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

High precision tests of QCD without scale or scheme ambiguities
The 40th anniversary of the Brodsky–Lepage–Mackenzie method

Leonardo Di Giustino leonardo.digiustino@uninsubria.it Stanley J. Brodsky sjbth@slac.stanford.edu Philip G. Ratcliffe philip.ratcliffe@uninsubria.it Xing-Gang Wu wuxg@cqu.edu.cn Sheng-Quan Wang sqwang@alu.cqu.edu.cn Department of Science and High Technology, University of Insubria, via Valleggio 11, I-22100, Como, Italy INFN, Sezione di Milano–Bicocca, 20126 Milano, Italy SLAC National Accelerator Laboratory, Stanford University, Stanford, California 94039, USA Department of Physics, Chongqing University, Chongqing 401331, P.R. China Department of Physics, Guizhou Minzu University, Guiyang 550025, P.R. China
Abstract

A key issue in making precise predictions in QCD is the uncertainty in setting the renormalization scale μr\mu_{r} and thus determining the correct values of the QCD running coupling αs(μr)\alpha_{s}(\mu_{r}) at each order in the perturbative expansion of a QCD observable. It has often been conventional to simply set the renormalization scale to the typical scale of the process QQ and vary it in the range μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] in order to estimate the theoretical error. This is the practice of Conventional Scale Setting (CSS). The resulting CSS prediction will however depend on the theorist’s choice of renormalization scheme and the resulting pQCD series will diverge factorially. It will also disagree with renormalization scale setting used in QED and electroweak theory thus precluding grand unification. A solution to the renormalization scale-setting problem is offered by the Principle of Maximum Conformality (PMC), which provides a systematic way to eliminate the renormalization scale-and-scheme dependence in perturbative calculations. The PMC method has rigorous theoretical foundations, it satisfies Renormalization Group Invariance (RGI) and preserves all self-consistency conditions derived from the renormalization group. The PMC cancels the renormalon growth, reduces to the Gell-Mann–Low scheme in the Nc0N_{c}\to 0 Abelian limit and leads to scale- and scheme-invariant results. The PMC has now been successfully applied to many high-energy processes. In this article we summarize recent developments and results in solving the renormalization scale and scheme ambiguities in perturbative QCD. In particular, we present a recently developed method the PMC and its applications, comparing the results with CSS. The method preserves the property of renormalizable SU(N)/U(1) gauge theories defined as Intrinsic Conformality (iCF).

This property underlies the scale invariance of physical observables and leads to a remarkably efficient method to solve the conventional renormalization scale ambiguity at every order in pQCD.

This new method reflects the underlying conformal properties displayed by pQCD at NNLO, eliminates the scheme dependence of pQCD predictions and is consistent with the general properties of the PMC. A new method to identify conformal and β\beta-terms, which can be applied either to numerical or to theoretical calculations is also shown. We present results for the thrust and CC-parameter distributions in e+ee^{+}e^{-} annihilation showing errors and comparison with the CSS. We also show results for a recent innovative comparison between the CSS and the PMC applied to the thrust distribution investigating both the QCD conformal window and the QED Nc0N_{c}\rightarrow 0 limit. In order to determine the thrust distribution along the entire renormalization group flow from the highest energies to zero energy, we consider the number of flavors near the upper boundary of the conformal window. In this flavor-number regime the theory develops a perturbative infrared interacting fixed point. These results show that PMC leads to higher precision and introduces new interesting features in the PMC. In fact, this method preserves with continuity the position of the peak, showing perfect agreement with the experimental data already at NNLO.

We also show a detailed comparison of the PMC with the other PMC approaches: the multi-scale-setting approach (PMCm) and the single-scale-setting approach (PMCs) by comparing their predictions for three important fully integrated quantities Re+eR_{e^{+}e^{-}}, RτR_{\tau} and Γ(Hbb¯)\Gamma(H\to b\bar{b}) up to the four-loop accuracy.

keywords:
QCD , Renormalization Group , Event Shape Variables , Higgs , Principle of Maximum Conformality
journal: Progress in Particle and Nuclear Physics

1   Introduction

The renormalization scale and scheme ambiguities

The renormalization scale and scheme ambiguities are an important source of errors in many processes in perturbative QCD preventing precise theoretical predictions for both standard model (SM) and beyond standard model (BSM) physics. In principle, an infinite perturbative series is devoid of this issue, given the scheme and scale invariance of entire physical quantities [1, 2, 3, 4, 5]; in practice, perturbative corrections are known up to a certain order of accuracy and scale invariance is only approximated in truncated series, leading to scheme and scale ambiguities [6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18].

Although perturbative calculations for theoretical predictions are also affected by other sources of errors, e.g. the top- and Higgs-mass uncertainties and the strong-coupling αs(MZ)\alpha_{s}(M_{Z}) uncertainty, the renormalization scale and scheme ambiguities remain among of the main sources of errors. These scale and scheme ambiguities play an important role for predictions in many fundamental processes in perturbative QCD, also with respect to the other sources of uncertainties. Processes such as gluon fusion in Higgs production [19] or bottom-quark production [20], which are essential for the physics investigated at the Large Hadron Collider (LHC) and at future colliders, are all affected by these ambiguities. In the present era, high-precision predictions are crucial for both SM and BSM physics, in order to test the theory in all sectors and to enhance sensitivity to possible new physics (NP) at colliders.

At the moment two basic strategies exist to deal with this problem. On one hand, according to conventional practice or conventional scale setting (CSS), this problem cannot be avoided and is responsible for part of the theoretical errors. The simple CSS procedure starts from the scale and scheme invariance of a given observable, which translates into complete freedom for the choice of renormalization scale. According to common practice, a first evaluation of the physical observable is obtained by calculating perturbative corrections in a given scheme (commonly used are MS\rm MS or MS¯\overline{\rm MS}) and at an initial renormalization scale μr=μrinit\mu_{r}=\mu_{r}^{\rm init}. Thus, the renormalization scale μr\mu_{r} is set to one of the typical scales of a process, QQ, and errors are estimated by varying the scale over a range [Q/2,2Q][Q/2,2Q].

This method is claimed to evaluate uncalculated contributions from higher-order terms and that, due to the perturbative nature of the expansion, the introduction of higher-order corrections would reduce the scheme and scale ambiguities order by order. There is no doubt that the higher the order of the loop corrections calculated, the greater is the precision of the theoretical estimates in comparison with the experimental data, but we cannot determine a priori the level of accuracy necessary for the CSS to achieve the desired precision. And in the majority of cases at present only the NNLO corrections are available. Moreover, the divergent nature of the asymptotic perturbative series and the presence of factorially growing terms (i.e. renormalons [21, 22, 23] severely compromise the theoretical predictions.

However, even though this procedure may give an indication as to the level of conformality and convergence reached by the truncated expansion, it leads to a numerical evaluation of theoretical errors that is quite unsatisfactory and strongly dependent on the value of the chosen scale. Moreover, this method gives predictions with large theoretical uncertainties that are comparable with the calculated order correction; different choices of the renormalization scale may lead to very different results when higher-order corrections are included. For example, the NLO correction to W+3W+3\,jets with the BlackHat code [24] can range from negligible to extremely severe, depending on the choice of the particular renormalization scale. One may argue that the proper renormalization scale for a fixed-order prediction may be judged by comparing theoretical results with experimental data, but this method would be strongly process dependent and would compromise the predictivity of the pQCD approach.

