1 Introduction

Nowadays the calculation of observables relevant for high-energy physics (HEP) colliders is extremely important, in view of the future experiments that will provide new data with accuracy not reached so far. Therefore, insight and support from the theory side is necessary to understand the new findings. Clearly, the HEP community has to be ready to tackle this kind of problems and various approaches have to be implemented or reformulated. In particular, the perturbative framework applied to Quantum Field Theories (QFTs) has shown to be very important for providing highly precise predictions. The so-called Next-to-Leading Order (NLO) revolution was possible due to the emergence of novel techniques inspired by clever ideas. Likewise, predictions at NNLO are currently being calculated, but a fully established framework as at NLO is not yet complete. There are indeed several ideas and already working methods that can, for some processes, produce NNLO predictions to be compared with available experimental measurements.

In the spirit of providing observables at NNLO and beyond, one encounters several obstacles that do not allow to easily perform an evaluation in the physical four space-time dimensions. For instance, the calculation of multi-loop Feynman integrals constitutes a challenge due to the presence of singularities. Hence, proper procedures to deal with infrared (IR) and ultraviolet (UV) divergences and with physical threshold singularities have to be devised and implemented. As seen in many applications, starting with an integrand free of singularities makes the evaluation more stable and leads to reliable numerical results. Such an integrand-level representation of physical observables, characterised by a point-by-point, or local cancellation of singularities, is one of the main topics of this manuscript and a valuable item in the HEP community wish-list.

This manuscript is one of the outcomes of the discussions and activities of the workshop “WorkStop/ThinkStart 3.0: paving the way to alternative NNLO strategies”, which took place on 4–6 November 2019 at the Galileo Galilei Institute for Theoretical Physics (GGI) in Florence. In this manuscript, we compare and summarise the several strategies, presented at the workshop, to locally cancel IR singularities and, thus, providing local integrand-level representations of physical observables in four space-time dimensions.Footnote 1 In order to analyse their features, we consider the NNLO correction to the scattering process \(e^{+}e^{-}\rightarrow \gamma ^{*}\rightarrow q{\bar{q}}\). The adopted techniques extend the ones summarised in Ref. [16], where thoughtful and complete descriptions of the NLO calculations for this process are provided. Although the full calculation of NNLO predictions requires several ingredients, we elucidate among the different frameworks their main features to perform such computations. Special emphasis is put on comparing the advantages and limitations of each strategy, in order to provide the reader with a better understanding of the techniques that are currently available.

It is clear that having a fully local representation of any physical observable allows for a smooth numerical evaluation and, thus, a more direct calculation of highly accurate predictions to be compared with the experimental data. In the context of this kind of calculations, taking care of the regularisation techniques applied to reach well-defined results is very important. Speaking in a wide sense, in this manuscript we consider two kinds of techniques, depending on the underlying dimensionality of the integration space: four vs. d dimensional implementations. Both alternatives are constrained to fulfil several conditions. In the following, we list a few of them.

  • A comparison between various versions of regularisation schemes that regulate singularities by treating the integration momenta in d dimensional space-time (dimensional schemes) and schemes that do not alter the space-time dimension (non-dimensional schemes) is carried out at NNLO, showing transition rules between both approaches at intermediate steps of the calculation.

  • The ultraviolet renormalisation, preserving all the symmetry properties of the amplitude, is under study in the different approaches. Some of these methods aim at an integrand-level renormalisation, which differs from the traditional integral-level framework. In other words, one is interested in extracting the usual UV counterterms directly from the bare amplitude rather than subtracting integrated counterterms. Once again, the main focus is put on the locality: subtracting the UV singularities directly from the amplitude leads to a local cancellation of non-integrable contributions in the UV region.

  • It is clear that multi-loop level calculations are, in general, contaminated not only by UV singularities. In fact, the presence of IR ones makes a direct integration more involved. This is because the standard IR singularities need to be canceled by the corresponding real corrections unless an IR counterterm is encountered.

A clear understanding of the aforementioned points will pave an avenue to provide a systematic procedure to generate NNLO calculations. On top of the d- or not to d-dimensional techniques, the treatment of \(\gamma _{5}\) might be elucidated to, thus, find agreement between the different schemes. Although this topic was not considered on of main the targets of this workshop, interesting discussions took place. In particular, at the closing discussion, which we summarise at the end of this manuscript.

Nevertheless, with the very interesting developments at the HEP colliders proposed for the near future, it is currently mandatory to consider higher-order predictions. Therefore, NNLO is no longer the ultimate goal and all methods need to overcome the obstacles of providing observables at \(\hbox {N}^3\)LO and \(\hbox {N}^4\)LO accuracy. Hence, for these reasons, presenting a collection of different methods has the intention of illustrating where we are and what we can do next. To this end, in the present manuscript, we comment on the features of the following regularisation/renormalisation schemes as well as methods that are only focused on the local cancellation of IR singularities:

  • Dimensional schemes: four dimensional helicity scheme (fdh) and dimensional reduction (dred)

  • Non-dimensional schemes: four-dimensional unsubtracted scheme (fdu), four dimensional regularisation/renormalisation (fdr), and implicit regularisation (ireg).

  • Subtraction methods: the \(q_{{\mathrm{T}}}\)-subtraction method, the antenna and the local analytic sector subtraction.

In the last section of this manuscript, we summarise the advantages and disadvantages of the above-mentioned methods. Furthermore, we provide a very brief summary of issues that were mentioned during the closing discussion session at the Workshop.

2 NNLO processes in FDH/DRED and FDR

The vast majority of higher-order calculations in QFT are done using conventional dimensional regularisation (cdr) to deal with ultraviolet (UV) and infrared (IR) singularities in intermediate expressions. As discussed in [\({\mathcal {O}}(\epsilon )\) terms in the real matrix element are compensated by similar modifications in the real-virtual corrections and adapted UV renormalisation, such that the physical cross section is scheme independent. This cancellation of the scheme dependence is best understood by treating the \(\epsilon \)-scalars that need to be introduced to consistently define fdh and dred as spurious physical particles. The UV and IR singularities of processes involving \(\epsilon \)-scalars cancel for physical processes after consistent UV renormalisation and combining double-virtual, real-virtual, and double-real parts. This leaves us with a finite contribution multiplied by \(n_\epsilon =2\epsilon \), the multiplicity of the \(\epsilon \)-scalars. Setting \(\epsilon \rightarrow 0\) in the final result the contribution of the \(\epsilon \)-scalars drops out or, equivalently, the scheme dependence cancels.

Since the contribution of \(\epsilon \)-scalars drops out for physical observables it is, of course, possible to leave them out from the very beginning. This is nothing but computing in cdr. However, including \(\epsilon \)-scalars sometimes offers advantages, as it is (from a technical point of view) equivalent to performing the algebra in four dimensions. We reiterate the statement that introducing \(\epsilon \)-scalars in diagrams and counterterms is nothing but a consistent procedure (also beyond leading order) to technically implement the often made instruction to “perform the algebra in four dimensions”.Footnote 2

Once we know how to transform from cdr to dimensional schemes where some degrees of freedom are kept in four dimensions, we ask the question whether the latter can be related to an entirely four-dimensional calculation using four-dimensional regularisation (fdr) [23,24,25].Footnote 3 It is clear that physical results must not depend on the regularisation scheme and, therefore, fdr (and fdh and dred) has to reproduce the results obtained in cdr, as long as the same renormalisation scheme (typically \({\overline{\textsc {ms}}}\)) is used. For that reason we investigate the possibility to relate individual parts of the calculation obtained in fdr to the corresponding ones in fdh and dred. Such a relation would deepen our understanding of the alternative schemes and tighten the argument that they are consistent (at least to the order investigated). On a more practical level, it opens up the possibility to compute individual parts in different schemes and consistently combine these results.

We start in Sect. 2.1 by presenting the new results for \(H\!\rightarrow \! b{\bar{b}}\) in dred and fdh and by discussing the results for \(\gamma \!\rightarrow \! q {{\bar{q}}}\). The corresponding NLO results and the \(N_F\) part of the NNLO results in fdr are presented in Sect. 2.2. In Sect. 2.3 we study the relation of the individual parts between the dimensional schemes and fdr.

2.1 FDH and DRED

The relation between cdr and other dimensional schemes such as fdh and dred has been investigated thoroughly in the literature [26, 27]. Before we consider the processes \(H\!\rightarrow \!b{\bar{b}}\) and \(\gamma ^*\!\rightarrow \!q{\bar{q}}\) we collect the renormalisation constants that are required. As is well known [28], in fdh and dred the evanescent coupling of an \(\epsilon \)-scalar to a fermion, \(a_e\!=\!\alpha _e/(4\pi )\), has to be distinguished from the gauge coupling \(a_s\!=\!\alpha _s/(4\pi )\). Identifying the renormalisation and regularisation scales, the relation between the bare couplings \(a_s^0\) and \(a_e^0\) to the \({\overline{\textsc {ms}}}\)-renormalised ones is given as

$$\begin{aligned}&a_{s}^{0} = Z_{a_s}^{{\overline{\textsc {ms}}}} \, a_{s}\nonumber \\&\quad = a_{s}\,S_{\epsilon }^{-1}\, \Big \{1- \frac{a_{s}}{\epsilon } \Big [ C_A\Big ( \frac{11}{3} -\frac{n_\epsilon }{6} \Big ) -\frac{2}{3}N_F \Big ] +{\mathcal {O}}(a^2) \Big \}, \end{aligned}$$
(2.1a)
$$\begin{aligned}&a_{e}^{0} = Z_{a_e}^{{\overline{\textsc {ms}}}} \, a_{e} = a_{e}\,S_{\epsilon }^{-1}\, \Big \{1- \frac{a_{s}}{\epsilon }\Big [ 6\,C_F \Big ] \nonumber \\&\quad -\frac{a_{e}}{\epsilon }\Big [ C_A(2\!-\!n_\epsilon ) -C_F(4\!-\!n_\epsilon ) -N_F \Big ] + {\mathcal {O}}(a^2) \Big \}, \end{aligned}$$
(2.1b)

with \(S_{\epsilon }\!=\!e^{-\epsilon \gamma _{E}}(4\pi )^{\epsilon }\). In cdr we only need (2.1a) with \(n_\epsilon \!\rightarrow \!0\), whereas in fdh and dred \(n_\epsilon \!=\!2\epsilon \). The corresponding values in the \({\overline{\textsc {dr}}}\) scheme are simply obtained by setting \(n_\epsilon \!=\!0\) in (2.1). As the \(N_F\) part does not depend on \(n_\epsilon \) at the one-loop level, it is the same both in \({\overline{\textsc {ms}}}\) and \({\overline{\textsc {dr}}}\). For later purpose we further need the difference between (2.1b) and (2.1a) which is given by

$$\begin{aligned} \delta _{Z}^{{\overline{\textsc {ms}}}}&\equiv S_{\epsilon } \Big [Z_{a_e}^{{\overline{\textsc {ms}}}}\!-\!Z_{a_s}^{{\overline{\textsc {ms}}}} \Big ]_{a_e=a_s} =\frac{a_s^2}{\epsilon }\left[ C_F\left( \!-\!2\!-\!n_\epsilon \right) \right. \nonumber \\&\quad \left. +C_A\left( \frac{5}{3}\!+\!\frac{5}{6}n_\epsilon \right) \!+\!\frac{1}{3}N_F\right] + {\mathcal {O}}(a^3). \end{aligned}$$
(2.2)

For \(H\!\rightarrow \!b {\bar{b}}\) we also need the renormalisation of the Yukawa coupling \(y_{b}^{0} \) which is defined as the ratio of the bottom quark mass and the Higgs vacuum expectation value, see e.g. [29, 30]. Apart from its appearance through the Yukawa coupling we will set the bottom quark mass to zero. While the renormalisation of \(y_{b}^{0}\) is well known in cdr, for fdh/dred this constitutes an additional calculational step. Using the \({\overline{\textsc {ms}}}\) scheme we get

$$\begin{aligned} y_{b}^{0}&= y_{b} \left[ 1 + a_s\, S_{\epsilon }^{-1}\, C_F\,\left( \!-\!\frac{3}{\epsilon } \!-\!\frac{n_\epsilon }{2\,\epsilon }\right) \right. \nonumber \\&\quad + a^2_s\, S_{\epsilon }^{-2}\,C^2_F\,\left( \frac{9}{2\,\epsilon ^2} - \frac{3}{4\,\epsilon } + \frac{2\,n_\epsilon }{\epsilon ^2} - \frac{n_\epsilon }{\epsilon } + \frac{3\,n_\epsilon ^2}{8\,\epsilon ^2} - \frac{9\,n_\epsilon ^2}{16\,\epsilon } \right) \nonumber \\&\quad + a^2_s\, S_{\epsilon }^{-2}\,C_A C_F\,\left( \frac{11}{2\,\epsilon ^2} - \frac{97}{12\,\epsilon } + \frac{n_\epsilon }{4\,\epsilon ^2} - \frac{19\,n_\epsilon }{24\,\epsilon } - \frac{n_\epsilon ^2}{4\,\epsilon ^2} + \frac{n_\epsilon ^2}{4\,\epsilon } \right) \nonumber \\&\quad \left. + a^2_s\, S_{\epsilon }^{-2}\, C_F N_F\,\left( - \frac{1}{\epsilon ^2} + \frac{5}{6\,\epsilon } - \frac{n_\epsilon }{4\,\epsilon ^2} + \frac{n_\epsilon }{8\,\epsilon } \right) \right] . \end{aligned}$$
(2.3)

Similar to (2.2) we again have set equal the renormalised(!) couplings \(a_e\!=\!a_s\). The Yukawa renormalisation can be obtained from the UV divergences of an off-shell computation of the \(H\!\rightarrow \!b{\bar{b}}\) Green functions; however a technically simpler determination is also possible [31] by taking the on-shell form factor and subtracting the IR divergences obtained from the known general structure [26].

