f_{14} =ϵ4st4m2stF0,1,1,0,1,1,1,0,0,\displaystyle\,=-\epsilon^{4}\sqrt{-s-t}\sqrt{4m^{2}-s-t}F_{0,1,1,0,1,1,1,0,0}\,,
f15\displaystyle f_{15} =12ϵ3t4m2t(2m2F0,1,1,0,1,2,1,0,0+2m2F0,2,1,0,1,1,1,0,0sF0,1,1,0,1,2,1,0,0),\displaystyle\,=\frac{1}{2}\epsilon^{3}\sqrt{-t}\sqrt{4m^{2}-t}\left(2m^{2}F_{0,1,1,0,1,2,1,0,0}+2m^{2}F_{0,2,1,0,1,1,1,0,0}-sF_{0,1,1,0,1,2,1,0,0}\right)\,,
f16\displaystyle f_{16} =ϵ3st4m2s4m2tF0,1,1,0,1,2,1,0,0,\displaystyle\,=\epsilon^{3}\sqrt{-s}\sqrt{-t}\sqrt{4m^{2}-s}\sqrt{4m^{2}-t}F_{0,1,1,0,1,2,1,0,0}\,,
f17\displaystyle f_{17} =12ϵ3s4m2s(2m2F0,1,1,0,1,1,2,0,0+2m2F0,1,1,0,1,2,1,0,0tF0,1,1,0,1,2,1,0,0),\displaystyle\,=\frac{1}{2}\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}\left(2m^{2}F_{0,1,1,0,1,1,2,0,0}+2m^{2}F_{0,1,1,0,1,2,1,0,0}-tF_{0,1,1,0,1,2,1,0,0}\right)\,,
f18\displaystyle f_{18} =ϵ3m2(4m2t)2(4m2st)F0,0,1,0,1,2,1,0,0ϵ3(4m2t)(2m2st)2(4m2st)F0,1,0,0,1,1,2,0,0\displaystyle\,=\frac{\epsilon^{3}m^{2}\left(4m^{2}-t\right)}{2\left(4m^{2}-s-t\right)}F_{0,0,1,0,1,2,1,0,0}-\frac{\epsilon^{3}\left(4m^{2}-t\right)\left(2m^{2}-s-t\right)}{2\left(4m^{2}-s-t\right)}F_{0,1,0,0,1,1,2,0,0}
+ϵ3m2s(4m2s)t(4m2st)F0,1,1,0,0,2,1,0,0ϵ3(ϵ10m2m23ϵs3ϵt)F0,1,1,0,1,1,1,0,0\displaystyle\,+\frac{\epsilon^{3}m^{2}s\left(4m^{2}-s\right)}{t\left(4m^{2}-s-t\right)}F_{0,1,1,0,0,2,1,0,0}-\epsilon^{3}\left(\epsilon 10m^{2}-m^{2}-3\epsilon s-3\epsilon t\right)F_{0,1,1,0,1,1,1,0,0}
ϵ3m2(s+t)tF0,2,1,0,1,1,1,0,1+ϵ2m2(4m2t)4(4m2st)F0,0,0,0,1,2,2,0,0\displaystyle\,-\frac{\epsilon^{3}m^{2}(s+t)}{t}F_{0,2,1,0,1,1,1,0,-1}+\frac{\epsilon^{2}m^{2}\left(4m^{2}-t\right)}{4\left(4m^{2}-s-t\right)}F_{0,0,0,0,1,2,2,0,0}
+ϵ3m2(8m44m2s6m2t+s2+2st+t2)4m2stF0,2,1,0,1,1,1,0,0\displaystyle\,+\frac{\epsilon^{3}m^{2}\left(8m^{4}-4m^{2}s-6m^{2}t+s^{2}+2st+t^{2}\right)}{4m^{2}-s-t}F_{0,2,1,0,1,1,1,0,0}
+ϵ2m2(4m2t)2(4m2st)F0,0,0,0,2,1,2,0,0+ϵ3m2(8m44m2s2m2t+s2+st)4m2stF0,1,1,0,1,1,2,0,0\displaystyle\,+\frac{\epsilon^{2}m^{2}\left(4m^{2}-t\right)}{2\left(4m^{2}-s-t\right)}F_{0,0,0,0,2,1,2,0,0}+\frac{\epsilon^{3}m^{2}\left(8m^{4}-4m^{2}s-2m^{2}t+s^{2}+st\right)}{4m^{2}-s-t}F_{0,1,1,0,1,1,2,0,0}