Besides the complexity of the higher-order calculations and the slow convergence of the perturbative series, there are many critical points in the CSS method:

  • 1.

    In general, the proper renormalization scale value, QQ, is not known, nor the correct range over which the scale and scheme parameters should be varied in order to obtain the correct error estimate. In fact, in some processes there can be more than one typical momentum scale that may be taken as the renormalization scale according to the CSS procedure; for example, in processes involving heavy quarks typical scales are either the center-of-mass energy s\sqrt{s} or also the heavy-quark mass. Moreover, the idea of the typical momentum transfer as the renormalization scale only sets the order of magnitude of the scale, but does not indicate the optimal scale.

  • 2.

    No distinction is made among different sources of errors and their relative contributions; e.g. in addition to the errors due to scale-scheme uncertainties there are also errors from missing uncalculated higher-order terms. In such an approach theoretical uncertainties can become quite arbitrary and unreliable.

  • 3.

    The convergence of the perturbative series in QCD is affected by uncancelled large logarithms as well as by “renormalon” terms that diverge as (n!βinαsn+1)\left(n!\beta_{i}^{n}\alpha_{s}^{n+1}\right) at higher orders [21, 22]; this is known as the renormalon problem [23]. Such renormalon terms can give sizable contributions to the theoretical estimates, as shown in e+ee^{+}e^{-} annihilation, τ\tau decay, deeply inelastic scattering and hard processes involving heavy quarks. These terms are responsible for important corrections at higher orders also in the perturbative region, leading to different predictions according to the different choices of scale (as shown in Ref. [24]). Large logarithms on the other hand can be resummed using the resummation technique [25, 26, 27, 28, 29, 30, 31] and results are IR renormalon free. This does not help for the renormalization scale and scheme ambiguities, which still affect theoretical predictions with or without resummed large logarithms. In fact, as recently shown in Ref. [20] for the bb¯b\bar{b}-production cross-section at NNLO-order accuracy at hadron colliders, the CSS scale setting leads to theoretical uncertainties that are of the same order as the NNLO corrections 2030%\sim 20-30\%, taking as the typical momentum scale the bb-quark mass mb4.92m_{b}\sim 4.92 GeV.

  • 4.

    In the Abelian limit Nc0N_{c}\rightarrow 0 at fixed αem=CFαs\alpha_{em}=C_{F}\alpha_{s} with CF=(Nc21)/2NcC_{F}=\left(N_{c}^{2}-1\right)/2N_{c}, a QCD case effectively approaches the analogous QED case [32, 33]. Thus, to be self-consistent, any QCD scale-setting method should also be extendable to QED and results should be in agreement with the Gell-Mann–Low (GM–L) scheme. This is an important requirement also from the perspective of a grand unified theory (GUT), where only a single global method for setting the renormalization scale may be applied and then it can be considered as a good criterion for verifying if a scale setting is correct or not. CSS leads to incorrect results when applied to QED processes. In the GM–L scheme, the renormalization scale is set with no ambiguity to the virtuality of the exchanged photon/photons, which naturally sums an infinite set of vacuum polarization contributions into the running coupling. Thus, the CSS approach of varying the scale by a factor 2 is not applicable to QED, since the scale is already optimized.

  • 5.

    The large amount of forthcoming high-precision experimental data, produced especially by the running at high collision energy and high luminosity of the Large Hadronic Collider (LHC), will require more accurate and refined theoretical estimates. The CSS appears to be more a “guess”; its results are afflicted by large errors and the perturbative series converges poorly with or without large-logarithm resummation or renormalon contributions. Moreover, within such a framework, it is almost impossible to distinguish between SM and BSM signals and in many cases, improved higher-order calculations are not expected to be available in the short term.

To sum up, the conventional scale-setting method assigns an arbitrary range and an arbitrary systematic error to fixed-order perturbative calculations that greatly affects the predictions of pQCD. On the other hand, various strategies for optimization of the truncated expansion have been proposed. We point out that since its very beginning, several attempts have been performed to improve the renormalization procedure and different schemes have been introduced mainly to improve the convergence of the perturbative series. An example is the introduction of the MS¯\overline{\rm MS}, as suggested in Refs.[34, 35] and later the introduction of the momentum subtraction scheme (MOM) as discussed by Celmaster and Gonsalves in Refs. [36, 6]. The MS scheme, introduced in Ref. [37], is in fact quite arbitrary and related to the particular regularization scheme used, leading to a rather higher second-order contribution in the DIS process, by reabsorbing the scheme term ln(4πγE)\ln(4\pi-\gamma_{E}) into a redefinition of the QCD scale parameter Λ\Lambda, Bardeen et al. noticed a significant improvement of the convergence of the perturbative series. Later, the introduction of the MOM scheme, though being gauge dependent, has been shown to also include other scheme terms into Λ\Lambda, leading to a scheme which is less dependent on the regularization procedure and tending to a more "physical" scale and nearly to an "optimization procedure" [6]. A recent discussion on the gauge dependence of the MOM scheme has been given in Ref. [38]. We discuss these schemes in more detail in Sec. 2.3.

In general, an optimization or scale-setting procedure is considered reliable if it preserves important self-consistency requirements. All Renormalization Group properties, such as uniqueness, reflexivity, symmetry and transitivity should also be preserved by the scale-setting procedure in order to be generally applied [39]. A fundamental requirement is the scheme independence; other requirements can be suggested by known tested theories (such as QED), by the convergence behavior of the series in particular kinematic regions or by important phenomenological results.

The Principle of Minimal Sensitivity (PMS) is an optimization procedure proposed by Stevenson [11, 10, 9, 8] based on the assumption that, since observables should be independent of the particular RS and scale, their optimal perturbative approximations should be stable under small RS variations. The RS-scheme parameters β2,β3,\beta_{2},\beta_{3},\dots and the scale parameter Λ\Lambda (or the subtraction point μr\mu_{r}) are considered “unphysical” and independent variables; their values are thus set in order to minimize the sensitivity of the estimate to their small variations. This is essentially the core of Optimized Perturbation Theory (OPT) [9], based on the PMS procedure. The convergence of the perturbative expansion is improved by requiring its independence from the choice of RS and μ\mu. Optimization is implemented by identifying the RS-dependent parameters in the truncated series (βi\beta_{i} for 2in2\leq i\leq n and Λ\Lambda), with the request that the partial derivative of the perturbative expansion of the observable with respect to the RS-dependent and scale parameters vanishes. In practice, the PMS scale setting is designed to eliminate the remaining renormalization and scheme dependence in the truncated expansions of the perturbative series.

We would argue that this approach is based more on convergence rather than physical criteria. In particular, the PMS is a procedure that can be extended to higher order and it can be generally applied to calculations obtained in arbitrary initial renormalization schemes. Although this procedure leads to results that are suggested to be unique and scheme independent, it unfortunately violates important properties of the renormalization group, as shown in Ref. [40], such as reflexivity, symmetry, transitivity and also the existence and uniqueness of the optimal PMS renormalization scheme are not guaranteed since they are strictly related to the presence of maxima and minima.