2.1.1 \(H\rightarrow b {{\bar{b}}}\) in FDH/DRED

In the following we consider one- and two-loop corrections to \(H\!\rightarrow \!b{{\bar{b}}}\) in fdh and dred. Our discussion follows closely [16] (for NLO) and [17] (for NNLO) where the corresponding calculations for the process \(e^+e^-\!\rightarrow \!\gamma ^*\!\rightarrow \! q{\bar{q}}\) are discussed. To start, we notice that disentangling the \(\epsilon \)-scalar contributions is actually simpler for \(H\!\rightarrow \!b{{\bar{b}}}\) than for \(\gamma ^*\!\rightarrow \!q {{\bar{q}}}\). This is due to the fact that the tree-level interaction is not mediated by a gauge boson and, accordingly, we therefore only have to split the gluon field in the one-loop contributions. Moreover, the major difference between fdh and dred is in the treatment of so-called ‘regular’ vector fields, see e.g. Table 1 of [16]. As in the present case only ‘singular’ vector fields contribute, the virtual correction is the same in both schemes, i.e.

$$\begin{aligned} {\mathcal {A}}_{{\textsc {fdh}}}(H\!\rightarrow \! b{\bar{b}})\ \equiv \ {\mathcal {A}}_{{\textsc {dred}}}(H\!\rightarrow \! b{\bar{b}})\ \equiv \ {\mathcal {A}}, \end{aligned}$$
(2.4)

see also Fig. 1. Using form factor coefficients \(F^{(n)}\) of loop-order n, the bare amplitude for \(H\!\rightarrow \! b {\bar{b}}\) can be written as

$$\begin{aligned} {\mathcal {A}}_{\text {bare}}&= {\mathcal {A}}^{(0)}_{\text {bare}} \left[ 1 + \sum _n \left( \frac{\mu _0^2}{\!-\!M_H^2}\right) ^{n\,\epsilon }\, S_{\epsilon }^{n}\,F^{(n)}_{\text {bare}} \right] , \end{aligned}$$
(2.5)

including the regularisation scale \(\mu _0\) and the Higgs mass \(M_H\). The tree-level amplitude \({\mathcal {A}}^{(0)}_{\text {bare}}\!=\! i\,y_{b}^{0}\, {\bar{u}}(p_b)v(p_{{\bar{b}}})\) results in the tree-level decay width \(\Gamma ^{(0)}\!=\!2\,M_H^2 y_b^2/(2 M_H)\).

Fig. 1
figure 1

The one-loop vertex correction to the \(H\!\rightarrow \!b{{\bar{b}}}\) amplitude. The diagram only contains a ‘singular’ gluon field. In dred and fdh there is an additional diagram with the gluon replaced by an \(\epsilon \)-scalar

2.1.2 NLO

The NLO virtual corrections are directly related the one-loop form factor for which we get in fdh/dred

$$\begin{aligned} F^{(1)}_{\text {bare}}&= a_{s}^{0}\,C_F\, \Big [ \!-\! \frac{2}{\epsilon ^2} \!-\! 2 \!+\! \frac{\pi ^2}{6} \!+\! {\mathcal {O}}(\epsilon ) \Big ] \nonumber \\&\quad + a_{e}^{0}\,C_F\,n_\epsilon \, \Big [ \frac{1}{\epsilon } \!+\! 2 \!+\! {\mathcal {O}}(\epsilon ) \Big ] . \end{aligned}$$
(2.6)

The well-known cdr result which is given e.g. in [29, 30] is obtained by setting \(n_\epsilon \!\rightarrow \!0\). Identifying the (bare) couplings in (2.6) before making replacement (2.1) corresponds to so-called ‘naive’ fdh and it has been shown that this leads to inconsistent results at higher orders [32,33,34]. Instead, we first have to renormalise \(a_{s}^{0}\) and \(a_{e}^{0}\) and only then we are allowed to identify the renormalised couplings \(a_e\!=\!a_s\).

Integrating over the phase space and using the conventions of [16] we then obtain for the UV-renormalised virtual cross section

$$\begin{aligned} \Gamma ^{(v)}_{{\textsc {ds}} }&= \Gamma ^{(0)}\,{C_F}\, \Phi _2(\epsilon )\,c_\Gamma (\epsilon )\, s^{-\epsilon } \Big \{a_s \Big [ \!-\!\frac{4}{\epsilon ^2} \!-\!\frac{6}{\epsilon } \!-\!4 \!+\!2\,\pi ^2 \!+\!{\mathcal {O}}(\epsilon ) \Big ] \nonumber \\&\quad + a_e \Big [ \!+\!\frac{n_\epsilon }{\epsilon } \!+\!{\mathcal {O}}(\epsilon ^0) \Big ] \Big \}, \end{aligned}$$
(2.7)

where we introduce the \(\epsilon \)-dependent prefactors

$$\begin{aligned} c_\Gamma (\epsilon )&=(4\pi )^{\epsilon } \frac{\Gamma (1+\epsilon )\,\Gamma ^2(1-\epsilon )}{\Gamma (1-2\epsilon )} =1+{\mathcal {O}}(\epsilon ), \end{aligned}$$
(2.8a)
$$\begin{aligned} \Phi _2(\epsilon )&=\Big (\frac{4\pi }{s}\Big )^\epsilon \frac{\Gamma (1-\epsilon )}{\Gamma (2-2\epsilon )} =1+{{{\mathcal {O}}}}(\epsilon ), \end{aligned}$$
(2.8b)
$$\begin{aligned} \Phi _3(\epsilon )&=\Big (\frac{4\pi }{s}\Big )^{2\epsilon }\frac{1}{\Gamma (2-2\epsilon )} = 1 + {{{\mathcal {O}}}}(\epsilon ). \end{aligned}$$
(2.8c)

The subscript ds in (2.7) and in what follows indicates that the results for all dimensional schemes can be obtained from this expression. For the evaluation of the real contribution we use the setup and notation of [16] and arrive at

$$\begin{aligned} \Gamma ^{(r)}_{{\textsc {ds}}}&= \Gamma ^{(0)}\,{C_F}\,\Phi _3(\epsilon )\, \Big \{a_s\Big [ \!+\!\frac{4}{\epsilon ^2} \!+\!\frac{6}{\epsilon } \!+\!21 \!-\!2\,\pi ^2 \!+\!{\mathcal {O}}(\epsilon ) \Big ] \nonumber \\&\quad + a_e\Big [ \!-\!\frac{n_\epsilon }{\epsilon } \!+\!{\mathcal {O}}(\epsilon ^0) \Big ] \Big \}. \end{aligned}$$
(2.9)

As for the virtual cross section, the result is the same in fdh and dred which is due to the absence of ‘regular’ gauge bosons at NLO for the considered process. The presence of a ‘singular’ gluon, however, leads to \(a_s\) contributions (which stem from the \({d}\)-dimensional gluon) and \(a_e\) contributions (which stem from the associated \(\epsilon \)-scalar gluon). Again, at NLO such a distinction is of course not strictly necessary; at NNLO, however, the different renormalisation of \(a_s\) and \(a_e\) has to be taken into account, see also Sect. 2.3.

Finally, since \(c_\Gamma \,\Phi _2 s^{-\epsilon }/\Phi _3 =1+\mathcal{O}(\epsilon ^3)\), the cancellation of singularities between (2.7) and (2.9) takes place and leaves us with the well known finite answer

$$\begin{aligned} \Gamma ^{(1)} =\Gamma ^{(0)} + \Gamma ^{(v)}_{{\textsc {ds}}} + \Gamma ^{(r)}_{{\textsc {ds}}}\Big |_{ {d}\rightarrow 4 } =\Gamma ^{(0)}\Big [ 1 + a_s\,17\,{C_F}\Big ]. \end{aligned}$$
(2.10)

In particular, the evanescent contributions \(\propto a_en_\epsilon \) drop out.

2.1.3 NNLO

In the following we provide separately the double-virtual, double-real, and real-virtual contributions to \(\Gamma \) in fdh/dred. All of these results are new and have not been published elsewhere before. For brevity reasons we keep the dependence on \(n_\epsilon \) explicit in the divergent terms, but drop finite \(n_\epsilon \) terms. Setting \(n_\epsilon \!\rightarrow \!0\ (2\epsilon ) \) then yields the cdr (fdh/dred) result. As before, we give the UV-renormalised results after having set \(a_e\!\equiv \!a_s\).

Writing the double virtual corrections as

$$\begin{aligned} \Gamma ^{(vv)}_{{\textsc {ds}}}&= \Gamma ^{(0)}\,\Phi _2(\epsilon )\,a^2_s\,{C_F}\,\Big [ C_F\, \Gamma ^{(vv)}_{{\textsc {ds}}}(C_F) \nonumber \\&\quad + C_A\, \Gamma ^{(vv)}_{{\textsc {ds}}}(C_A) + N_F\, \Gamma ^{(vv)}_{{\textsc {ds}}}(N_F) \Big ], \end{aligned}$$
(2.11)

we then obtain for the individual parts

$$\begin{aligned} \Gamma ^{(vv)}_{{\textsc {ds}}}(C_F)&= + \frac{8}{\epsilon ^4} + \frac{24-4n_\epsilon }{\epsilon ^3} + \frac{1}{\epsilon ^2} \Big ( 34-\frac{28\pi ^2}{3}-23n_\epsilon \Big )\nonumber \\&\quad + \frac{1}{\epsilon } \Big ( \frac{109}{2} \!-\!12\pi ^2 \!-\!\frac{184\zeta _3}{3} -62 n_\epsilon \!+\!\frac{20\pi ^2 n_\epsilon }{3} \!+\! \frac{31n_\epsilon ^2}{8} \Big )\nonumber \\&\quad +128 \!-\! \frac{40\pi ^2}{3} \!+\! \frac{137\pi ^4}{45} \!-\! 116\zeta _3, \end{aligned}$$
(2.12a)
$$\begin{aligned} \Gamma ^{(vv)}_{{\textsc {ds}}}(C_A)&= + \frac{22-n_\epsilon }{2\,\epsilon ^3} + \frac{1}{\epsilon ^2} \Big ( \frac{32}{9} \!+\!\frac{\pi ^2}{3} \!-\!\frac{19n_\epsilon }{18} \!+\!\frac{n_\epsilon ^2}{2} \Big ) \nonumber \\&\quad +\frac{1}{\epsilon } \Big ( \!-\!\frac{961}{54} \!-\!\frac{11\pi ^2}{6} \!+\!26\zeta _3 \!+\!\frac{761n_\epsilon }{108} \!+\!\frac{ \pi ^2 n_\epsilon }{12} \Big )\nonumber \\&\quad -\frac{934}{81} \!+\!\frac{701\pi ^2}{54} \!-\!\frac{8\pi ^4}{45} \!+\!\frac{302\zeta _3}{9}, \end{aligned}$$
(2.12b)
$$\begin{aligned} \Gamma ^{(vv)}_{{\textsc {ds}}}(N_F)&= - \frac{2}{\epsilon ^3} + \frac{1}{\epsilon ^2} \Big ( \!-\! \frac{8}{9} \!+\!\frac{n_\epsilon }{2} \Big )+ \frac{1}{\epsilon } \Big ( \frac{65}{27} \!+\!\frac{\pi ^2}{3} \!-\!\frac{3n_\epsilon }{4} \Big )\nonumber \\&\quad + \frac{400}{81} \!-\! \frac{55\pi ^2}{27} \!+\! \frac{4\zeta _3}{9}. \end{aligned}$$
(2.12c)

We note that in (2.11) we could have pulled out additional prefactors such as \(c_\Gamma ^2\), in analogy to (2.7). This would of course modify the subleading poles and finite terms of (2.12) (as well as (2.13) and (2.14) below). The conclusions, however, we will draw in Sect. 2.3 are not affected by the choice of the prefactor.