+ϵ3(8m44m2s2m2t+s2+st)4(4m2st)F0,1,1,0,1,2,0,0,0\displaystyle\,+\frac{\epsilon^{3}\left(8m^{4}-4m^{2}s-2m^{2}t+s^{2}+st\right)}{4\left(4m^{2}-s-t\right)}F_{0,1,1,0,1,2,0,0,0}
+ϵ2(8m44m2s2m2t+s2+st)8(4m2st)F0,2,1,0,0,2,0,0,0\displaystyle\,+\frac{\epsilon^{2}\left(8m^{4}-4m^{2}s-2m^{2}t+s^{2}+st\right)}{8\left(4m^{2}-s-t\right)}F_{0,2,1,0,0,2,0,0,0}
+ϵ2(8m44m2s2m2t+s2+st)4(4m2st)F0,2,2,0,0,1,0,0,0\displaystyle\,+\frac{\epsilon^{2}\left(8m^{4}-4m^{2}s-2m^{2}t+s^{2}+st\right)}{4\left(4m^{2}-s-t\right)}F_{0,2,2,0,0,1,0,0,0}
+12ϵ3(8m42m2s2m2t+st)F0,1,1,0,1,2,1,0,0,\displaystyle\,+\frac{1}{2}\epsilon^{3}\left(8m^{4}-2m^{2}s-2m^{2}t+st\right)F_{0,1,1,0,1,2,1,0,0}\,,
f19\displaystyle f_{19} =ϵ2sF2,0,2,1,0,0,0,0,0,\displaystyle\,=-\epsilon^{2}sF_{2,0,2,1,0,0,0,0,0}\,,
f20\displaystyle f_{20} =ϵ2ss4m2sF2,1,2,1,0,0,0,0,0,\displaystyle\,=\epsilon^{2}\sqrt{-s}s\sqrt{4m^{2}-s}F_{2,1,2,1,0,0,0,0,0}\,,
f21\displaystyle f_{21} =ϵ2s4m2sF2,1,0,0,0,2,0,0,0,\displaystyle\,=-\epsilon^{2}\sqrt{-s}\sqrt{4m^{2}-s}F_{2,1,0,0,0,2,0,0,0}\,,
f22\displaystyle f_{22} =12ϵ3sF0,0,1,1,2,1,0,0,0ϵ2sF0,0,2,1,1,2,0,1,0,\displaystyle\,=-\frac{1}{2}\epsilon^{3}sF_{0,0,1,1,2,1,0,0,0}-\epsilon^{2}sF_{0,0,2,1,1,2,0,-1,0}\,,
f23\displaystyle f_{23} =ϵ3s4m2sF0,0,1,1,2,1,0,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,0,1,1,2,1,0,0,0}\,,
f24\displaystyle f_{24} =ϵ3s4m2sF0,0,1,1,1,2,0,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,0,1,1,1,2,0,0,0}\,,
f25\displaystyle f_{25} =ϵ3s4m2sF2,0,1,1,0,0,1,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{2,0,1,1,0,0,1,0,0}\,,
f26\displaystyle f_{26} =ϵ3sF1,1,0,0,0,2,1,0,0ϵ2sF2,1,0,0,0,2,1,0,1,\displaystyle\,=\epsilon^{3}sF_{1,1,0,0,0,2,1,0,0}-\epsilon^{2}sF_{2,1,0,0,0,2,1,0,-1}\,,
f27\displaystyle f_{27} =ϵ3s4m2sF1,1,0,0,0,2,1,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{1,1,0,0,0,2,1,0,0}\,,
f28\displaystyle f_{28} =ϵ3sF1,1,1,1,0,1,0,0,02sϵ4F1,1,1,1,0,1,0,0,0,\displaystyle\,=\epsilon^{3}sF_{1,1,1,1,0,1,0,0,0}-2s\epsilon^{4}F_{1,1,1,1,0,1,0,0,0}\,,