Another optimization procedure, namely the Fastest Apparent Convergence (FAC) criterion was introduced by Grunberg and is based on the idea of effective charges. As pointed out by Grunberg [12, 13, 14], any perturbatively calculable physical quantity can be used to define an effective coupling, or “effective charge”, by entirely incorporating the radiative corrections into its definition. Effective charges can be defined from an observable starting from the assumption that the infinite series of a given quantity is scheme and scale invariant. The effective charge satisfies the same renormalization group equations as the usual coupling. Thus, the running behavior for both the effective coupling and the usual coupling are the same if their RG equations are calculated in the same renormalization scheme. This idea has been discussed in more detail in Refs. [41, 42].

An important suggestion is that all effective couplings defined in the same scheme satisfy the same RG equations. While different schemes or effective couplings would lead to different renormalization group equations. Hence, any effective coupling can be used as a reference for the particular renormalization procedure. In general, this method can be applied to any observable calculated in any RS, also in processes with large higher-order corrections. The FAC scale setting, as has been shown in Ref. [40], preserves the RG self-consistency requirements, although the FAC method can be considered more an optimization approach rather than a proper scale-setting procedure to extend order by order. FAC results depend sensitively on the quantity to which the method is applied. In general, when the NLO correction is large, the FAC proves to be a resummation of the most important higher-order corrections and thus a RG-improved perturbation theory is achieved.

PMS and FAC are procedures commonly in use for scale setting in perturbative QCD together with CSS and an introduction to these methods can be found in Refs. [40, 43]. However, as shown in Refs. [44], these optimization methods not only have the same difficulties of CSS, but they also lead to incorrect and unphysical results in particular kinematic regions.

A solution to the renormalization scale setting problem is offered by the Principle of Maximum Conformality (PMC) [45, 46, 47, 48]. This method is the generalization and extension of the original Brodsky-Lepage-Mackenzie (BLM) method [15] to all theories, to all orders and to all observables and it satisfies all the theoretical requirements of a reliable scale-setting procedure at once, leading to accurate and consistent results. The primary purpose of the PMC method is to solve the scale-setting ambiguity; it has been extended to all orders [49, 50] and it determines the correct running coupling and the correct momentum flow according to RGE invariance [51, 52]. This leads to results that are invariant with respect to the initial renormalization scale and in agreement with the requirement of scale invariance of observables in pQCD [40]. The approach provides a systematic method to eliminate renormalization scheme and scale ambiguities from first principles by absorbing the β\beta terms, which govern the behavior of the running coupling via the renormalization group equation. Thus, the divergent renormalon terms cancel, improving convergence of the perturbative QCD series. Furthermore, the resulting PMC predictions do not depend on the particular scheme used, thereby preserving the principles of renormalization group invariance [39, 51]. The PMC procedure is also consistent with the standard Gell-Mann–Low method in the Abelian limit, Nc0N_{c}\rightarrow 0 [32]. Moreover, in a theory unifying all forces (electromagnetic, weak and strong interactions), such as Grand Unified Theories, one cannot trivially apply a different scale-setting or analytic procedure to different sectors of the theory. The PMC offers the possibility to apply the same method to all sectors of a theory, starting from first principles, eliminating renormalon growth, scheme dependence, scale ambiguity and satisfying the QED Gell-Mann–Low scheme in the zero-color limit Nc0N_{c}\to 0.

PMC and schemes

We remark that the fundamental task of the PMC is to solve the renormalization scale and scheme ambiguities and in order to achieve this, it makes use of the RG-equations to reabsorb the nfn_{f}-terms that are related to the UV-divergent diagrams. The procedure of the PMC works with any initial definition of the scheme or scale for the running coupling, MS or MS¯\overline{\rm MS} or the ’t Hooft scheme are all equivalent, since the entire scheme dependence of the observable is reabsorbed into the running coupling and into the PMC scale in the final series. In particular, the lower order terms, β0,β1\beta_{0},\beta_{1}, of the β\beta-function are scheme independent; thus a scheme transformation cannot lead to a prescription for determining the scale. In fact, also in the case of the ’t Hooft scheme (as shown in Refs.[53, 54]) one can eliminate all scheme dependent βi\beta_{i} coefficients, but there is no prescription for the renormalization scale, which can be considered as the first scheme parameter. Moreover, what seems a good prescription for the running coupling is not necessarily a good prescription for the entire fixed-order calculated quantity. In fact, still considering the ’t Hooft case, the entire series of βi,i2\beta_{i},i\geq 2 is removed from the definition of the coupling, but in a fixed-order calculation these βi\beta_{i} coefficients would have an impact on the coefficients of the series at each order of calculation and on the convergence of the series. On the other hand, the PMC gives a prescription for fixing the scale and reabsorbing all scheme-dependent terms of a cross-section into the running coupling and into the PMC scale, exposing the perturbative series to a minimization/cancellation of the effects of the scale and scheme uncertainties. An important consequence of the PMC procedure is the RS-invariance of the resulting series. We refer to RS-scheme invariance as the invariance under the extended-renormalization group and its equations (xRGE)[9, 55, 56, 17]. It can be shown that the PMC procedure may be performed either way using the same xRGE111We will discuss this procedure in detail in a future work soon.. The xRGE show that different scheme definitions can be related at the lowest order by a scale transformation for the case of the minimal subtraction schemes (MS and MS¯\overline{\rm MS}) and the momentum space subtraction (MOM) scheme, but using the PMC, results would be scheme independent. A recent argument on the scheme dependence of the first conformal coefficient by Stevenson [57] is incorrect. The reason is that he assumes the v1v_{1} coefficient of the scheme to be totally free with respect to the structure of the color factors and that it can be reabsorbed into the Λ\Lambda parameter. There are several reasons why this certainly leads to wrong results. First, two different couplings in two different schemes at NLO can only be related by a scale transformation according to the RG group, i.e. by a shift of the type: β0log(μ12μ22)\beta_{0}\log(\frac{\mu_{1}^{2}}{\mu_{2}^{2}}) as shown also in Ref. [17]. In fact, given the scheme invariance of the β0,β1\beta_{0},\beta_{1}, the only parameter that can be varied at NLO is the renormalization scale μr\mu_{r}, which can have any real value or at most any value in the complex plane. Thus, the only scheme terms that can be reabsorbed into the scale (either the Λ\Lambda or the renormalization scale μr\mu_{r}) are those related to a shift of the scale using the standard RG group equations. This corresponds to the freedom of subtracting out any finite value together with the pole, in the renormalization procedure (e.g. MS and MS¯\overline{\rm MS}). If one takes the freedom to vary the scheme according to any other relation that does not correspond to a "proper" RG equation at NLO, this consequently modifies the structure of the color factors for the coupling and thus would obtain a wrong result. If one follows this misleading assumption, one can obtain any result outside of a given initial theory and this is certainly not the aim of the xRGE, which should preserve the invariance of the final result. Moreover, if one changes the color structure for the coupling, one should be aware that, to be consistent, also the structure of the entire fixed-order calculation should be varied accordingly; though we do not agree with this procedure since a change of the color structure correspond to a change of the initial SU(N) theory (e.g. from QCD to QED Ref.[32]). Thus, any relation among couplings in different schemes, which can have perturbative validity in QCD but which cannot be considered as "proper" xRG transformations, should be considered as matching relations among quantities defined in different approximations or obtained using different approaches. Also in this case, results can be improved using the PMC and the residual dependence on the particular implicit definition of the "scheme" can be suppressed perturbatively by adding higher-order calculations. It can be shown that also in this case the results obtained are scheme-independent (e.g. see Ref. [58]). One may point out that, though the LO and the NLO conformal coefficient are scheme invariant, the higher-order conformal coefficients can be scheme dependent, we can answer that in the worst case this scheme dependence for any of the PMCs procedures is highly suppressed. Another argument in Ref. [57] by Stevenson, which is based on the principle of minimum sensitivity (PMS), we hold to be incorrect. Since the PMS is based on the assumption that all the unknown higher-order terms give zero contribution to the pQCD series [9], its prediction directly breaks the standard renormalization group invariance [39, 51], its pQCD series does not have normal perturbative features [59] and can be treated as an effective prediction only when we know the series up to high enough orders and the conventional series has already shown good perturbative features [60]. On the other hand, the PMC respects all features of the renormalization group, and its prediction satisfies all the requirements of standard renormalization group invariance [39, 51, 58, 52]. Moreover, given that the PMC preserves the RG invariance, it is possible to define CSR - Commensurate Scale Relations[50] among the effective charges relating observables in different "schemes" preserving all the group properties. Applying the PMC and the CSRs, one can relate effective couplings, as also conformal coefficients, leading to scheme-independent results for the observables. We remark that in order to apply the PMC correctly, one should be able to distinguish among the nature of the different nfn_{f}-terms: whether they are related to the running of the coupling, to the running of masses or to UV-finite diagrams and in a deeper analysis also to the particular UV-divergent diagram (as discussed in Refs. [61, 48]). Once all nfn_{f}-terms have been associated with the correct diagram or parameter, conformal coefficients are RG invariant and match the coefficients of a conformal theory. Applications of the PMC to different quantities (see Ref. [62, 63]) have recently shown a direct improvement of theoretical predictions. Moreover, a deeper insight into the QCD coupling αs(Q)\alpha_{s}(Q) at all scales, including Q2=0Q^{2}=0, has been recently achieved (e.g. see Refs. [64, 65]), showing results consistent with the PMC.