As mentioned before, the one- and two-loop renormalisation of the Yukawa coupling is included in (2.12) in order to get consistent results. We would like to stress, that in fdh and dred there are finite terms associated with the \({\overline{\textsc {ms}}}\) renormalisation factors. The terms \(\propto n_\epsilon ^m/\epsilon ^n\) (\(m,n\!>\!0\)) that are potentially finite when setting \(n_\epsilon \!=\!2\epsilon \) cancel when combining the double-virtual with the real-virtual and double-real contribution, as will be shown below. However, if the double-real (and real-virtual) corrections are computed by doing the algebra in four dimensions, the \(n_\epsilon \) terms are not disentangled any longer and, therefore, all these terms need to be included to obtain a consistent result. Moreover, as no \(\epsilon \)-scalars are present at the tree level, (2.1) is sufficient for the renormalisation of \(a_s^0\) and \(a_e^0\).

Splitting the real-virtual and double-real contributions in a similar way as before, we then get

$$\begin{aligned} \Gamma ^{(rv)}_{{\textsc {ds}}}(C_F)&= - \frac{16}{\epsilon ^4} - \frac{48\!-\!8n_\epsilon }{\epsilon ^3} \nonumber \\&\quad + \frac{1}{\epsilon ^2} \Big ( \!-\!146 \! +\!\frac{64\pi ^2}{3} \!+\!42n_\epsilon \!-\!\frac{n_\epsilon ^2}{2} \Big ) \nonumber \\&\quad +\frac{1}{\epsilon } \Big ( \!-\!524 \!+\!46\pi ^2 \!+\!\frac{848\zeta _3}{3} \!+\!147n_\epsilon \nonumber \\&\quad -\frac{37\pi ^2n_\epsilon }{3} \!-\!\frac{13n_\epsilon ^2}{2} \Big ) \nonumber \\&\quad -1879 \!+\!170\pi ^2 \!-\!\frac{38\pi ^4}{9} \!+\!624\zeta _3, \end{aligned}$$
(2.13a)
$$\begin{aligned} \Gamma ^{(rv)}_{{\textsc {ds}}}(C_A)&= -\frac{2}{\epsilon ^4} -\frac{62\!-\!5n_\epsilon }{3\,\epsilon ^3}\nonumber \\&\quad +\frac{1}{\epsilon ^2} \Big ( \!-\!52 \!+\!\frac{7\pi ^2}{3} \!+\!n_\epsilon \!-\!\frac{n_\epsilon ^2}{2} \Big ) \nonumber \\&\quad + \frac{1}{\epsilon } \Big ( \!-\!209 \!+\!\frac{158\pi ^2}{9} \!+\!\frac{16\zeta _3}{3} \!+\!\frac{25n_\epsilon }{2} \!-\!\frac{17\pi ^2n_\epsilon }{9} \Big )\nonumber \\&\quad - \frac{4769}{6} \!+\! \frac{355\pi ^2}{6} \!-\! \frac{47\pi ^4}{36} \!+\! \frac{2000\zeta _3}{9}, \end{aligned}$$
(2.13b)
$$\begin{aligned} \Gamma ^{(rv)}_{{\textsc {ds}}}(N_F)&= + \frac{8}{3\,\epsilon ^3} + \frac{1}{\epsilon ^2} \Big (4-n_\epsilon \Big ) + \frac{1}{\epsilon } \Big ( 14 \!-\!\frac{14\pi ^2}{9} \!-\!\frac{5n_\epsilon }{2} \Big )\nonumber \\&\quad + \frac{127}{3} \!-\! \frac{7\pi ^2}{3} \!-\! \frac{200\zeta _3}{9}\, \end{aligned}$$
(2.13c)

and

$$\begin{aligned} \Gamma ^{(rr)}_{{\textsc {ds}}}(C_F)&= + \frac{8}{\epsilon ^4} + \frac{24\!-\!4n_\epsilon }{\epsilon ^3} \nonumber \\&\quad + \frac{1}{\epsilon ^2} \Big ( 112 \!-\!12\pi ^2 \!-\!19n_\epsilon \!+\!\frac{n_\epsilon ^2}{2} \Big ) \nonumber \\&\quad +\frac{1}{\epsilon } \Big ( \frac{939}{2} \!-\! 34\pi ^2 \!-\!\frac{664\zeta _3}{3} \!-\!85n_\epsilon +\frac{17\pi ^2 n_\epsilon }{3} \!+\!\frac{21n_\epsilon ^2}{8} \Big ) \nonumber \\&\quad + \frac{7695}{4} \!-\! \frac{488\pi ^2}{3} \!+\! \frac{53\pi ^4}{45} \!-\! 544\zeta _3, \end{aligned}$$
(2.14a)
$$\begin{aligned} \Gamma ^{(rr)}_{{\textsc {ds}}}(C_A)&= + \frac{2}{\epsilon ^4} + \frac{58\!-\!7n_\epsilon }{6\epsilon ^3} + \frac{1}{\epsilon ^2} \Big ( \frac{436}{9} \!-\!\frac{8\pi ^2}{3} \!-\!\frac{89n_\epsilon }{18} \Big ) \nonumber \\&\quad + \frac{1}{\epsilon } \Big ( \frac{12247}{54} \!-\! \frac{283\pi ^2}{18} \!-\! \frac{94\zeta _3}{3} \!-\! \frac{2111n_\epsilon }{108} \!+\! \frac{65 \pi ^2 n_\epsilon }{36} \Big )\nonumber \\&\quad + \frac{333595}{324} \!-\! \frac{2047\pi ^2}{27} \!+\! \frac{89\pi ^4}{69} \!-\!\frac{2860\zeta _3}{9}, \end{aligned}$$
(2.14b)
$$\begin{aligned} \Gamma ^{(rr)}_{{\textsc {ds}}}(N_F)&= - \frac{2}{3\epsilon ^3} + \frac{1}{\epsilon ^2} \Big ( \!-\!\frac{28}{9} \!+\!\frac{n_\epsilon }{2} \Big )\nonumber \\&\quad + \frac{1}{\epsilon } \Big ( \!-\!\frac{443}{27} \!+\!\frac{11\pi ^2}{9} \!+\!\frac{13n_\epsilon }{4} \Big )\nonumber \\&\quad \!-\! \frac{12923}{162} \!+\! \frac{136\pi ^2}{27} \!+\! \frac{268\zeta _3}{9}. \end{aligned}$$
(2.14c)

To get the double-real contribution we used the integrals of [35] and the FORM code of [36]. As for the process \(\gamma ^*\!\rightarrow \!q{\bar{q}}\) [17], the double-real corrections in dred/fdh are simply obtained by integrating the four-dimensional matrix element squared over the phase space. Their determination is therefore significantly simplified compared to the case of cdr.

Finally, combining these results, we can extend (2.10) to NNLO asFootnote 4

$$\begin{aligned} \Gamma ^{(2)}&=\Gamma ^{(1)} + \Gamma ^{(vv)}_{{\textsc {ds}}} + \Gamma ^{(rv)}_{{\textsc {ds}}} + \Gamma ^{(rr)}_{{\textsc {ds}}} \Big |_{{d}\rightarrow 4} \end{aligned}$$
(2.15a)
$$\begin{aligned}&=\Gamma ^{(0)}\Big [ 1 + a_s\,17\,{C_F}+ a^2_s\, C^2_F \Big ( \frac{691}{4} \!-\!6\pi ^2 \!-\!36\zeta _3 \Big ) \nonumber \\&\quad + a^2_s\, C_F C_A \Big ( \frac{893}{4} \!-\!\frac{11\pi ^2}{3} \!-\!62\zeta _3 \Big ) \nonumber \\&\quad + a^2_s\, C_F N_F \Big ( \!-\!\frac{65}{2} \!+\!\frac{2\pi ^2}{3} \!+\! 8\zeta _3 \Big ) \Big ]\, \end{aligned}$$
(2.15b)

in agreement with [37]. As expected, the divergent \(n_\epsilon \) parts cancel in the final result which is nothing but the scheme-independence of the physical result.

2.1.4 \(\gamma \rightarrow q {{\bar{q}}}\) in FDH/DRED

The scheme dependence of the cross section \(e^+e^-\!\rightarrow \!\gamma ^*\!\rightarrow \!q{\bar{q}}\) at NNLO is discussed in detail in [17]. For this particular process the individual results in dred and fdh differ, as in dred there are \(\epsilon \)-scalar photons present in the tree-level process. While [17] mainly deals with dred, here we only recall a few points that are relevant for the comparison of fdh with fdr. In particular, we want to extend to NNLO the investigation of the interplay between fdh and fdr presented at NLO in [16].

2.1.5 NLO

Copying the results given in Section 2.3 of [16], we write the virtual and real cross sections as

$$\begin{aligned} \sigma ^{(v)}_\mathrm{FDH}&= \sigma ^{(0)}\,{C_F}\, \Phi _2(\epsilon )\,c_\Gamma (\epsilon )\, s^{-\epsilon } \nonumber \\&\quad \times \Big \{a_s \Big [ \!-\!\frac{4}{\epsilon ^2} \!-\!\frac{6}{\epsilon } \!-\!16 \!+\!2\pi ^2 \!+\!{\mathcal {O}}(\epsilon ) \Big ] \nonumber \\&\quad + a_e \Big [ \!+\!\frac{n_\epsilon }{\epsilon } \!+\!{\mathcal {O}}(\epsilon ^0) \Big ] \Big \}, \end{aligned}$$
(2.16)
$$\begin{aligned} \sigma ^{(r)}_\mathrm{FDH}&= \sigma ^{(0)}\,{C_F}\, \Phi _3(\epsilon )\,\nonumber \\&\quad \times \Big \{ a_s\Big [ \!+\!\frac{4}{\epsilon ^2} \!+\!\frac{6}{\epsilon } \!+\!19 \!-\!2\pi ^2 \!+\!{\mathcal {O}}(\epsilon ) \Big ] \nonumber \\&\quad + a_e\Big [ \!-\!\frac{n_\epsilon }{\epsilon } \!+\!{\mathcal {O}}(\epsilon ^0) \Big ] \Big \} , \end{aligned}$$
(2.17)

where \(\sigma ^{(0)}\!=\!e^4/(4\pi )\,Q_q N_c/(3s)\) includes the electric charge and the colour number of the quark as well as the flux factor 1/(2s). Combining these two contributions we find the well-known regularisation-scheme independent physical cross section

$$\begin{aligned} \sigma ^{(1)} =\sigma ^{(0)} + \sigma ^{(v)}_{{\textsc {fdh}}} + \sigma ^{(r)}_{{\textsc {fdh}}}\Big |_{{d}\rightarrow 4} =\sigma ^{(0)}\Big [ 1 + a_s\,3\,{C_F}\Big ]. \end{aligned}$$
(2.18)

2.1.6 NNLO

Moving on to NNLO, we first split the cross section into a double-virtual, a double-real, and a real-virtual part, i.e.

$$\begin{aligned} \sigma ^{{\textsc {nnlo}}}_{{\textsc {fdh}}}= \sigma ^{(vv)}_{{\textsc {fdh}}}+ \sigma ^{(rr)}_{{\textsc {fdh}}}+ \sigma ^{(rv)}_{{\textsc {fdh}}} . \end{aligned}$$
(2.19)

For the comparison with fdr in Sect. 2.3, we here focus on the \(N_F\) part of the respective contributions. The double-virtual part can be extracted from the dred result given in (3.8b) of [17] as

$$\begin{aligned} \sigma ^{(vv)}_{{\textsc {fdh}}}(N_F)&=\sigma ^{(0)}\,\Phi _2(\epsilon )\, a_{s}^2\,C_F N_F \Big [ \!-\!\frac{2}{\epsilon ^3} \nonumber \\&\quad -\frac{8}{9 \epsilon ^2} \!+\!\frac{1}{\epsilon } \Big (\frac{92}{27} \!+\!\frac{\pi ^2}{3}\Big ) \!+\!\frac{1921}{81} \!-\!\frac{91\pi ^2}{27} \!+\!\frac{4}{9}\zeta _3 \Big ]. \end{aligned}$$
(2.20)

The other two contributions are given by

$$\begin{aligned} \sigma ^{(rv)}_{{\textsc {fdh}}}(N_F)&=\sigma ^{(0)}\,\Phi _2(\epsilon )\, a_{s}^2\,C_F N_F \nonumber \\&\quad \times \Big [ \frac{8}{3\epsilon ^3} \!+\!\frac{4}{\epsilon ^2} \!+\!\frac{1}{\epsilon } \Big (\frac{32}{3} \!-\!\frac{14\pi ^2}{9}\Big ) \nonumber \\&\quad +\frac{94}{3} \!-\!\frac{7\pi ^2}{3} \!-\!\frac{200}{9}\zeta _3 \Big ], \end{aligned}$$
(2.21)
$$\begin{aligned} \sigma ^{(rr)}_{{\textsc {fdh}}}(N_F)&=\sigma ^{(0)}\,\Phi _2(\epsilon )\, a_{s}^2\,C_F N_F \nonumber \\&\quad \times \Big [ \!-\!\frac{2}{3\epsilon ^3} \!-\!\frac{28}{9 \epsilon ^2} \!+\!\frac{1}{\epsilon } \Big (\!-\!\frac{380}{27} \!+\!\frac{11\pi ^2}{9}\Big )\nonumber \\&\quad -\frac{5350}{81} \!+\!\frac{154\pi ^2}{27} \!+\!\frac{268}{9}\zeta _3 \Big ]. \end{aligned}$$
(2.22)

The combination of all three parts results in

$$\begin{aligned}&\sigma ^{(vv)}_{{\textsc {fdh}}}(N_F) + \sigma ^{(rv)}_{{\textsc {fdh}}}(N_F) + \sigma ^{(rr)}_{{\textsc {fdh}}}(N_F) \nonumber \\&\quad = \sigma ^{(0)}\,a_{s}^2\,C_F N_F \big [\!-\!11\!+\!8 \zeta _3 \big ] \end{aligned}$$
(2.23)

in agreement with the literature [38, 39]. Note that the constant terms of (2.21) and (2.22) differ from the corresponding results in dred, as given in (3.20) and (3.17b) of [17].