f29\displaystyle f_{29} =ϵ4s4m2sF1,0,1,1,1,1,0,0,0,\displaystyle\,=-\epsilon^{4}\sqrt{-s}\sqrt{4m^{2}-s}F_{1,0,1,1,1,1,0,0,0}\,,
f30\displaystyle f_{30} =ϵ4s2F1,1,1,1,1,1,0,0,04ϵ4m2sF1,1,1,1,1,1,0,0,0,\displaystyle\,=\epsilon^{4}s^{2}F_{1,1,1,1,1,1,0,0,0}-4\epsilon^{4}m^{2}sF_{1,1,1,1,1,1,0,0,0}\,,
f31\displaystyle f_{31} =ϵ3s2F2,1,1,1,0,0,1,0,04ϵ3m2sF2,1,1,1,0,0,1,0,0,\displaystyle\,=\epsilon^{3}s^{2}F_{2,1,1,1,0,0,1,0,0}-4\epsilon^{3}m^{2}sF_{2,1,1,1,0,0,1,0,0}\,,
f32\displaystyle f_{32} =ϵ4ss4m2sF1,1,1,1,0,1,1,0,0\displaystyle\,=\epsilon^{4}\sqrt{-s}s\sqrt{4m^{2}-s}F_{1,1,1,1,0,1,1,0,0}\,
f33\displaystyle f_{33} =ϵ3s4m2sF1,1,0,0,1,2,1,0,1ϵ3st4m2sF1,1,0,0,1,2,1,0,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{1,1,0,0,1,2,1,0,-1}-\epsilon^{3}\sqrt{-s}t\sqrt{4m^{2}-s}F_{1,1,0,0,1,2,1,0,0}\,,
f34\displaystyle f_{34} =ϵ3st4m2s4m2tF1,1,0,0,1,2,1,0,0,\displaystyle\,=\epsilon^{3}\sqrt{-s}\sqrt{-t}\sqrt{4m^{2}-s}\sqrt{4m^{2}-t}F_{1,1,0,0,1,2,1,0,0}\,,
f35\displaystyle f_{35} =ϵ3s4m2sF0,0,1,1,1,2,1,1,0,\displaystyle\,=-\epsilon^{3}\sqrt{-s}\sqrt{4m^{2}-s}F_{0,0,1,1,1,2,1,-1,0}\,,
f36\displaystyle f_{36} =ϵ3st4m2tF0,0,1,1,1,2,1,0,0+ϵ3st4m2tF0,0,1,1,2,1,1,0,0,\displaystyle\,=\epsilon^{3}s\sqrt{-t}\sqrt{4m^{2}-t}F_{0,0,1,1,1,2,1,0,0}+\epsilon^{3}s\sqrt{-t}\sqrt{4m^{2}-t}F_{0,0,1,1,2,1,1,0,0}\,,
f37\displaystyle f_{37} =ϵ3s4m6s(m2t)2F0,0,1,1,1,2,1,0,0,\displaystyle\,=\epsilon^{3}\sqrt{-s}\sqrt{4m^{6}-s\left(m^{2}-t\right)^{2}}F_{0,0,1,1,1,2,1,0,0}\,,
f38\displaystyle f_{38} =ϵ4st4m2tF1,0,1,1,1,1,1,0,0,\displaystyle\,=\epsilon^{4}s\sqrt{-t}\sqrt{4m^{2}-t}F_{1,0,1,1,1,1,1,0,0}\,,
f39\displaystyle f_{39} =ϵ4s4m2s(m2F1,0,1,1,1,1,1,0,0F0,1,1,0,1,1,1,0,0+F1,0,1,1,1,1,1,1,0),\displaystyle\,=-\epsilon^{4}\sqrt{-s}\sqrt{4m^{2}-s}\left(m^{2}F_{1,0,1,1,1,1,1,0,0}-F_{0,1,1,0,1,1,1,0,0}+F_{1,0,1,1,1,1,1,-1,0}\right)\,,
f40\displaystyle f_{40} =ϵ4sst4m2s4m2tF1,1,1,1,1,1,1,0,0,\displaystyle\,=-\epsilon^{4}s\sqrt{-s}\sqrt{-t}\sqrt{4m^{2}-s}\sqrt{4m^{2}-t}F_{1,1,1,1,1,1,1,0,0}\,,