PMC applications

So far, the PMC approach has been successfully applied to many high-energy processes, including Higgs boson production at the LHC [66], Higgs boson decays to γγ\gamma\gamma [67, 68], gggg and bb¯b\bar{b} [69, 70, 71, 72, 73, 63] processes, top-quark pair production at the LHC and Tevatron [74, 47, 75, 76, 77, 78, 79], decay process [80, 81], semihard processes based on the BFKL approach [82, 83, 84], electron–positron annihilation to hadrons [49, 50, 51], hadronic Z0Z^{0} boson decays [85, 86], the event shapes in electron–positron annihilation [87, 88, 89, 90, 91], the electroweak parameter ρ\rho [92, 93], Υ(1S)\Upsilon(1S) leptonic decay [94, 95], and charmonium production [96, 97, 98] and decay [99, 100, 101, 102]. As shown in Ref. [103], by using the PMC, one can obtain a smooth transition for the running behavior between Bjorken sum rule effective coupling αsg1(Q)\alpha_{s}^{g_{1}}(Q) in the V-scheme and the Light-Front Holographic QCD (LFHQCD) coupling, going from the perturbative to the non nonperturbative domain respectively. In addition, the PMC provides a possible solution to the BππB\to\pi\pi puzzle [104] and to the γγηc\gamma\gamma^{*}\rightarrow\eta_{c} puzzle [105].

In particular, with regard to the PMC application to top-quark pair hadroproduction, the resulting production cross-sections agree with precise experimental data, and the large discrepancies of the top-quark forward-backward asymmetries between SM estimations and experimental measurements are greatly reduced [47, 74, 75, 76, 77]. Recently, an improved QCD prediction for the top-quark decay process obtained by using the PMC has also been presented in Ref.[80].

Refer to caption
Figure 1.1: Summary of the top-quark pole masses, where the PMC result and previous determinations from collider measurements at different energies and different techniques are presented. The top-quark pole mass from the PDG [106] is also presented as the shaded band for reference.

A precise top-quark mass can be extracted by the comparison of precise PMC predictions of production cross-sections with experimental data measured at LHC [78, 79]. The determined top-quark pole mass mt=172.5±1.4m_{t}=172.5\pm 1.4 GeV, from the LHC measurement at s=13\sqrt{s}=13 TeV [79] agrees with the world average cited by the Particle Data Group (PDG) [106]. More explicitly, we present a summary of the top-quark pole masses in Fig.(1.1), where our PMC result and previous determinations from collider measurements at different energies and different techniques are presented. Owing to unknown higher-order contributions, this leads to two kinds of residual scale dependence for PMC predictions. However, these two residual scale dependencies are distinct from the conventional scale ambiguities, they are given by the unknown higher-order uncalculated contributions, while the scale dependence of the CSS is related more to the calculated orders. Thus, the residual scale dependence of the PMC results from the unknown higher-order terms and is an unavoidable intrinsic feature of the perturbative approach. It is to be noted that two of the three PMC ambiguities identified in Ref.[107] correspond exactly to two kinds of residual scale dependence. These two residual scale dependencies are highly suppressed; however, if the pQCD convergence of the perturbative series of either the PMC scale or the pQCD approximant is weak, such residual scale dependence could be significant. A recent discussion of the residual scale dependence of PMC predictions can also be found in the review [52].

Outline

In this review we display recent developments in solving the renormalization-scale and -scheme ambiguities based on the PMC and present some fundamental applications and results.

With respect to the previous reviews on the PMC (Refs. [40, 52]), in this review we focus on the recently developed method, namely the Infinite-Order Scale-Setting using the Principle of Maximum Conformality (PMC), showing first some of its applications and new features, and then a comparison of this method with the CSS and the other PMC approaches, such as the multi-scale (PMCm), and the single-scale PMC (PMCs), on the Event Shape Variables and on some crucial fully integrated quantities: Re+eR_{e^{+}e^{-}}, RτR_{\tau} and Γ(Hbb¯)\Gamma(H\to b\bar{b}).

More in detail: in Section I we give an introduction to the renormalization scale setting problem in QCD; in Section II we recall fundamental equations of the renormalization group and their extended version, starting from the renormalization procedure of the strong coupling αs(μ)\alpha_{s}(\mu) and its renormalization scale dependence; in Section III we summarize formulas and basic concepts of the PMC approach with its features and methods (PMCm, PMCs); in Section IV we introduce the newly developed method PMC and its new features; in Section V we show results for the application of the PMC to the event-shape variables: thrust and CC-parameter, we show particularly new features of the PMC performing interesting limits for thrust (IR conformal and QED limit) and comparing the results under the CSS and the PMC methods; in this section we also show a novel method to determine the strong coupling and its behavior αs(Q)\alpha_{s}(Q) over a wide range of scales, from a single experiment at a single scale, using the event-shape variable results; in Section VI we present a detailed comparison of the CSS, PMCs, PMCm and PMC methods tested on fundamental fully integrated quantities at 4 loops: Re+eR_{e^{+}e^{-}}, RτR_{\tau} and Γ(Hbb¯)\Gamma(H\to b\bar{b}); finally, in Section VII we summarize and discuss the results.