2.2 FDR: four-dimensional regularisation/renormalisation

2.2.1 \(H\rightarrow b {{\bar{b}}}\) in FDR

2.2.2 NLO

Here we describe the fdr NLO calculation of the decay rate \(\Gamma _{H\rightarrow b {{\bar{b}}}(g)}\). The strong coupling constant does not appear at the tree-level and as in Sect. 2.1 we use the \({\overline{\textsc {ms}}}\) value with \(a=\alpha _s/{4 \pi }\). As for the unrenormalised bottom mass, it is denoted by \(m_0\) and again it is taken to be different from zero only in the Yukawa coupling. The one-loop relation between \(m_0\) and the physical pole mass m, defined as the value of the four-momentum at which the bottom quark propagator develops a pole, is obtained by evaluating the diagram of Fig. 2,

$$\begin{aligned}&m_0 = m (1+a \delta m), \quad \delta m = - {C_F}(3 L^{\prime \prime }+5),\nonumber \\&L^{\prime \prime } :=\ln \mu ^2-\ln m^2. \end{aligned}$$
(2.24)
Fig. 2
figure 2

The one-loop correction to the bottom mass

The unphysical mass \(\mu ^2\) in (2.24) is the fdr UV regulator, which in the present calculation is taken to coincide with the fdr IR regulator.Footnote 5 Once the decay amplitude is renormalised (namely expressed in terms of physical quantities only) all the \(\mu ^2\)s of UV origin get replaced by physical scales, and the remaining ones are IR regulators which cancel in the sum of virtual and real contributions. As a matter of fact, we will encounter two additional combinations besides \(L^{\prime \prime }\),

$$\begin{aligned} L^\prime := \ln \mu ^2-\ln (-s-i 0^+)\quad {\mathrm{and}}\quad L := \ln \mu ^2-\ln s,\nonumber \\ \end{aligned}$$
(2.25)

where for \(H\rightarrow b {\bar{b}}\) we have \(s=M_H^2\), the Higgs mass squared. The amplitude contributing to \(H\rightarrow b {{\bar{b}}}\) can be written as

$$\begin{aligned} A_2 = m_0 A^{(0)}_2 (1+ a A_2^{(v)}), \end{aligned}$$
(2.26)

where \(A^{(0)}_2\) is the tree level result and the one-loop correction of Fig. 1 reads

$$\begin{aligned} A_2^{(v)} = -{C_F}(L^\prime )^2. \end{aligned}$$
(2.27)

Inserting (2.24) in (2.26) produces the renormalised one-loop amplitude

$$\begin{aligned} A_2^{(1)}= m A^{(0)}_2 \left[ 1-a {C_F}\big (5+3 L^{\prime \prime }+(L^\prime )^2 \big ) \right] . \end{aligned}$$
(2.28)

Upon integration over the 2-body phase-space, the square of (2.28) gives the virtual part of the NLO corrections,

$$\begin{aligned} \Gamma ^{(v)}_{\mathrm{FDR}}(m^2)= & {} -a {C_F}\Gamma ^{(0)}(m^2)\, \nonumber \\&\quad \times {{{\mathcal {R}}}}e \left[ 2 (L^\prime )^2+6 L^{\prime \prime }+10 \right] . \end{aligned}$$
(2.29)

This can be translated to the \({\overline{\textsc {ms}}}\) scheme by expressing \(m^2\) in terms of the running \({\overline{\textsc {ms}}}\) bottom mass \({\underline{m}}^2(M^2_H)\) [40]

$$\begin{aligned} m^2= & {} {\underline{m}}^2(M^2_H) \left[ 1+a{C_F}\big (8-6 \ln \frac{m^2}{M^2_H}\big ) \right] \nonumber \\= & {} {\underline{m}}^2(M^2_H) \left[ 1+a{C_F}\big (8-6 (L - L^{\prime \prime })\big ) \right] . \end{aligned}$$
(2.30)

With the corresponding change in the Yukawa we obtain

$$\begin{aligned} \Gamma ^{(v)}_{\mathrm{FDR}}= -a {C_F}\Gamma ^{(0)}\, {{{\mathcal {R}}}}e \left[ 2 (L^\prime )^2+6 L+2 \right] , \end{aligned}$$
(2.31)

which can be directly compared with (2.7). The real counterpart is obtained by squaring the diagrams of Fig. 3 and integrating over a 3-body phase-space in which all final-state particles acquire a small mass \(\mu \) [41],

$$\begin{aligned} \Gamma ^{(r)}_{\mathrm{FDR}}= & {} \Gamma ^{{\mathrm{R}}}_{H\rightarrow b {{\bar{b}}} g}\nonumber \\= & {} a {C_F}\Gamma ^{(0)}_{H\rightarrow b {{\bar{b}}}}(m^2) \left[ 2 L^2+6 L+19-2 \pi ^2 \right] . \end{aligned}$$
(2.32)

Replacing \({{{\mathcal {R}}}}e(L')^2 = L^2+\pi ^2\), the sum of (2.31) and (2.32) gives the UV and IR finite decay rate up to the NLO accuracy

$$\begin{aligned} \Gamma ^{(1)}_{\mathrm{FDR}}= \Gamma ^{(0)}\, \left[ 1+ 17 a {C_F}\right] , \end{aligned}$$
(2.33)

which agrees with (2.10).

Fig. 3
figure 3

The two diagrams contributing to the \(H\rightarrow b {{\bar{b}}} g\) amplitude. In dred and fdh there are additional diagrams with gluons replaced by \(\epsilon \)-scalars

2.2.3 NNLO, \(N_F\) part

The \(N_F\) parts contributing to the NNLO cross section in fdr are given in (4.19) of [25] in terms of the Yukawa coupling defined through the pole mass of the bottom quark. If we express these results in terms of the \({\overline{\textsc {ms}}}\) Yukawa couplings, as done for (2.31), we get

$$\begin{aligned} \Gamma ^{(vv)}_{{\textsc {fdr}}}(N_F)&=\Gamma ^{(0)}\, a_{s}^2\,C_F N_F \Big [ \!-\!\frac{8}{9} L^3 \!+\!\frac{2}{9} L^2 \nonumber \\&\quad + L \Big (\frac{278}{27}\!+\!\frac{4}{9}\pi ^2\Big ) \!+\!\frac{3425}{162} -\frac{40}{27}\pi ^2 \!-\!\frac{16}{3}\zeta _3 \Big ], \end{aligned}$$
(2.34)
$$\begin{aligned} \Gamma ^{(rv)}_{{\textsc {fdr}}}(N_F)&=\Gamma ^{(0)}\, a_{s}^2\,C_F N_F \Big [ \!+\!\frac{4}{3} L^3 \!+\! 4 L^2+ L \Big (\frac{38}{3}\!-\!\frac{4}{3}\pi ^2\Big ) \Big ], \end{aligned}$$
(2.35)
$$\begin{aligned} \Gamma ^{(rr)}_{{\textsc {fdr}}}(N_F)&=\Gamma ^{(0)}\, a_{s}^2\,C_F N_F \Big [ \!-\!\frac{4}{9} L^3 \!-\!\frac{38}{9} L^2- L \Big (\frac{620}{27}\!-\!\frac{8}{9}\pi ^2\Big )\nonumber \\&\quad -\frac{4345}{81} \!+\!\frac{58}{27}\pi ^2 \!+\!\frac{40}{3}\zeta _3 \Big ]. \end{aligned}$$
(2.36)

2.2.4 \(\gamma \rightarrow q {{\bar{q}}}\) in FDR

2.2.5 NLO

The computation of \(\gamma \rightarrow q {{\bar{q}}}\) in fdr at NLO has been discussed in [16]. Here we just list the results for the virtual and real corrections, as given in (5.2) and (5.35) of [16]. They read

$$\begin{aligned} \sigma ^{(v)}_\mathrm{FDR}&= \sigma ^{(0)}\,a_s\,{C_F}\, \Big [ \!-\!2 L^2 \!-\! 6 L \!-\!14 \!+\!2\pi ^2 \Big ], \end{aligned}$$
(2.37)
$$\begin{aligned} \sigma ^{(r)}_\mathrm{FDR}&= \sigma ^{(0)}\,a_s\,{C_F}\, \Big [ \!+\! 2 L^2 \!+\!6 L \!+\!17 \!-\!2\pi ^2 \Big ], \end{aligned}$$
(2.38)

where L is defined in (2.25), with s now the centre-of-mass energy squared.

2.2.6 NNLO, \(N_F\) part

The \(N_F\) parts contributing to the NNLO cross section in fdr are given in (5.13) of [25] and read

$$\begin{aligned} \sigma ^{(vv)}_{{\textsc {fdr}}}(N_F)&=\sigma ^{(0)}\, a_{s}^2\,C_F N_F \Big [ \!-\!\frac{8}{9} L^3 \!+\!\frac{2}{9} L^2 \nonumber \\&\quad + L \Big (\frac{278}{27}\!+\!\frac{4}{9}\pi ^2\Big ) +\frac{3355}{81} \!-\!\frac{76}{27}\pi ^2 \!-\!\frac{16}{3}\zeta _3 \Big ], \end{aligned}$$
(2.39)
$$\begin{aligned} \sigma ^{(rv)}_{{\textsc {fdr}}}(N_F)&=\sigma ^{(0)}\, a_{s}^2\,C_F N_F \Big [ \!+\!\frac{4}{3} L^3 \!+\! 4 L^2+ L \Big (\frac{34}{3}\!-\!\frac{4}{3}\pi ^2\Big ) \Big ], \end{aligned}$$
(2.40)
$$\begin{aligned} \sigma ^{(rr)}_{{\textsc {fdr}}}(N_F)&=\sigma ^{(0)}\, a_{s}^2\,C_F N_F \Big [ \!-\!\frac{4}{9} L^3 \!-\!\frac{38}{9} L^2\nonumber \\&\quad + L \Big ( \!-\!\frac{584}{27} \!+\!\frac{8}{9}\pi ^2 \Big )-\frac{4246}{81} \!+\!\frac{76}{27}\pi ^2 \!+\!\frac{40}{3}\zeta _3 \Big ]. \end{aligned}$$
(2.41)

2.3 Relations between FDH and FDR

2.3.1 NLO

The relation between IR divergences of virtual (and therefore real) one-loop results obtained in fdh and fdr has been established in (5.41) of [16] for the process \(\gamma ^*\!\rightarrow \! q {\bar{q}}\) and reads

$$\begin{aligned} \frac{1}{\epsilon ^2} \leftrightarrow \frac{1}{2} L^2_{},\quad \frac{1}{\epsilon } \leftrightarrow L_{}. \end{aligned}$$
(2.42)

These relations are here confirmed by the results of \(H\!\rightarrow \!b {\bar{b}}\). In fact, (2.9) and (2.32) are related through (2.42) and so are (2.7) and (2.31) if the \({\overline{\textsc {ms}}}\) mass (2.30) is used. Moreover, if \(\epsilon \!=\!0\) in prefactors that are related to integration in \({d}\) dimensions, i.e. \(c_{\Gamma }(\epsilon \!=\!0)\), \(\Phi _{2}(\epsilon \!=\!0)\), and \(\Phi _{3}(\epsilon \!=\!0)\) in (2.8), the finite terms of the individual parts are the same in fdh and fdr, including the \(\pi ^2\) terms.