f41\displaystyle f_{41} =ϵ4ss4m2s(F1,1,1,1,1,1,1,0,1+tF1,1,1,1,1,1,1,0,0),\displaystyle\,=-\epsilon^{4}s\sqrt{-s}\sqrt{4m^{2}-s}(F_{1,1,1,1,1,1,1,0,-1}+tF_{1,1,1,1,1,1,1,0,0})\,,
f42\displaystyle f_{42} =ϵ4(4m4sF1,1,1,1,1,1,1,0,0m2s2F1,1,1,1,1,1,1,0,0\displaystyle\,=\epsilon^{4}\left(4m^{4}sF_{1,1,1,1,1,1,1,0,0}-m^{2}s^{2}F_{1,1,1,1,1,1,1,0,0}\right.
+4m2sF1,1,1,1,1,1,1,1,0s2F1,1,1,1,1,1,1,1,0),\displaystyle\,\left.+4m^{2}sF_{1,1,1,1,1,1,1,-1,0}-s^{2}F_{1,1,1,1,1,1,1,-1,0}\right)\,,
f43\displaystyle f_{43} =ϵ4m2sF1,0,1,1,1,1,1,0,0ϵ4m2sF1,1,1,1,1,1,1,0,1ϵ3m2sF0,1,1,0,1,1,2,0,0\displaystyle\,=\epsilon^{4}m^{2}sF_{1,0,1,1,1,1,1,0,0}-\epsilon^{4}m^{2}sF_{1,1,1,1,1,1,1,0,-1}-\epsilon^{3}m^{2}sF_{0,1,1,0,1,1,2,0,0}
ϵ3m2sF0,1,1,0,1,2,1,0,0+ϵ412s2tF1,1,1,1,1,1,1,0,0ϵ412s2F1,1,1,1,0,1,1,0,0\displaystyle\,-\epsilon^{3}m^{2}sF_{0,1,1,0,1,2,1,0,0}+\epsilon^{4}\frac{1}{2}s^{2}tF_{1,1,1,1,1,1,1,0,0}-\epsilon^{4}\frac{1}{2}s^{2}F_{1,1,1,1,0,1,1,0,0}
+ϵ412s2F1,1,1,1,1,1,1,0,1+12ϵ3stF0,1,1,0,1,2,1,0,0ϵ3stF1,1,0,0,1,2,1,0,0,\displaystyle\,+\epsilon^{4}\frac{1}{2}s^{2}F_{1,1,1,1,1,1,1,0,-1}+\frac{1}{2}\epsilon^{3}stF_{0,1,1,0,1,2,1,0,0}-\epsilon^{3}stF_{1,1,0,0,1,2,1,0,0}\,,
+ϵ3stF2,1,1,1,0,0,1,0,0+ϵ4sF0,1,1,0,1,1,1,0,0,ϵ4sF1,0,1,1,1,1,0,0,0+ϵ4sF1,0,1,1,1,1,1,1,0\displaystyle\,+\epsilon^{3}stF_{2,1,1,1,0,0,1,0,0}+\epsilon^{4}sF_{0,1,1,0,1,1,1,0,0},-\epsilon^{4}sF_{1,0,1,1,1,1,0,0,0}+\epsilon^{4}sF_{1,0,1,1,1,1,1,-1,0}
ϵ4sF1,1,1,1,1,1,1,1,112ϵ3sF0,1,1,0,0,2,1,0,0+14ϵ3sF0,1,1,0,1,2,0,0,0+ϵ3sF1,1,0,0,0,2,1,0,0\displaystyle\,-\epsilon^{4}sF_{1,1,1,1,1,1,1,-1,-1}-\frac{1}{2}\epsilon^{3}sF_{0,1,1,0,0,2,1,0,0}+\frac{1}{4}\epsilon^{3}sF_{0,1,1,0,1,2,0,0,0}+\epsilon^{3}sF_{1,1,0,0,0,2,1,0,0}
ϵ3sF1,1,0,0,1,2,1,0,1+12ϵ3sF1,1,1,1,0,1,0,0,0ϵ218sF0,2,1,0,0,2,0,0,0ϵ214sF0,2,2,0,0,1,0,0,0.\displaystyle\,-\epsilon^{3}sF_{1,1,0,0,1,2,1,0,-1}+\frac{1}{2}\epsilon^{3}sF_{1,1,1,1,0,1,0,0,0}-\epsilon^{2}\frac{1}{8}sF_{0,2,1,0,0,2,0,0,0}-\epsilon^{2}\frac{1}{4}sF_{0,2,2,0,0,1,0,0,0}\,.