2   Renormalization Scale and Scheme invariance

The focus of this section is the strong coupling constant and its renormalization-scale and -scheme dependence. We summarize fundamental basic theoretical results and updated formulas regarding the renormalization group.

2.1   The Renormalization Group

QCD is a renormalizable theory, which means that the infinite number of ultraviolet (UV) singularities that arise in loop integration may be reabsorbed into a finite number of parameters entering the Lagrangian: the masses, coupling constant and fields.

The procedure starts from the assumption that the variables entering the Lagrangian are not the effective quantities measured in experiments, but are unknown functions affected by singularities. The origin of the ultraviolet singularities is often interpreted as a manifestation that a QFT is a low-energy effective theory of a more fundamental yet unknown theory. The use of regularization UV cut-offs shields the very short-distance domain, where the perturbative approach to QFT ceases to be valid.

Once the coupling has been renormalized to a measured value and at a given energy scale, the effective coupling is no longer sensitive to the ultraviolet (UV) cut-off, nor to any unknown phenomena arising beyond this scale. Thus, the scale dependence of the coupling can be well understood formally and phenomenologically. Actually, gauge theories are affected not only by UV, but also by infrared (IR) divergencies. The cancellation of the latter is guaranteed by the Kinoshita-Lee-Nauenberg (KLN) theorem [108, 109].

Considering first the Lagrangian of a massless theory, which is free of any particular scale parameter, in order to deal with these divergences a regularization procedure is introduced. Referring to the dimensional regularization procedure [37, 110, 111], one varies the dimension of the loop integration, D=42εD=4-2\varepsilon and introduces a scale μ\mu in order to restore the correct dimension of the coupling.

In order to determine the renormalized gauge coupling, we consider the quark-quark-gluon vertex and its loop corrections. UV-divergences arise from loop integration for higher-order contributions for both the external fields and the vertex. The renormalization constants of the vertices are related by Slavnov-Taylor identities, in particular:

Zαs1=(Z3Z2/Z1)2,Z^{-1}_{\alpha_{s}}=(\sqrt{Z_{3}}Z_{2}/Z_{1})^{2}, (2.1)

where ZαsZ_{\alpha_{s}} is the coupling renormalization constant, Z1(Q)Z_{1}\left(Q\right) is the vertex renormalization constant, Z3(Q)Z_{3}\left(Q\right) and Z2(Q)Z_{2}\left(Q\right) are:

Aa,μ=Z31/2(Q)Aa,μR(Q)andψ=Z21/2(Q)ψR(Q),A_{a,\mu}=Z_{3}^{1/2}\left(Q\right)A_{a,\mu}^{R}\left(Q\right)\quad\text{and}\quad\psi=Z_{2}^{1/2}\left(Q\right)\psi^{R}\left(Q\right),

the renormalization constants for gluon and quark fields respectively. The superscript RR indicates renormalized fields. The renormalization constants in dimensional regularization are given by:

Z1(Q)\displaystyle Z_{1}(Q) =\displaystyle= 1αs(Q)4π(Nc+CF)1ε\displaystyle 1-\frac{\alpha_{s}(Q)}{4\pi}\big{(}N_{c}+C_{F}\big{)}\frac{1}{\varepsilon} (2.2)
Z2(Q)\displaystyle Z_{2}(Q) =\displaystyle= 1αs(Q)4πCF1ε\displaystyle 1-\frac{\alpha_{s}(Q)}{4\pi}C_{F}\frac{1}{\varepsilon} (2.3)
Z3(Q)\displaystyle Z_{3}(Q) =\displaystyle= 1+αs(Q)4π(53Nc23Nf)1ε\displaystyle 1+\frac{\alpha_{s}(Q)}{4\pi}\left(\frac{5}{3}N_{c}-\frac{2}{3}N_{f}\right)\frac{1}{\varepsilon} (2.4)

where ε{\rm\varepsilon} is the regulator parameter for the UV-ultraviolet singularities. We have labelled NfN_{f} the number of flavors related to the UV singular diagrams.222Throughout the paper we use the following notation: nfn_{f} for the general number of active flavors, NfN_{f} to indicate the number of flavors related to the UV-divergent and NFN_{F} for the UV-finite diagrams. By substitution, we have that the UV divergence:

Zα(Q)=1αs(Q)4πβ01ε,Z_{\alpha}\left(Q\right)=1-\frac{\alpha_{s}\left(Q\right)}{4\pi}\beta_{0}\frac{1}{\varepsilon}, (2.5)

with

β0=112Nf3.\beta_{0}=11-\frac{2N_{f}}{3}. (2.6)

The singularities related to the UV poles are subtracted out by a redefinition of the coupling. In the MS\rm MS scheme, the renormalized strong coupling αs(Q)\alpha_{s}(Q) is related to the bare coupling αs¯\overline{\alpha_{s}} by:

αs¯=Q2εZα(Q)αs(Q).\overline{\alpha_{s}}=Q^{2\varepsilon}Z_{\alpha}\left(Q\right)\alpha_{s}\left(Q\right). (2.7)

In the minimal subtraction scheme (MS\rm MS) only the pole 1/ε1/\varepsilon related to the UV singularity is subtracted out. A more suitable scheme is MS¯\overline{\rm MS} [35, 112, 113], where the constant term ln(4π)γE\ln(4\pi)-\gamma_{E} is also subtracted out. Different schemes can also be related by scale redefinition, e.g. μ24πμ2eγE\mu^{2}\rightarrow 4\pi\mu^{2}e^{-\gamma_{E}}. We remark that the renormalization procedure leads to a unique renormalization constant ZαsZg2Z_{\alpha_{s}}\equiv Z^{2}_{g} for the strong coupling. In fact, the other renormalization constants, such as Z1VZ_{1}^{V}, V(3g,4g,ccg,qqg)V\in(3g,4g,ccg,qqg), i.e. the renormalization constants for 3-gluon, 4-gluon, ghost-ghost-gluon, quark-quark-gluon vertex respectively, are related to it via the Slavnov–Taylor identities [114]:

Zg\displaystyle Z_{g} =\displaystyle= Z13g(Z3)3/2,\displaystyle Z_{1}^{3g}\left(Z_{3}\right)^{-3/2}, (2.8)
Zg\displaystyle Z_{g} =\displaystyle= Z14g(Z3)1,\displaystyle\sqrt{Z_{1}^{4g}}\left(Z_{3}\right)^{-1}, (2.9)
Zg\displaystyle Z_{g} =\displaystyle= Z1ccg(Z3)1/2(Z3c)1,\displaystyle Z_{1}^{ccg}\left(Z_{3}\right)^{-1/2}\left(Z_{3}^{c}\right)^{-1}, (2.10)
Zg\displaystyle Z_{g} =\displaystyle= Z1ψψg(Z3)1/2(Z2)1.\displaystyle Z_{1}^{\psi\psi g}\left(Z_{3}\right)^{-1/2}\left(Z_{2}\right)^{-1}. (2.11)

Thus, the renormalization procedure depends both on the particular choice of scheme and on the subtraction point μ\mu. Hence, even though there are no dimensionful parameters in the initial bare Lagrangian, a mass scale μ\mu is acquired during the renormalization procedure. The emergence of μ\mu from a Lagrangian with no explicit scale is called dimensional transmutation [115]. The value of μ\mu is arbitrary and is the momentum at which the UV divergences are subtracted out. Hence μ\mu is called the subtraction point or renormalization scale. Thus, the definition of the renormalized coupling αsMS¯(μ)\alpha_{s}^{\overline{\rm MS}}(\mu) depends at the same time on both the chosen scheme, MS¯\overline{\rm MS} and the renormalization scale μ\mu.