In order to find similar transition rules at NNLO, we also follow a second approach for the comparison between fdh and fdr which is slightly different. To start, we multiply a generic virtual one-loop result obtained in fdh by an \(\epsilon \)-dependent function \(\phi _1(\epsilon )\) and demand equality with the corresponding fdr result, i.e.

$$\begin{aligned} \phi _1(\epsilon )\,\Gamma ^{(v)}_{ \textsc {fdh}} \,\equiv \, \Gamma ^{(v)}_{ \textsc {fdr}}, \quad \phi _1(\epsilon )\,\sigma ^{(v)}_{ \textsc {fdh}} \,\equiv \, \sigma ^{(v)}_{ \textsc {fdr}}\, \end{aligned}$$
(2.43)

with

$$\begin{aligned} \phi _1(\epsilon ) = 1 +\epsilon \, b_{11} +\epsilon ^2\, b_{12}. \end{aligned}$$
(2.44)

The coefficients \(b_{11}\) and \(b_{12}\) are so far unknown and can be obtained by inserting known fdh and fdr results. Before we can do this, however, we have to rescale powers of \(1/\epsilon \) by using

$$\begin{aligned} \Big (\frac{1}{\epsilon }\Big )^{k_1} \rightarrow (\lambda _1\,L)^{k_1},\quad k_1\in \{1,2\}. \end{aligned}$$
(2.45)

The scale factor \(\lambda _1\) is so far unknown and sets the size of the fdh \(1/\epsilon \)-pole with respect to the fdr logarithm L at NLO. Evaluating (2.43) by using for instance (2.16) and (2.37) we find after expanding, applying shift (2.45), and drop** \({\mathcal {O}}(\epsilon )\) terms

$$\begin{aligned} b_{11}&\ = \ -\frac{3}{2}+\frac{3}{2\,\lambda _1}, \end{aligned}$$
(2.46a)
$$\begin{aligned} b_{12}&\ = \ +\frac{9}{4}-\frac{9}{4\,\lambda _1}, \end{aligned}$$
(2.46b)
$$\begin{aligned} (\lambda _1)^2&\ = \ \frac{1}{2}. \end{aligned}$$
(2.46c)

Equations (2.44)–(2.46) are equivalent to (2.42) in that they allow to translate virtual (and therefore real) one-loop results obtained in fdh to fdr and vice versa, including the finite terms. The validity of the rules has been checked explicitly for NLO corrections to \(H\!\rightarrow \!b{\bar{b}}\) and \(\gamma ^*\!\rightarrow \!q{\bar{q}}\).

2.3.2 NNLO, double-virtual

We start our NNLO comparisons with the IR divergences in the \(N_F\) part of the double-virtual contributions. As mentioned repeatedly, to get correct and unitary NNLO results in fdh it is crucial to distinguish evanescent couplings at NLO, see for instance (2.6) and the text below as well as (2.16). In fdr, on the other hand, unitarity is restored by treating UV-divergent one-loop subdiagrams appearing in two-loop amplitudes in the same way they are treated at NLO. This can be achieved by introducing ‘extra-integrals’, as explained in Appendix A of [25]. After taking this contribution into account, it is possible to compare the IR divergences of the double-virtual results obtained in fdh and fdr.

In order to find transition rules between the two schemes we follow the approach described in the previous section and adjust the involved quantities accordingly, i.e. we demand

$$\begin{aligned}&\phi _2(\epsilon )\,\Gamma ^{(vv)}_{ \textsc {fdh}}(N_F) \ \equiv \ \Gamma ^{(vv)}_{ \textsc {fdr}}(N_F),\nonumber \\&\phi _2(\epsilon )\,\sigma ^{(vv)}_{ \textsc {fdh}}(N_F) \ \equiv \ \sigma ^{(vv)}_{ \textsc {fdr}}(N_F)\, \end{aligned}$$
(2.47)

with

$$\begin{aligned} \phi _2(\epsilon ) = 1 +\epsilon \, b_{21} +\epsilon ^2\, b_{22} +\epsilon ^3\, b_{23}\, \end{aligned}$$
(2.48)

and get after the rescaling

$$\begin{aligned} \Big (\frac{1}{\epsilon }\Big )^{k_2} \rightarrow (\lambda _2\,L)^{k_2},\quad k_2\in \{1,2,3\}\, \end{aligned}$$
(2.49)

and the use of (2.12c) and (2.20) the two-loop coefficients

$$\begin{aligned} b_{21}&= -\frac{4}{9} \!-\!\frac{1}{9\,(\lambda _{2})^2}, \end{aligned}$$
(2.50a)
$$\begin{aligned} b_{22}&= \frac{154}{81} \!-\!\frac{139}{27\,\lambda _{2}} \!+\!\frac{4}{81\,(\lambda _{2})^2} \!+\!\pi ^2 \Big ( \frac{1}{6} \!-\!\frac{2}{9\,\lambda _{2}} \Big ), \end{aligned}$$
(2.50b)
$$\begin{aligned} b_{23}&= -\frac{7621}{729} \!+\!\frac{556}{243\,\lambda _{2}} \!-\!\frac{154}{729\,(\lambda _{2})^2}\nonumber \\&\quad -\pi ^2 \Big ( \frac{23}{54} \!-\!\frac{8}{81\,\lambda _{2}} \!+\!\frac{1}{54\,(\lambda _{2})^2} \Big ) \!+\!\frac{26}{9}\zeta _3 , \end{aligned}$$
(2.50c)
$$\begin{aligned} (\lambda _2)^3&\ = \ \frac{4}{9}. \end{aligned}$$
(2.50d)

Similar to the one-loop case, we have reabsorbed the difference between the two schemes in the prefactor \(\phi _2(\epsilon )\), including the finite terms. Restricting ourselves to the \(N_F\) part of the double-virtual corrections for a single process, (2.48)–(2.50) can be seen as using four parameters \(\{\lambda _2, b_{21}, b_{22}, b_{23}\}\) to enforce four equalities between the \(L^n\) and \(\epsilon ^{-n}\) terms for \(n\in \{3,2,1,0\}\). What is remarkable is that the same relations hold for \(H\!\rightarrow \!b{\bar{b}}\) and \(\gamma ^*\!\rightarrow \!q{\bar{q}}\). Despite their apparent similarity, these two processes have a completely different behaviour regarding UV renormalisation. The question whether these rules also apply to the \(C_F^2\) and \(C_A\,C_F\) part of the aforementioned processes or to other processes can not be answered at the moment. Of course, the precise form of the subleading poles depends on whether or not prefactors like \(c_\Gamma ^2(\epsilon )\) are factored out in the fdh results. This explains the \(\pi ^2\) and \(\zeta _3\) terms in (2.50b) and (2.50c). Moreover, note that the two-loop scaling factor (2.50d) is different from the one-loop scaling (2.46c).

2.3.3 NNLO, real-virtual

Regarding the real-virtual corrections we first notice that for the considered processes the \(N_F\) part solely stems from the (sub)renormalisation of the bare couplings in the real one-loop result. In other words, it is given by the product of two one-loop quantities: the real NLO contribution (which contains double and single IR divergences) times a one-loop renormalisation constant (which contains a single UV divergence).

In fdh, the corresponding terms originate from inserting (2.1a) and (2.1b) in (2.9) and (2.17) as the Yukawa coupling \(y_b^0\) does not depend on \(N_F\) at NLO. Similar to the double-virtual contributions it is crucial to distinguish gauge and evanescent couplings in order to avoid the wrong UV-renormalisation of the \(a_e\) terms. Ignoring this distinction, however, leads to results in ‘naive’ fdh which are different from fdh:

$$\begin{aligned} {\textsc {fdh}}{:}\quad&\Gamma ^{(rv)}_{{\textsc {fdh}}}(N_F)= \frac{8}{3\,\epsilon ^3} + \frac{4}{\epsilon ^2}+ \frac{1}{\epsilon } \Big ( 12 \!-\!\frac{14\pi ^2}{9} \Big ) +{\mathcal {O}}(\epsilon ^0), \end{aligned}$$
(2.51)
$$\begin{aligned}&\sigma ^{(rv)}_{{\textsc {fdh}}}(N_F)\propto \frac{8}{3\,\epsilon ^3} + \frac{4}{\epsilon ^2} + \frac{1}{\epsilon } \Big ( \frac{32}{3} \!-\!\frac{14\pi ^2}{9} \Big ) +{\mathcal {O}}(\epsilon ^0), \end{aligned}$$
(2.52)
$$\begin{aligned} \text {`naive'}\, \textsc {fdh}{:}\quad&\Gamma ^{(rv)}_{{\textsc {fdh}'}}(N_F)= \frac{8}{3\,\epsilon ^3} + \frac{4}{\epsilon ^2} + \frac{1}{\epsilon } \Big ( \frac{38}{3} \!-\!\frac{14\pi ^2}{9} \Big ) +{\mathcal {O}}(\epsilon ^0) , \end{aligned}$$
(2.53)
$$\begin{aligned}&\sigma ^{(rv)}_{{\textsc {fdh}'}}(N_F)\propto \frac{8}{3\,\epsilon ^3} + \frac{4}{\epsilon ^2} + \frac{1}{\epsilon } \Big ( \frac{34}{3} \!-\!\frac{14\pi ^2}{9} \Big ) +{\mathcal {O}}(\epsilon ^0). \end{aligned}$$
(2.54)

In fdr, the corresponding results are given in (2.35) and (2.41) and read

$$\begin{aligned} {\textsc {fdr}}{:}\quad \Gamma ^{(rv)}_{{\textsc {fdh}}}(N_F)&= \frac{4}{3} L^3 \!+\! 4 L^2 \!+\! L \Big (\frac{38}{3}\!-\!\frac{4}{3}\pi ^2\Big ), \end{aligned}$$
(2.55)
$$\begin{aligned} \sigma ^{(rv)}_{{\textsc {fdh}}}(N_F)&= \frac{4}{3} L^3 \!+\! 4 L^2 \!+\! L \Big (\frac{34}{3}\!-\!\frac{4}{3}\pi ^2\Big ). \end{aligned}$$
(2.56)

Similar to fdh, they stem from the \(a_s\)-renormalisation in (2.32) and (2.38), see also (3.1) and (3.2) of [25]. In contrast to the double-virtual contributions, however, the different renormalisation of \(a_e\) and \(a_s\) is not taken into account via ‘extra-integrals’. Accordingly, (2.55) and (2.56) correspond to ‘naive’ fdh, i.e. (2.53) and (2.54), rather than fdh. Since only one-loop quantities are involved, the transition between ‘naive’ fdh and fdr is already known and given by e.g. (2.42) times the transition \(1/\epsilon _{{\textsc {uv}}}\leftrightarrow L\) for the UV divergence, i.e.

$$\begin{aligned}&\frac{1}{\epsilon ^2_{{\textsc {ir}}}}\times \frac{1}{\epsilon _{{\textsc {uv}}}} \leftrightarrow \frac{1}{2} L^2 \times L,\quad \frac{1}{\epsilon _{{\textsc {ir}}}}\times \frac{1}{\epsilon _{{\textsc {uv}}}} \leftrightarrow L \times L, \nonumber \\&(\epsilon _{{\textsc {ir}}})^0\times \frac{1}{\epsilon _{{\textsc {uv}}}} \leftrightarrow 1 \times L. \end{aligned}$$
(2.57)

Note that the transition of the divergent \(\pi ^2\) terms depends on the \({d}\)-dependent prefactors that have been factored out in the fdh result, similar to the double-virtual contributions.

Regarding the finite terms it is clear that a transition between (‘naive’) fdh and fdr can not exist. The reason is that for the considered processes the \(N_F\) part of the real-virtual contribution in fdr is obtained via multiplying (2.32) and (2.38) by a pure UV divergence. As a consequence, (2.55) and (2.56) only contain pure divergences (which are parametrized as powers of L) and no finite terms. Therefore, the \(N_F\) part of the real-virtual contribution alone does not contain a finite part which is different from any dimensional scheme.

2.3.4 NNLO, double-real

In the previous section we have seen that a transition rule for real-virtual contributions between fdh and fdr does not exist. As the physical result has to be scheme independent, this is also the case for the double-real contributions. Given the fact that a transition rule exists for the double-virtual part, however, it is clear that (2.48) can also be used to translate the sum of the real-virtual and double-real components from fdh to fdr. This is due to the fact that the divergences are the same (apart from their sign) and that the finite term is given by the difference between the physical result and the finite part of the double-virtual contribution. We have checked this explicitly and find indeed

$$\begin{aligned}&\phi _2(\epsilon ) \Big [ \Gamma ^{(rv)}_{{\textsc {fdh}}}(N_F) +\Gamma ^{(rr)}_{{\textsc {fdh}}}(N_F) \Big ] = \Gamma ^{(rv)}_{{\textsc {fdr}}}(N_F) +\Gamma ^{(rr)}_{{\textsc {fdr}}}(N_F), \end{aligned}$$
(2.58)
$$\begin{aligned}&\phi _2(\epsilon ) \Big [ \sigma ^{(rv)}_{{\textsc {fdh}}}(N_F) +\sigma ^{(rr)}_{{\textsc {fdh}}}(N_F) \Big ] = \sigma ^{(rv)}_{{\textsc {fdr}}}(N_F) +\sigma ^{(rr)}_{{\textsc {fdr}}}(N_F). \end{aligned}$$
(2.59)

2.3.5 Unitarity restoration in FDH and FDR

As we have commented many times, fdh and fdr use different strategies to restore unitarity. In this paragraph, we report on an attempt towards a comparison of the two unitarity restoration methods.