Appendix B Elliptic multiple polylogarithms

In this appendix we collect some technical material related to eMPLs not reviewed in section 3.2.

In the case where all the indices Ai=(nici)A_{i}=\left(\begin{smallmatrix}n_{i}\\ c_{i}\end{smallmatrix}\right) are equal to (±10)\left(\begin{smallmatrix}\pm 1\\ 0\end{smallmatrix}\right), the integral in eq. (21) is divergent and requires a special treatment and we define instead,

4(A1Ak;x;a)=1k!logkx+\displaystyle\cal E_{4}(A_{1}\ldots A_{k};x;\vec{a})=\frac{1}{k!}\log^{k}x+ l=0km=1lσ(1)l+m(kl)!logklx\displaystyle\sum_{l=0}^{k}\sum_{m=1}^{l}\sum_{\sigma}\frac{(-1)^{l+m}}{(k-l)!}\log^{k-l}x (59)
×4R(Aσ(1)(m)Aσ(m1)(m)Aσ(m+1)(m)Aσ(l)(m)|Am;x;a),\displaystyle\times\cal E_{4}^{\textrm{R}}\left(A^{(m)}_{\sigma(1)}\ldots A^{(m)}_{\sigma(m-1)}A^{(m)}_{\sigma(m+1)}\ldots A^{(m)}_{\sigma(l)}\Big{|}A_{m};x;\vec{a}\right)\,,

with Ai(m)=AiA_{i}^{(m)}=A_{i} if i<mi<m and Ai(m)=(10)A_{i}^{(m)}=\left(\begin{smallmatrix}1\\ 0\end{smallmatrix}\right) otherwise. The third sum runs over all shuffles σ\sigma of (A1(m)Am1(m))\left(A^{(m)}_{1}\ldots A^{(m)}_{m-1}\right) and (Am+1(m)Al(m))\left(A^{(m)}_{m+1}\ldots A^{(m)}_{l}\right), i.e., over all permutations of their union that preserve the relative ordering within each of the two lists. The 4R\cal E_{4}^{\textrm{R}} are iterated integrals with suitable subtractions to render the integrations finite,

4R(n1nk00|na0;x;a)=0x𝑑t1Ψn1(0,t1)0t10tk1𝑑tk(Ψna(0,tk)Ψ1(0,tk)).\!\!\!\cal E_{4}^{\textrm{R}}\left(\begin{smallmatrix}n_{1}&\ldots&n_{k}\\ 0&\ldots&0\end{smallmatrix}|\begin{smallmatrix}n_{a}\\ 0\end{smallmatrix};x;\vec{a}\right)=\int_{0}^{x}dt_{1}\Psi_{n_{1}}(0,t_{1})\int_{0}^{t_{1}}\ldots\int_{0}^{t_{k-1}}dt_{k}\left(\Psi_{n_{a}}(0,t_{k})-\Psi_{1}(0,t_{k})\right)\,. (60)

The rather ad-hoc looking form of eq. (59) is determined essentially uniquely by requiring the regularised eMPLs to share the same algebraic and differential properties as their convergent analogues. We refer to ref. Broedel:2018qkq for a detailed discussion.

The functions Z4(x,a)Z_{4}(x,\vec{a}) and G(a)G_{\ast}(\vec{a}) that appear in eq. (25) can be expressed in terms of incomplete elliptic integrals of the first and second kind Broedel:2017kkb ; Broedel:2018qkq :

F(x|λ)=0xdt(1t2)(1λt2),E(x|λ)=0x𝑑t1λt21t2.\begin{split}\textrm{F}(x|\lambda)&\,=\int_{0}^{x}\frac{dt}{\sqrt{(1-t^{2})(1-\lambda t^{2})}}\,,\\ \textrm{E}(x|\lambda)&\,=\int_{0}^{x}dt\,\sqrt{\frac{1-\lambda t^{2}}{1-t^{2}}}\,.\end{split} (61)

References