Renormalization scale invariance is recovered by introducing the Renormalization Group Equations. The scale dependence of the coupling can be determined by considering that the bare coupling αs¯\overline{\alpha_{s}} and renormalized couplings, αs\alpha_{s}, at different scales are related by:

αs¯=Q2εZα(Q)αs(Q)=μ2εZα(μ)αs(μ),\overline{\alpha_{s}}=Q^{2\varepsilon}Z_{\alpha}\left(Q\right)\alpha_{s}\left(Q\right)=\mu^{2\varepsilon}Z_{\alpha}\left(\mu\right)\alpha_{s}\left(\mu\right), (2.12)

where ε\varepsilon is the regularization parameter, integrals are carried out in 42ε4-2\varepsilon dimensions and the UV divergences are regularized to 1/ε1/\varepsilon poles. The ZiZ_{i} are constructed as functions of 1/ε1/\varepsilon, such that they cancel all 1/ε1/\varepsilon poles. From Eq. (2.12) we obtain the relation from two different couplings at two different scales:

αs(Q)=𝒵α(Q,μ)αs(μ),\alpha_{s}\left(Q\right)=\mathcal{Z}_{\alpha}\left(Q,\mu\right)\alpha_{s}\left(\mu\right), (2.13)

with 𝒵α(Q,μ)(μ2ε/Q2ε)[Zα(Q)/Zα(μ)]\mathcal{Z}_{\alpha}\left(Q,\mu\right)\equiv\left(\mu^{2\varepsilon}/Q^{2\varepsilon}\right)\left[Z_{\alpha}\left(Q\right)/Z_{\alpha}\left(\mu\right)\right]. The 𝒵α\mathcal{Z}_{\alpha} then clearly form a group with a composition law

𝒵α(Q,μ)=𝒵α(Q,μ0)𝒵α(μ0,μ),\mathcal{Z}_{\alpha}\left(Q,\mu\right)=\mathcal{Z}_{\alpha}\left(Q,\mu_{0}\right)\mathcal{Z}_{\alpha}\left(\mu_{0},\mu\right), (2.14)

a unity element 𝒵α(Q,Q)=1\mathcal{Z}_{\alpha}\left(Q,Q\right)=1 and an inverse: 𝒵α(Q,μ)=𝒵α1(μ,Q)\mathcal{Z}_{\alpha}\left(Q,\mu\right)=\mathcal{Z}_{\alpha}^{-1}\left(\mu,Q\right). The fundamental properties of the Renormalization Group are: reflexivity, symmetry and transitivity. Thus, the scale invariance of a given perturbatively calculated quantity is recovered by the invariance of the theory under the Renormalization Group Equations (RGE).

2.2   The evolution of αs(μ)\alpha_{s}(\mu) in perturbative QCD

As shown in the previous section the renormalization procedure is not void of ambiguities. The subtraction of the singularities depends on the subtraction point or renormalization scale μ\mu and on the renormalization scheme (RS). Physical observables cannot depend on the particular scheme or scale, given that the theory stems from a conformal Lagrangian. This implies that scale invariance must be recovered imposing the invariance of the renormalized theory under the renormalization group equation (RGE). In this section we discuss the dependence of the renormalized coupling αs(Q)\alpha_{s}(Q) on the scale QQ. As shown in QED by Gell-Mann and Low, this dependence can be described introducing the β\beta-function given by:

14πdαs(Q)dlogQ2=β(αs)\frac{1}{4\pi}\frac{d\alpha_{s}(Q)}{d\log Q^{2}}=\beta(\alpha_{s}) (2.15)

and

β(αs)=(αs4π)2n=0(αs4π)nβn.\beta\left(\alpha_{s}\right)=-\left(\frac{\alpha_{s}}{4\pi}\right)^{2}\sum_{n=0}\left(\frac{\alpha_{s}}{4\pi}\right)^{n}\beta_{n}. (2.16)

Neglecting quark masses, the first two β\beta-terms are RS independent and have been calculated in Refs. [116, 117, 118, 119, 120] for the MS¯\overline{\rm MS} scheme:

β0\displaystyle\beta_{0} =\displaystyle= 113CA43TRNf,\displaystyle\frac{11}{3}C_{A}\!-\!\frac{4}{3}T_{R}N_{f}, (2.17)
β1\displaystyle\beta_{1} =\displaystyle= 343CA24(53CA+CF)TRNf\displaystyle\frac{34}{3}C_{A}^{2}\!-\!4\left(\frac{5}{3}C_{A}\!+\!C_{F}\right)T_{R}N_{f} (2.18)

where CF=(Nc21)2NcC_{F}=\frac{\left(N_{c}^{2}-1\right)}{2N_{c}}, CA=NcC_{A}=N_{c} and TR=1/2T_{R}=1/2 are the color factors for the SU(3){\rm SU(3)} gauge group [121]. At higher loops βi,\beta_{i}, i2i\geq 2 are scheme dependent; results for MS¯\overline{\rm MS} are given in Ref. [122]

β2\displaystyle\beta_{2} =\displaystyle= 28572503318Nf+32554Nf2,\displaystyle\frac{2857}{2}-\frac{5033}{18}N_{f}+\frac{325}{54}N_{f}^{2}, (2.19)

in Ref. [123]

β3\displaystyle\beta_{3} =\displaystyle= (1497536+3564ζ3)(1078361162+650827ζ3)Nf\displaystyle\left(\frac{149753}{6}+3564\zeta_{3}\right)-\left(\frac{1078361}{162}+\frac{6508}{27}\zeta_{3}\right)N_{f} (2.20)
+(50065162+647281ζ3)Nf2+1093729Nf3,\displaystyle\quad+\left(\frac{50065}{162}+\frac{6472}{81}\zeta_{3}\right)N_{f}^{2}+\frac{1093}{729}N_{f}^{3},

and in Ref. [124]