Our starting point is ‘naive’ fdh  in which no distinction is made between gauge and evanescent couplings, and we try to extract from fdr the contribution of the fdh evanescent \(a_e\) terms. This is achieved by interpreting the fdr ‘extra-integrals’ as UV-divergent dimensionally regulated integrals subtracted at the integrand level, as explained in Section 6 of [\(EEI_b\)).Footnote 7 These integrals contain now \(1/\epsilon \) poles of UV origin suitable to be combined with the ‘naive’ fdh expressions. For the regarded processes, their contribution corresponds to the evanescent \(a_e\) terms in (2.6) and (2.16), respectively, times the difference of the renormalisation constants \(\delta _Z\!=\!(Z_{\alpha _e}\!-\!Z_{\alpha _s})\), whose value in the \({\overline{\textsc {ms}}}\)-scheme is given in (2.2),

$$\begin{aligned} \Gamma ^{(vv)}_{EEI_b} \ =\&\Gamma ^{(0)}_{\text {}}\times \delta _Z^{}\times C_F\,n_\epsilon \Big [\frac{2}{\epsilon }+4\Big ], \end{aligned}$$
(2.60a)
$$\begin{aligned} \sigma ^{(vv)}_{EEI_b} \ =\&\sigma ^{(0)}_{\text {}}\times \delta _Z^{}\times C_F\,n_\epsilon \Big [\frac{1}{\epsilon }+1\Big ]. \end{aligned}$$
(2.60b)

More precisely, (2.6) and (2.16) are reproduced if the contribution of (2.60) is added to the ‘naive’ FDH results. Note that, since \(\delta _Z\!=\!{\mathcal {O}}(\epsilon ^{\!-\!1})\) and \(n_\epsilon \!=\!2\epsilon \), the \(EEI_b\) are of \({\mathcal {O}}(\epsilon ^{\!-\!1})\) and that (2.60) refers to the full contributions, not only the \(N_F\) parts. Finally, it should be mentioned that the contributions in (2.60) have been extracted from off-shell diagrams, while the unitarity restoring fdr procedure to be used on-shell is slightly different [fdu framework to obtain the NNLO QED corrections to the \(N_f\) terms associated to \(\gamma ^{*}\!\rightarrow \! q {{\bar{q}}} (g)\).

We would like to highlight that the fdu/ltd framework has been extended and improved since the last WorkStop/ThinkStart meeting in 2016 [16]. Therefore, we briefly review the new features and properties that have been recently improved.

3.1 Dual representations from the LTD theorem

The ltd theorem is based on a clever application of Cauchy’s residue theorem (CRT) to the loop integrals. The original version was developed in Refs. [47, 49], where CRT was used to integrate out the energy component of each loop momenta. This procedure reduces loop amplitudes to collections of tree-level-like diagrams with a modification of the customary Feynman prescription. Thus, given a loop line associated to the momentum \(q_i\), we should use the rule

$$\begin{aligned} G_F(q_i)= & {} \frac{1}{q_i^2-m_i^2+\imath 0} \rightarrow G_D(q_i;q_j) \nonumber \\= & {} \frac{1}{q_i^2-m_i^2+\imath \, \eta \cdot (q_j-q_i)} , \end{aligned}$$
(3.1)

to replace the Feynman propagators when the momentum \(q_j\) is set on-shell. In this expression, \(\eta \) is a future-like vector that defines the explicit dependence of the prescription on the momentum flow, and the integration contours are closed on the lower half-plane such that only those poles with negative imaginary components are selected. It is important to notice that this dual representation is equivalent to the usual Feynman tree theorem (ftt) [63, 64], and the multiple cut information is encoded within the momentum-dependent dual prescription.

Recently, a new representation of dual integrals was achieved through the iterated calculation of residues, leading to what we call nested residues [54,55,56,57]. This strategy turns out to be more efficient computationally, since it allows a straightforward algorithmic implementation. Hence, in order to elucidate how ltd formalism works, let us consider an L-loop scattering amplitude with N external legs, \(\{p_j\}_N\) and n internal lines, in the Feynman representation,

$$\begin{aligned}&{\mathcal {A}}_N^{(L)}(1,\ldots , n) = \int _{\ell _1, \ldots , \ell _L} {\mathcal {A}}_F^{(L)}(1,\ldots , n) \nonumber \\&\quad = \int _{\ell _1, \ldots , \ell _L} {\mathcal {N}}( \{ \ell _s\}_L, \{ p_j\}_N) \, G_F(1,\ldots , n) . \end{aligned}$$
(3.2)

This amplitude is naturally defined over the Minkowski space of the L loop momenta, \(\{\ell _s\}_L\). The singular structure is associated to the denominators introduced by the Feynman propagators,

$$\begin{aligned} G_F(1,\ldots , n) = \prod _{i\in 1\cup \ldots \cup n} \left( G_F(q_i) \right) ^{a_i} , \end{aligned}$$
(3.3)

with

$$\begin{aligned} G_{F}\left( q_{i}\right)&=\frac{1}{q_{i}^{2}-m_{i}^{2}+\imath 0}\nonumber \\&=\frac{1}{\left( q_{i,0}-q_{i,0}^{\left( +\right) }\right) \left( q_{i,0}+q_{i,0}^{\left( +\right) }\right) }, \end{aligned}$$
(3.4)

where the \(q_{i},m_{i}\) and \(+\imath 0\) correspond to the loop momentum, its mass, and the infinitesimal Feynman prescription. Furthermore, we explicitly pull out the dependence on the energy component of the loop momentum \(q_{i,0}\) together with its on-shell energy,

$$\begin{aligned} q_{i,0}^{\left( +\right) }&=\sqrt{{\varvec{q}}_{i}^{2}+m_{i}^{2}-\imath 0}, \end{aligned}$$
(3.5)

that is expressed in terms of the spatial components of the loop momentum.

Besides, \(a_i\) and \(\{1,\ldots ,n \}\) in Eq. (3.2) correspond to the arbitrary positive integers, raising the powers of the propagators, and sets containing internal momenta of the form \(q_{i_j}=k_{j}+\ell _i \in i\), respectively. In the following, the \(a_j\) exponents will not be included in the notation because the treatment of the expressions is independent of their explicit values.

As in the previous ltd representation, the dual contributions are obtained by integrating out one degree of freedom per loop through the Cauchy residue theorem. Iterating this procedure, we can write [54],

$$\begin{aligned}&{\mathcal {A}}_D^{(L)}(1,\ldots , r; r+1,\dots , n) \nonumber \\&\quad =-2\pi \imath \sum _{i_r \in r} {\mathrm{Res}} ({\mathcal {A}}_D^{(L)}(1, \ldots , r-1;r, \ldots , n),\nonumber \\&\qquad {\mathrm{Im}}(\eta \cdot q_{i_r})<0), \end{aligned}$$
(3.6)

starting from

$$\begin{aligned}&{\mathcal {A}}_D^{(L)}(1; 2, \ldots , n) \nonumber \\&\quad =-2\pi \imath \sum _{i_1 \in 1} {\mathrm{Res}} ({\mathcal {A}}_F^{(L)}(1, \dots , n), \,{\mathrm{Im}}(\eta \cdot q_{i_1})<0). \end{aligned}$$
(3.7)

All the sets in Eq. (3.6) before the semicolon contain one propagator that has been set on shell, while all the propagators belonging the sets that appear after the semicolon remain off shell. The sum over all possible on-shell configurations is implicit. Regarding the prescription, the contour choice is the same used in the original ltd formulation. It is worth mentioning that the ltd representation is independent of the coordinate system, and that there are some non-trivial cancellations when iterating the residue loop-by-loop. The last result implies that only those poles whose loci is always on the lower complex half-plane will lead to non-vanishing contributions; the others, called displaced poles will cancel at each iterative step [57]. Moreover, these contributions can be mapped onto the usual cut diagrams, thus allowing a graphical interpretation.

Finally, we would like to emphasise that the ltd representation corresponds to tree-level like objects integrated in the Euclidean space. In this way, the application of this novel representation of multi-loop scattering amplitudes allows to open loops into trees, aiming at finding a natural connection with the structures exhibited in the real emission contributions.

3.2 Local cancellation of IR singularities through kinematic map**s

Once the ltd theorem is applied to any multi-loop multi-leg scattering amplitude, we obtain a representation involving tree-level like objects and phase-space integrals. This leads to an important conceptual simplification to understand the origin of IR and threshold singularities, at integrand level. By analysing the intersection of the integration hyperboloids associated to the dual contributions for different cuts [fdu formalism has been successfully proven to deal with NLO corrections to physical observables, such as cross sections and decay rates [

$$\begin{aligned} \sigma ^{(0)} = \int {\mathrm{dPS}}^{1\rightarrow n} |{{{\mathcal {M}}}}^{(0)}_{n}|^2 \, {{{\mathcal {S}}}}_0(\{p_i\}) , \end{aligned}$$
(3.8)

with \(|{{{\mathcal {M}}}}^{(0)}_{n}|^2\) the Born squared amplitude and \(\mathcal{S}_0\) the IR-safe measure function that defines the physical observable. On the one hand, the virtual one-loop contribution is

$$\begin{aligned} \sigma _V^{(1)} = \int {\mathrm{dPS}}^{1\rightarrow n} \, \int _{\ell } 2 {\mathrm{Re}}({{{\mathcal {M}}}}^{(1)}_{n} {{{\mathcal {M}}}}^{(0),*}_{n}) \, {{{\mathcal {S}}}}_0(\{p_i\}) , \end{aligned}$$
(3.9)

where \({\mathrm{Re}}({{{\mathcal {M}}}}^{(1)}_{n} {{{\mathcal {M}}}}^{(0),*}_{n})\) corresponds to the interference between the one-loop and the Born amplitude, including also factors that might be related to self-energy contributions. On the other hand, we need to take into account the real-emission contribution,

$$\begin{aligned} \sigma _R^{(1)} = \int {\mathrm{dPS}}^{1\rightarrow n+1} \, |\mathcal{M}^{(0)}_{n+1}|^2 \, {{{\mathcal {S}}}}_1(\{{p'}_i\}) , \end{aligned}$$
(3.10)

which is characterised by the presence of an additional external particle. Notice that the measure function, \(\mathcal{S}_1(\{{p'}_i\})\), is extended to include the extra radiation and it must fulfil the reduction property \({{{\mathcal {S}}}}_1 \rightarrow {{{\mathcal {S}}}}_0\) in the IR limits. Moreover, the corresponding IR singularities can be disentangled and isolated into disjoint regions of the real-emission phase-space. Following a slicing strategy, we introduce a partition \({{{\mathcal {R}}}}_i\) and define

$$\begin{aligned} \sigma _{R,i}^{(1)} = \int {\mathrm{dPS}}^{1\rightarrow n+1} \, d\sigma _R^{(1)} \, {{{\mathcal {R}}}}_i , \end{aligned}$$
(3.11)

that fulfils \(\sum _ i \sigma _{R,i}^{(1)} = {\sigma }_R^{(1)}\) and that only one IR divergent configuration is allowed inside different \({{{\mathcal {R}}}}_i\).

Regarding the virtual contribution, the application of ltd to Eq. (3.9) will produce a sum of cuts leading to dual contributions, namely

$$\begin{aligned} \sigma _D^{(1)}= & {} \int {\mathrm{dPS}}^{1\rightarrow n} \sum _{i=1}^{N} \, \int _{\vec {\ell }} I_i(q_i) \, {{{\mathcal {S}}}}_0(\{p_i\}) \nonumber \\\equiv & {} \, \int {\mathrm{dPS}}^{1\rightarrow n} \sum _{i=1}^{N} \sigma _{D,i}^{(1)} \ , \end{aligned}$$
(3.12)

with N the number of different internal lines. Each line is characterised by a momenta \(q_i\), which is set on shell in the different dual terms. The presence of this extra on-shell momenta allows to establish a connection with the real emission contributions, through a proper momentum map**. Explicitly, for the NLO case, we have a bijective transformation,

$$\begin{aligned} {{{\mathcal {T}}}}_i(\{p_1,\ldots \,p_n,q_i\}) \rightarrow \{p'_1,\ldots \,p'_{n+1} \} \, \end{aligned}$$
(3.13)

restricted to some partition \({{{\mathcal {R}}}}_i\). At this point, the dual contributions can be understood as local counterterms for the real corrections, whilst the development of the transformations \({{{\mathcal {T}}}}_i\) is guided by the structure of the partition \(\mathcal{R}_i\). This partition is based on the FKS or CS strategy, i.e. splitting the phase-space into disjoint regions containing at most one infrared singularity. Then, through a proper study of the cut structure, a connection among dual integrals and partitions is established, in such a way that a map** \({{{\mathcal {T}}}}_i\) exists and leads to local cancellations of IR singularities.

3.3 Multi-loop local UV renormalisation

The successful identification and cancellation of IR singularities lead to cross sections with only UV singularities. The standard procedure to remove those divergences requires renormalisation of the field wave-functions and couplings, this is what we aim to reach through the construction of local UV counterterms. Since all these elements have to cast only UV divergences in the ltd framework, it is important to study carefully the analytic properties of the UV integrands that will be added to the real radiation. At one-loop, it has been considered the regime of massless and massive particles propagating in the loop [42,43,44]. The transition between massless and massive renormalisation constants has found to be smooth and singularities are well understood in this framework. Let us point out a crucial difference between the standard renormalisation constant in dreg and ltd.

Wave-function renormalisation constants emerge from the computation of self-energy diagrams. In particular, massless bubble diagrams in dreg do not present any problem since the IR and UV divergences are considered as equal, therefore, the full integral vanishes. On the contrary, the same integral in the fdu/ltd framework will contribute to the IR and UV regions because they are treated separately, therefore, the integral cannot be removed even if the full integral is actually zero. We stress that integrands in the fdu/ltd are split into the IR and UV domains, and they have to be keep in this way in order to render the full integrands free of singularities in the fdu formalism.

Let us review the basic ideas of one-loop renormalisation constant; these ideas are implemented and improved for the two-loop case and beyond. The massive wave-function renormalisation constant, in the Feynman gauge, at one-loop can be obtained from,

$$\begin{aligned}&\Delta Z_2(p_1) \nonumber \\&\quad = -g_{s}^2 \, {C_F}\, \int _{\ell } G_F(q_1) \, G_F(q_3) \, \left[ (d-2)\frac{q_1 \cdot p_2}{p_1 \cdot p_2} \right. \nonumber \\&\qquad +4\, M^2 \left. \left( 1- \frac{q_1 \cdot p_2}{p_1 \cdot p_2}\right) G_F(q_3)\right] , \end{aligned}$$
(3.14)

which represent the unintegrated form. This integral is obtained by the standard Feynman rules procedure. It is important to remark that the limit of massless case is straightforward achieved from Eq. (3.14), so this expression is the most general of \(\Delta Z_2(p_1; M)\). Since \(\Delta Z_2(p_1; M)\) contains singularities associated to the UV domain, it is important to find the UV component of the Eq. (3.14), \(\Delta Z_2^{{{\mathrm{UV}}}}\), and subtract it in order to find the UV-free wave-function renormalisation constant, \(\Delta Z_2^{{{\mathrm{IR}}}}\). The UV part is extracted by performing an expansion of the integrand around the UV propagator \(G_F(q_{\mathrm{UV}}) = (q^2_{\mathrm{UV}}-\mu _{\mathrm{UV}}^2 + \imath 0)^{-1}\). In particular, for Eq. (3.14), it is found

$$\begin{aligned} \Delta Z_2^{\text {UV}}(p_1)&= (2-d)\, g_{s}^2 \, {C_F}\, \nonumber \\&\quad \times \int _{\ell } \big [G_F(q_\text {UV})\big ]^2 \, \left( 1+\frac{q_\text {UV} \cdot p_2}{p_1 \cdot p_2}\right) \nonumber \\&\quad \times \Big [1\!-\!G_F(q_\text {UV})(2\, q_\text {UV} \cdot p_1 + \mu ^2_\text {UV})\Big ] . \end{aligned}$$
(3.15)

Finally, \(\Delta Z_2^{\mathrm{IR}}\) is given by,

$$\begin{aligned} \Delta Z_2^{\text {IR}} = \Delta Z_2 - \Delta Z_2^{\text {UV}}, \end{aligned}$$
(3.16)

which contains only IR singularities and they are needed to cancel the remaining singularities at the cross section level.

We focus now on subtraction of UV singularities for two-loop amplitudes. In general, for the two-loop case, the integrand involved is cumbersome, therefore, we will use this document to emphasise the procedure to renormalise locally the UV behaviour of any two-loop amplitude. The procedure is valid if the integrand is free of infrared singularities. In this sense, it is mandatory to subtract first all IR divergences and then apply the following algorithm. This algorithm has been extensively discussed in [51,52,53], with applications of one- and two-loop scattering amplitudes.

Let us now consider a generic two-loop scattering amplitude free of IR singularities,

$$\begin{aligned} {\mathcal {A}}^{(2)} = \int _{\ell _1}\int _{\ell _2} {\mathcal {I}}(\ell _1, \ell _2) , \end{aligned}$$
(3.17)

where the integrand is a function of the integration variables \(\ell _1\) and \(\ell _2\). UV divergences shall appear when the limit of \(|\vec {\ell }_1|\) and \(|\vec {\ell }_2|\) go to infinity. In the two-loop case, there are three UV limits to be considered. The limit when \(|\vec {\ell }_1|\) goes to infinity and \(|\vec {\ell }_2|\) remains fixed, the other way around, and when \(|\vec {\ell }_1|\) and \(|\vec {\ell }_2|\) go to infinity simultaneously. Based on the ideas developed at one-loop, the UV divergences can be extracted from the integrand by making the replacement,Footnote 9

$$\begin{aligned}&{\mathcal {S}}_{j,{{\mathrm{UV}}}} : \{ \ell _j^2\vert \ell _j \cdot k_i\} \rightarrow \{\lambda ^2 q_{j, {{\mathrm{UV}}}}^2 \nonumber \\&\quad + (1-\lambda ^2)\mu _{{\mathrm{UV}}}^2 \vert \lambda \, q_{j,{{\mathrm{UV}}}}\cdot k_i \} , \end{aligned}$$
(3.18)

for a given loop momentum \(\ell _j\) and expanding the expression up to logarithmic order around the UV propagator. This construction is represented by the \(L_{\lambda }\) operator. It is worth mentioning that the result shall generate a finite part after integration that has to be fixed to reproduce the correct value of the integral. Therefore, the first counterterms will be obtained by

$$\begin{aligned} {\mathcal {A}}_{j,{{\mathrm{UV}}}}^{(2)} =\,&L_{\lambda } \left( {\mathcal {A}}^{(2)}\Big |_{{\mathcal {S}}_{j,{\mathrm{UV}}}} \right) \nonumber \\&-\,d_{j,{{\mathrm{UV}}}} \, \mu _{{\mathrm{UV}}}^2 \int _{\ell _j} (G_F(q_{j,{{\mathrm{UV}}}}))^3 , \end{aligned}$$
(3.19)

where \(d_{j,{\mathrm{UV}}}\) is the fixing parameter which makes the finite part of integral to be zero in the \(\overline{{\mathrm{MS}}}\) scheme.

Now, after the complete subtraction of these counterterms, the remaining divergences shall occur when both \(|\vec {\ell }_1|\) and \(|\vec {\ell }_1|\) approach to infinity simultaneously. In this case, the following replacement is implemented,

$$\begin{aligned}&{\mathcal {S}}_{{{\mathrm{UV}}}^2} : \{ \ell _j^2\vert \ell _j \cdot \ell _k\vert \ell _j \cdot k_i\} \nonumber \\&\quad \rightarrow \{\lambda ^2 q_{j, {{\mathrm{UV}}}}^2 + (1-\lambda ^2)\mu _{{{\mathrm{UV}}}}^2 \vert \lambda ^2 q_{j,{{\mathrm{UV}}}} \cdot q_{k, {{\mathrm{UV}}}} \nonumber \\&\qquad +(1-\lambda ^2) \, \mu _{{{\mathrm{UV}}}}^2/2 \vert \lambda \, q_{j,{{\mathrm{UV}}}}\cdot k_i \} \, \end{aligned}$$
(3.20)

on the subtracted integrand. Then, the application of the \(L_{\lambda }\) operation has to be made and the fixing parameter is again needed in order to build properly the counterterm, \({\mathcal {A}}^{(2)}_{{{\mathrm{UV}}}^2}\). Explicitly,

$$\begin{aligned} {\mathcal {A}}_{{{\mathrm{UV}}}^2}^{(2)} =\,&L_{\lambda } \left( \left( {\mathcal {A}}^{(2)} -\sum _{j=1,2}{\mathcal {A}}^{(2)}_{j,{{\mathrm{UV}}}} \right) \Bigg |_{\mathcal {S_{{{\mathrm{UV}}}^2}}} \right) \nonumber \\&-d_{{{\mathrm{UV}}}^2} \, \mu _{{\mathrm{UV}}}^4 \int _{\ell _1} \int _{\ell _2} (G_F(q_{1,{{\mathrm{UV}}}}))^3(G_F(q_{12,{{\mathrm{UV}}}}))^3 , \end{aligned}$$
(3.21)

where \(d_{{{\mathrm{UV}}}^2}\) is the fixing parameter of the double limit.

Finally, the original amplitude can be renormalised by the subtraction of all UV counterterms, such thatFootnote 10

$$\begin{aligned} {\mathcal {A}}_{\mathrm{R}}^{(2)} = {\mathcal {A}}^{(2)} - {\mathcal {A}}_{1,{{\mathrm{UV}}}}^{(2)}- {\mathcal {A}}_{2,{\mathrm{UV}}}^{(2)} - {\mathcal {A}}_{{{\mathrm{UV}}}^2}^{(2)} , \end{aligned}$$
(3.22)

is free of IR and UV singularities.

Before closing this discussion, let us deepen into the multi-loop case. In this scenario, multiple ultraviolet poles will appear, since all loops have the possibility to tend to infinity at different speed. However, if the amplitude is free of IR singularities, the algorithm presented along this section is still valid. For an L-loop integral, there are \(2^L-1\) UV counterterms at most with the same number of fixing parameters. Therefore, after the proper knowledge of all UV counterterms, the renormalisation of the original L-loop amplitude is achieved and the four-dimensional representation of the integrand can be obtained.

Fig. 4
figure 4

Tree, one- and \(N_F\) two-loop Feynman diagrams for \(\gamma ^*\rightarrow q{\bar{q}}\)

3.4 \(\gamma ^{*}\rightarrow \,q{{\bar{q}}}\) at NNLO

Let us first consider the kinematics of decay of \(\gamma ^{*}\rightarrow q{\bar{q}}\), in which, to keep a simple structure at integrand level, we work with massless particles in the loop. We remark that this choice does not generate difficulties in the evaluation of integrals, within the ltd framework. The latter pattern is because of the way how integrands are expressed, which are inherit of the masses. Equivalently, their dependence is stored in the fixed energy components, \(q_{i,0}^{\left( +\right) }\). Hence, here and in the following processes, \(k=p_{1}+p_{2}+\cdots +p_{n}\), with \(p_{i}^{2}=0,k^{2}=s\) and \(n\le 4\). With this notation in mind, the squared tree-level amplitude can be cast as,

$$\begin{aligned} \omega _{0}=\overline{\left| {\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\right| ^{2}} =\left\langle {\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\right\rangle =\frac{2}{3}e^{2}Q_{q}^{2}N_{c}\left( d-2\right) s, \end{aligned}$$
(3.23)

with \(Q_{q}=\frac{1}{3},-\frac{2}{3}\) and \(N_{c}\) the electric charge and the colour number of the quarks, respectively.

3.4.1 Virtual contributions

The contribution at one-loop, after only performing Dirac algebra and expressing scalar product in terms of denominators, becomes,

$$\begin{aligned}&\left\langle {\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}}^{\left( 1\right) }\right\rangle \nonumber \\&\quad = \imath \,C_{F}g_{s}^{2}\omega _{0}\Bigg [2sI_{111}^{\left( 1\right) }-(d-8)I_{101}^{\left( 1\right) }+\frac{2}{s}I_{1-11}^{\left( 1\right) }\nonumber \\&\qquad +\frac{2}{s}\left( I_{010}^{\left( 1\right) }-I_{100}^{\left( 1\right) }-I_{001}^{\left( 1\right) }\right) -2\left( I_{110}^{\left( 1\right) }+I_{011}^{\left( 1\right) }\right) \Bigg ], \end{aligned}$$
(3.24)

with,

$$\begin{aligned}&I_{\alpha _{1}\alpha _{2}\alpha _{3}}^{\left( 1\right) } =\int _{\ell _{1}}\prod _{i=1}^{3}G_{F}\left( q_{i}^{\alpha _{i}}\right) ,\nonumber \\&\quad q_{1}=\ell _{1},\quad q_{2}=\ell _{1}-p_{1},\quad q_{3}=\ell _{1}-p_{12}. \end{aligned}$$
(3.25)

As mentioned in the former sections, we would like to emphasise that within ltd and, therefore, fdu, scaleless integrals are not set directly to zero as conventionally carried out in dreg. The main reason to do this is to achieve a complete cancellation of singularities at integrand level by kee** as much as possible control on the local structure of the integrands. Thus, if we were in dreg, one finds that the second line in Eq. (3.24) vanishes and,

$$\begin{aligned} I_{1-11}^{\left( 1\right) }&=-\frac{s}{2}I_{101}. \end{aligned}$$
(3.26)

These relations amount to

$$\begin{aligned}&\left\langle {\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}}^{\left( 1\right) }\right\rangle = \imath \,C_{F}g_{s}^{2}\omega _{0}\left[ 2sI_{111}^{\left( 1\right) }-(d-7)I_{101}^{\left( 1\right) }\right] , \end{aligned}$$
(3.27)

where it is straightforward to identify the IR and UV singularities, which come from \(I_{111}^{\left( 1\right) }\) and \(I_{101}^{\left( 1\right) }\), respectively. This is indeed what is traditionally carried out by means of the tensor reduction.