β4\displaystyle\beta_{4} =\displaystyle= {815745516+6218852ζ3882092ζ4288090ζ5\displaystyle\left\{\frac{8157455}{16}+\frac{621885}{2}\zeta_{3}-\frac{88209}{2}\zeta_{4}-288090\zeta_{5}\right. (2.21)
+Nf[3364608131944481116481ζ3+339356ζ4+135899527ζ5]\displaystyle\quad+N_{f}\left[-\frac{336460813}{1944}-\frac{4811164}{81}\zeta_{3}+\frac{33935}{6}\zeta_{4}+\frac{1358995}{27}\zeta_{5}\right]
+Nf2[259609131944+69853181ζ3105269ζ438176081ζ5]\displaystyle\quad+N_{f}^{2}\left[\frac{25960913}{1944}+\frac{698531}{81}\zeta_{3}-\frac{10526}{9}\zeta_{4}-\frac{381760}{81}\zeta_{5}\right]
+Nf3[630559583248722243ζ3+161827ζ4+4609ζ5]+Nf4[1205291615281ζ3]},\displaystyle\quad+N_{f}^{3}\left[-\frac{630559}{5832}-\frac{48722}{243}\zeta_{3}+\frac{1618}{27}\zeta_{4}+\frac{460}{9}\zeta_{5}\right]+\left.N_{f}^{4}\left[\frac{1205}{2916}-\frac{152}{81}\zeta_{3}\right]\right\},

with ζ31.20206\zeta_{3}\simeq 1.20206, ζ41.08232\zeta_{4}\simeq 1.08232 and ζ51.03693\zeta_{5}\simeq 1.03693, the Riemann zeta function. Given the renormalizability of QCD, new UV singularities arising at higher orders can be cancelled by redefinition of the same parameter, i.e. the strong coupling. This procedure leads to the renormalization constant:

Za(μ)\displaystyle Z_{a}(\mu) =\displaystyle= 1β0ϵa+(β02ϵ2β12ϵ)a2(β03ϵ376β0β1ϵ2+β23ϵ)a3\displaystyle 1-\frac{\beta_{0}}{\epsilon}a+\left(\frac{\beta_{0}^{2}}{\epsilon^{2}}-\frac{\beta_{1}}{2\epsilon}\right)a^{2}-\left(\frac{\beta_{0}^{3}}{\epsilon^{3}}-\frac{7}{6}\frac{\beta_{0}\beta_{1}}{\epsilon^{2}}+\frac{\beta_{2}}{3\epsilon}\right)a^{3} (2.22)
+(β04ϵ423β1β0212ϵ3+5β2β06ϵ2+3β128ϵ2β34ϵ)a4+,\displaystyle\qquad+\left(\frac{\beta_{0}^{4}}{\epsilon^{4}}-\frac{23\beta_{1}\beta_{0}^{2}}{12\epsilon^{3}}+\frac{5\beta_{2}\beta_{0}}{6\epsilon^{2}}+\frac{3\beta_{1}^{2}}{8\epsilon^{2}}-\frac{\beta_{3}}{4\epsilon}\right)a^{4}+\cdots,

where a=αs(μ)/(4π)a=\alpha_{s}(\mu)/(4\pi) in the MS\rm MS scheme. Given the arbitrariness of the subtraction procedure of also including part of the finite contributions (e.g. the constant [ln(4π)γE][\ln(4\pi)-\gamma_{E}] in MS¯\overline{\rm MS}), there is an inherent ambiguity for these terms that translates into the RS dependence. In order to solve any truncated version of Eq. (2.15), this being a first order differential equation, we need an initial value of αs\alpha_{s} at a given energy scale αs(μ0)\alpha_{s}(\mu_{0}). For this purpose, we set the initial scale μ0=MZ\mu_{0}=M_{Z} the Z0Z^{0} mass, with the value αs(MZ)\alpha_{s}(M_{Z}) being determined phenomenologically. In QCD the number of colors NcN_{c} is set to 3, while NfN_{f}, i.e. the number of active flavors, varies with energy scale across quark thresholds.

2.2.1   One-loop result and asymptotic freedom

When all quark masses are set to zero, two physical parameters dictate the dynamics of the theory and these are the numbers of flavors NfN_{f} and colors NcN_{c}. In this section we determine the exact analytical solution to the truncated Eq. (2.15). Considering the formula:

αs(μ0)αs(μ)14πdαsβ(αs)=μ02μ2dQ2Q2,\int_{\alpha_{s}(\mu_{0})}^{\alpha_{s}(\mu)}\frac{1}{4\pi}\frac{d\alpha_{s}}{\beta(\alpha_{s})}=-\int_{\mu_{0}^{2}}^{\mu^{2}}\frac{dQ^{2}}{Q^{2}}, (2.23)

and retaining only the first term:

Q2αs2αsQ2=14πβ0\frac{Q^{2}}{\alpha_{s}^{2}}\frac{\partial\alpha_{s}}{\partial Q^{2}}=-\frac{1}{4\pi}\beta_{0} (2.24)

we achieve the solution for the coupling:

4παs(μ0)4παs(Q)=β0ln(μ02Q2).\frac{4\pi}{\alpha_{s}\left(\mu_{0}\right)}-\frac{4\pi}{\alpha_{s}\left(Q\right)}=\beta_{0}\ln\left(\frac{\mu_{0}^{2}}{Q^{2}}\right). (2.25)

This solution can be given in the explicit form:

αs(Q)=αs(μ0)1+β0αs(μ0)4πln(Q2/μ02).\alpha_{s}(Q)=\frac{\alpha_{s}(\mu_{0})}{1+\beta_{0}\frac{\alpha_{s}(\mu_{0})}{4\pi}\ln(Q^{2}/\mu_{0}^{2})}. (2.26)

This solution relates one known (measured value) of the coupling at a given scale μ0\mu_{0} with an unknown value αs(Q)\alpha_{s}(Q). More conveniently, the solution can be given introducing the QCD scale parameter Λ\Lambda. At β0\beta_{0} order, this is defined as:

Λ2μ2e4πβ0αs(μ),\Lambda^{2}\equiv\mu^{2}e^{-\frac{4\pi}{\beta_{0}\alpha_{s}\left(\mu\right)}}, (2.27)

which yields the familiar one-loop solution:

αs(Q)=4πβ0ln(Q2/Λ2).\alpha_{s}\left(Q\right)=\frac{4\pi}{\beta_{0}\ln\left(Q^{2}/\Lambda^{2}\right)}.

Already at the one-loop level one can distinguish two regimes of the theory. For the number of flavors larger than 11Nc/211N_{c}/2 (i.e. the zero of the β0\beta_{0} coefficient) the theory possesses an infrared noninteracting fixed point and at low energies the theory is known as non-Abelian quantum electrodynamics (non-Abelian QED). The high-energy behavior of the theory is uncertain, it depends on the number of active flavors and there is the possibility that it could develop a critical number of flavors above which the theory reaches an UV fixed point [125] and therefore becomes safe. When the number of flavors is less than 11Nc/211N_{c}/2, the noninteracting fixed point becomes UV in nature and then we say that the theory is asymptotically free.

It is straightforward to check the asymptotic limit of the coupling in the deep-UV region:

limsαs(s)=0.\lim_{s\rightarrow\infty}\alpha_{s}(s)=0. (2.28)

This result is known as asymptotic freedom and it is the outstanding result that has justified QCD as the most accredited candidate for the theory of strong interactions. On the other hand, we have that the perturbative coupling diverges at the scale Λ(200300)MeV\Lambda\sim(200-300)\,{\rm MeV}. This is sometimes referred to as the Landau ghost pole to indicate the presence of a singularity in the coupling that is actually unphysical and implies the breakdown of the perturbative regime. This itself is not an explanation for confinement, though it might indicate its presence. When the coupling becomes too large, the use of a nonperturbative approach to QCD is mandatory in order to obtain reliable results. We remark that the scale parameter Λ\Lambda is RS dependent and its definition depends on the order of accuracy of the coupling αs(Q)\alpha_{s}(Q). Considering that the solution αs\alpha_{s} at order β0\beta_{0} or β1\beta_{1} is universal, the definition of Λ\Lambda at the first two orders is usually preferred, i.e. Λ\Lambda given at 1-loop by Eq. (2.27) or at 2-loops (see later) by Eq. (2.33).