In ltd, we keep scaleless integrals to perform a local UV renormalisation as well as IR subtraction from real corrections, following the lines of the fdu scheme. Thus, applying ltd to Eq. (3.24),

$$\begin{aligned}&\left\langle {\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}}^{\left( 1\right) }\right\rangle =\imath \,C_{F}g_{S}^{2}\omega _{0}\int _{\varvec{\ell }_{1}} \bigg [-\frac{{\mathcal {I}}_{101}^{d}}{2s}\left( -4\left( q_{1,0}^{\left( +\right) }\right) ^{2}\nonumber \right. \\&\left. \qquad +4\left( q_{2,0}^{\left( +\right) }\right) ^{2}+(2d-15)s\right) \nonumber \\&\qquad +\frac{2}{s}\left( {\mathcal {I}}_{010}^{d}-{\mathcal {I}}_{100}^{d}\right) +2s{\mathcal {I}}_{111}^{d}-4{\mathcal {I}}_{110}^{d}\bigg ], \end{aligned}$$
(3.28)

with

$$\begin{aligned}&{\mathcal {I}}_{100}^{d}=\frac{1}{2q_{1,0}^{\left( +\right) }},\quad {\mathcal {I}}_{010}^{d}=\frac{1}{2q_{2,0}^{\left( +\right) }}, \end{aligned}$$
(3.29a)
$$\begin{aligned}&{\mathcal {I}}_{101}^{d}=-\frac{1}{4\left( q_{1,0}^{\left( +\right) }\right) ^{2}}\left( \frac{1}{\lambda _{3}^{+}}+\frac{1}{\lambda _{3}^{-}}\right) , \end{aligned}$$
(3.29b)
$$\begin{aligned}&{\mathcal {I}}_{110}^{d}=-\frac{1}{4q_{1,0}^{\left( +\right) }q_{2,0}^{\left( +\right) }}\left( \frac{1}{\lambda _{1}^{+}}+\frac{1}{\lambda _{1}^{-}}\right) , \end{aligned}$$
(3.29c)
$$\begin{aligned}&{\mathcal {I}}_{111}^{d}=\frac{1}{4\left( q_{1,0}^{\left( +\right) }\right) ^{2}q_{2,0}^{\left( +\right) }}\left( \frac{1}{\lambda _{1}^{+}\lambda _{1}^{-}}+\frac{1}{\lambda _{1}^{+}\lambda _{2}^{-}}+\frac{1}{\lambda _{1}^{-}\lambda _{2}^{+}}\right) , \end{aligned}$$
(3.29d)

where,

$$\begin{aligned}&\lambda _{1}^{\pm }=q_{1,0}^{\left( +\right) }+q_{2,0}^{\left( +\right) }\pm \frac{\sqrt{s}}{2},&\lambda _{2}^{\pm }=2q_{1,0}^{\left( +\right) }\mp \sqrt{s}, \end{aligned}$$
(3.29e)

We remark that the integrands of Eq. (3.28) are computed, without the loss of generality, in the center-of-mass frame, allowing us to have \(q_{3,0}^{\left( +\right) }=q_{1,0}^{\left( +\right) }\). Additionally, the integrands (3.29) are expressed in the multi-loop ltd representation, displaying structure depending only on physical singularities. A noteworthy comment on the structure of (3.29) is in order. The structure of these integrands can easily be related to one-, two- and three-point functions, which have been explicitly computed, independently on the number of loops, in [55]. Although there a simple recipe for the calculation of dual integrals through ltd, it as possible to profit of the explicit causal structure of multi-loop topologies. This is indeed what is carried out in the two- and three-point integrands, where their structure correspond to the Maximal-Loop and Next-to-Maximal-Loop topologies, respectively. A detailed discussion of the structure and the features of these topologies is presented in [55, 57, 68]. Interestingly from Eq. (3.29), the treatment of physical thresholds is straightforward because of the structure the latter hold. In this configuration, in fact, it is possible to obtained up to two “entangled causal” thresholds that are observed from the structure of \(\lambda _{i}^{\pm }\).

Hence, by following the idea presented in the one-loop case, we generate the two-loop contribution, in which we elaborate, for the sake of simplicity, on the two-loop integral,

$$\begin{aligned} I_{\alpha _{1}\cdots \alpha _{7}}^{\left( 2\right) }&=\int _{\ell _{1},\ell _{2}}\prod _{i=1}^{7}G_{F}\left( q_{i}^{\alpha _{i}}\right) , \end{aligned}$$
(3.30)

with

$$\begin{aligned}&q_{1}=\ell _{1},&q_{5}=\ell _{1}-p_{12},\nonumber \\&q_{2}=\ell _{2},&q_{6}=\ell _{2}-p_{12},\nonumber \\&q_{3}=\ell _{1}-p_{1},&q_{7}=\ell _{2}-p_{1}.\nonumber \\&q_{4}=\ell _{1}+\ell _{2}-p_{1}, \end{aligned}$$
(3.31)

Let us remark that the integrand obtained from the Feynman diagram depicted in Fig. 4c one only has five propagators, the additional ones, \(q_{6}\) and \(q_{7}\), are needed to express all scalar products in terms of denominators. This is done to simplify the structure of the integrand, but it is not mandatory and this step can be avoided. In fact, the explicit dependence on the energy component of the loop momenta can always pull out.

Hence, with the above considerations, the two-loop contribution turns out to be,

$$\begin{aligned} \left\langle {\mathcal {A}}_{q{\bar{q}}}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}}^{\left( 2\right) }\right\rangle =C_{F}N_{F}g_{S}^{4}\omega _{0}\left[ \text {some two-loop integrals}\right] . \end{aligned}$$
(3.32)

These two-loop integrals can be further reduced by means of the integration-by-parts identities (IBPs). Moreover, in order not to alter the local structure of the integrands, we do not make use of the zero sector symmetries and loop momentum redefinition. This is done to carefully combine and thus match virtual with real corrections.

Fig. 5
figure 5

Tree level Feynman diagrams for \(\gamma ^*\rightarrow q{\bar{q}}g\) and \(\gamma ^*\rightarrow q{\bar{q}}q'{\bar{q}}'\)

3.4.2 Real contributions

In this section, we list the tree-level amplitudes that are needed to perform a cancellation of the IR singularities, \(\gamma ^{*}\rightarrow q{\bar{q}}g\) and \(\gamma ^{*}\rightarrow q{\bar{q}}q'{\bar{q}}'\).

\(\varvec{\gamma ^{*}\rightarrow q{\bar{q}}g}\)

$$\begin{aligned}&\left\langle {\mathcal {A}}_{q{\bar{q}}g}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}g}^{\left( 0\right) }\right\rangle = \frac{2(d-2)e^{2}N_{c}C_{F}Q_{f}^{2}g_{S}^{2}}{s_{12}s_{23}}\nonumber \\&\quad \times \left( (d-2)\left( s_{12}+s_{23}\right) ^{2}+4\left( s_{13}s_{123}-s_{23}s_{12}\right) \right) , \end{aligned}$$
(3.33)

with \(s_{123}=s_{12}+s_{23}+s_{13}\).

\(\varvec{\gamma ^{*}\rightarrow q{\bar{q}}q'{\bar{q}}'}\)

$$\begin{aligned}&\left\langle {\mathcal {A}}_{q{\bar{q}}q'{\bar{q}}'}^{\left( 0\right) }\Big |{\mathcal {A}}_{q{\bar{q}}q'{\bar{q}}'}^{\left( 0\right) }\right\rangle =\frac{4e^{2}N_{c}C_{F}Q_{f}^{2}g_{S}^{4}}{s_{23}^{2}s_{123}^{2}s_{234}^{2}} \Big [4(d-2)Q^{2}s_{23}s_{123}s_{234}\nonumber \\&\quad -s_{1234}\Big ((d-2)s_{123}^{2}\left( (d-2)s_{23}^{2}\nonumber \right. \\&\quad \left. +4s_{34}s_{23}+4\left( s_{34}-s_{234}\right) {}^{2}\right) +2s_{234}s_{123}\nonumber \\&\quad +\left( s_{23}\left( ((d-10)d+20)s_{23}+2(d-2)s_{34}\right) \nonumber \right. \\&\quad \left. +2(d-2)s_{12}\left( s_{23}+2s_{34}-2s_{234}\right) \right) \nonumber \\&\quad +(d-2)s_{234}^{2}\left( (d-2)s_{23}^{2}+4s_{12}^{2}+4s_{23}s_{12}\right) \Big )\nonumber \\&\quad +s_{23}s_{123}s_{234}\Big (4(d-2)s_{12}^{2}\nonumber \\&\quad +4s_{12}\left( -(d-2)s_{123}+(d-4)\left( s_{234}-2s_{34}\right) +2s_{23}\right) \nonumber \\&\quad +4(d-2)s_{34}^{2}+(d-2)^{2}s_{123}^{2}\nonumber \\&\quad +4s_{34}\left( 2s_{23}-(d-2)s_{234}\right) \nonumber \\&\quad +(d-2)\left( (d-2)s_{234}^{2}\right. \left. +4s_{23}^{2}-4s_{234}s_{23}\right) \nonumber \\&\quad +2s_{123}\left( (d-4)\left( (d-4)s_{234}+2s_{34}\right) \nonumber \right. \\&\quad \left. -2(d-2)s_{23}\right) \Big )\Big ]. \end{aligned}$$
(3.34)

In this equation, we have not performed any additional collection of terms.

It is remarkable the simplicity of the full squared amplitudes displayed in (3.33) and (3.34). However, to keep track of the divergencies, within the ltd/fdu framework, one can consider individual Feynman diagrams and perform the match between real and virtual corrections. In fact, by kee** the ordering of the diagrams depicted in Fig. 5, one notices that, when squaring the amplitude, the interference terms account for contributions coming from the virtual diagrams. While the remaining diagrams account for the wave function corrections.

3.4.3 Map**s

As stated in the former sections, one of the distinctive features of the fdu approach is the real-virtual integrand-level combination through kinematical map**s. At NNLO, these map**s should relate the one- and two-loop amplitudes with the double-real emission terms. A similar strategy to the phase-space slicing method must be applied, but now there will be more singular regions, which might also overlap. In this way, a generic NNLO map** could be a complicated transformation with highly non-trivial dependencies.

However, when computing the NNLO QCD corrections proportional to \(N_f\) for the process \(\gamma ^* \rightarrow q {{\bar{q}}}\), a huge simplification takes place: only the double-virtual (i.e. two-loop) and the double-real emission contributes. Thus, we will need a transformation to generate a \(1 \rightarrow 4\) physical configuration starting from double-cuts, which are described by a physical \(1 \rightarrow 2\) process plus two additional on-shell momenta (i.e. the cut lines, \(q_i\) and \(q_j\)). A preliminar proposal for such map** is the following,

$$\begin{aligned}&p'_k = q_i ,\quad p'_l = q_j,\quad p'_1 = p_1 + \alpha _i q_i + \alpha _j q_j, \nonumber \\&p'_1 = p_1 + (1-\alpha _i) q_i + (1-\alpha _j) q_j , \end{aligned}$$
(3.35)

where momentum conservation is automatically fulfilled and the coefficients \(\alpha _i\) and \(\alpha _j\) must be adjusted to verify that \((p'_1)^2=p_1^2\) and \((p'_2)^2=p_2^2\). We can appreciate that the cut lines behave as real final state radiation, and we expect this transformation to link the IR singularities present in both contributions to achieve a fully local cancellation in four space-time dimensions.Footnote 11

3.5 Discussion

The ultimate goal of the ltd/fdu framework consists in achieving a fully local regularisation of both IR and UV singularities, thus leading to a four-dimensional representation of physical observables. The key observations are:

  • IR singularities cancel among virtual and real contributions, which is supported by the well-known KLN theorem;

  • and IR singularities inside the dualised virtual terms can be isolated into a compact region of the integration domain.

In this way, the singular regions in both contributions can be mapped to the same points, leading to a complete cancellation and skip** the need of introducing additional regularisation techniques (such as dreg). Alternatively, we can think that real emission is being used as an IR local counterterm for the dual contributions. Of course, renormalisation counterterms must be introduced, as well as potential initial-state radiation (ISR) subtraction terms.

Since the first proof-of-concept of the fdu framework, we have developed several new strategies to tackle the problem of obtaining integrable representations of IR-safe observables in four space-time dimensions. In particular, during last year, we have progressed a lot in understanding the location of IR and threshold singularities in the virtual amplitudes, as well as elucidating a novel dual representation through the application of nested residues. The path looks very promising to address some of the current limitations of our approach, namely a fully automated multi-loop local renormalisation and the cancellation of ISR singularities in a universal way. Other strategies, such as \(q_T\)-subtraction/resummation have shown to be perfectly adapted to attack these problems, although they still lack of locality. Thus, we believe that a conceptual combination of other methodologies might shed light to solve the current limitations to extend the ltd/fdu framework.