2.2.2   Two-loop solution and the perturbative conformal window

In order to determine the solution for the strong coupling αs\alpha_{s} at NNLO, it is convenient to introduce the following notation: x(μ)αs(μ)2πx(\mu)\equiv\frac{\alpha_{s}(\mu)}{2\pi}, t=log(μ2/μ02)t=\log(\mu^{2}/\mu_{0}^{2}), B=12β0B=\frac{1}{2}\beta_{0} and C=12β1β0C=\frac{1}{2}\frac{\beta_{1}}{\beta_{0}}, x1Cx^{*}\equiv-\frac{1}{C}. The truncated NNLO approximation of the Eq. (2.15) leads to the differential equation:

dxdt=Bx2(1+Cx)\frac{dx}{dt}=-Bx^{2}(1+Cx) (2.29)

An implicit solution of Eq. (2.29) is given by the Lambert W(z)W(z) function:

WeW=zWe^{W}=z (2.30)

with: W=(xx1)W=\left(\frac{x^{*}}{x}-1\right). The general solution for the coupling is:

x\displaystyle x =\displaystyle= x1+W,\displaystyle\frac{x^{*}}{1+W}, (2.31)
z\displaystyle z =\displaystyle= exx01(xx01)(μ2μ02)xB.\displaystyle e^{\frac{x^{*}}{x_{0}}-1}\left(\frac{x^{*}}{x_{0}}-1\right)\left(\frac{\mu^{2}}{\mu_{0}^{2}}\right)^{x^{*}B}. (2.32)

Here we shall discuss the solutions to Eq. (2.29) with respect to the particular initial phenomenological value x0αs(MZ)/(2π)=0.01877±0.00014x_{0}\equiv\alpha_{s}(M_{Z})/(2\pi)=0.01877\pm 0.00014 given by the coupling determined at the Z0Z^{0} mass scale [126].

The signs of β0,β1\beta_{0},\beta_{1} and consequently of B,xB,x^{*}, depend on the values of the Nc,NfN_{c},N_{f}, since the number NcN_{c} is set by the SU(Nc)(N_{c}) theory, we discuss the possible regions varying only the number of flavors NfN_{f}. We point out that different regions are defined by the signs of the β0,β1\beta_{0},\beta_{1}, which have zeros in Nf0¯=112Nc\bar{N_{f}^{0}}=\frac{11}{2}N_{c}, Nf1¯=34Nc313Nc23\bar{N_{f}^{1}}=\frac{34N_{c}^{3}}{13N_{c}^{2}-3} respectively with Nf0¯>Nf1¯\bar{N_{f}^{0}}>\bar{N_{f}^{1}}. In the range Nf<Nf1¯N_{f}<\bar{N_{f}^{1}} and Nf>Nf0¯N_{f}>\bar{N_{f}^{0}} we have B>0B>0, C>0C>0 and the physical solution is given by the W1W_{-1} branch, while for Nf1¯<Nf<Nf0¯\bar{N_{f}^{1}}<N_{f}<\bar{N_{f}^{0}} the solution for the strong coupling is given by the W0W_{0} branch. By introducing the phenomenological value x0x_{0}, we define a restricted range for the IR fixed point discussed by Banks and Zaks [127]. Given the value N¯f=x1(x0)=15.222±0.009\bar{N}_{f}=x^{*-1}(x_{0})=15.222\pm 0.009, we have that in the range 34Nc313Nc23<Nf<N¯f\frac{34N_{c}^{3}}{13N_{c}^{2}-3}<N_{f}<\bar{N}_{f} the β\beta-function has both a UV and an IR fixed point, while for Nf>N¯fN_{f}>\bar{N}_{f} we no longer have the asymptotically free UV behavior. The two-dimensional region in the number of flavors and colors where asymptotically free QCD develops an IR interacting fixed point is colloquially known as the conformal window of pQCD.

Thus, the actual physical range of a conformal window for pQCD is given by 34Nc313Nc23<Nf<N¯f\frac{34N_{c}^{3}}{13N_{c}^{2}-3}<N_{f}<\bar{N}_{f}. The behavior of the coupling is shown in Fig. 2.1. In the IR region the strong coupling approaches the IR finite limit, xx^{*}, in the case of values of NfN_{f} within the conformal window (e.g. black dashed curve of Fig. 2.1), while it diverges at

Λ=μ0(1+|x|x0)12B|x|e12Bx0\Lambda=\mu_{0}\left(1+\frac{|x^{*}|}{x_{0}}\right)^{\frac{1}{2B|x^{*}|}}e^{-\frac{1}{2Bx_{0}}} (2.33)

outside the conformal window given the solution for the coupling with W1W_{-1} (e.g. the solid red curve of Fig. 2.1). The solution of the NNLO equation for the case B>0,C>0B>0,C>0, i.e. Nf<34Nc313Nc23N_{f}<\frac{34N_{c}^{3}}{13N_{c}^{2}-3}, can also be given using the standard QCD scale parameter Λ\Lambda of Eq. (2.33),

x\displaystyle x =\displaystyle= x1+W1,\displaystyle\frac{x^{*}}{1+W_{-1}}, (2.34)
z\displaystyle z =\displaystyle= 1e(μ2Λ2)xB.\displaystyle-\frac{1}{e}\left(\frac{\mu^{2}}{\Lambda^{2}}\right)^{x^{*}B}. (2.35)

Different solutions can be achieved using different schemes, i.e. different definitions of the scale parameter Λ\Lambda [128]. We stress that the presence of a Landau “ghost” pole in the strong coupling is only an effect of the breaking of the perturbative regime, including nonperturbative contributions, or using nonperturbative QCD, a finite limit is expected at any NfN_{f} [43]. Both solutions have the correct UV asymptotically free behavior. In particular, for the case N¯f<Nf<112Nc\bar{N}_{f}<N_{f}<\frac{11}{2}N_{c}, we have a negative zz, a negative CC and a multi-valued solution, one real and the other imaginary, actually only one (the real) is acceptable given the initial conditions, but this solution is not asymptotically free. We thus restrict our analysis to the range Nf<N¯fN_{f}<\bar{N}_{f}, where we have the correct UV behavior. In general, IR and UV fixed points of the β\beta-function can also be determined at different values of the number of colors NcN_{c} (different gauge group SU(N)SU(N)) and NfN_{f} also extending this analysis to other gauge theories [129].

Refer to caption
Figure 2.1: The strong running coupling αs(μ)\alpha_{s}(\mu) for Nf=12N_{f}=12 (blue dashed) and for Nf=5N_{f}=5 (solid red). [90]

2.2.3   αs\alpha_{s} at higher loops

The 3-loop truncated RG equation 2.15, written using the same normalization as Eq. (2.29) is given by

β(x)=dxdt=Bx2(1+Cx+C2x2),\beta(x)=\frac{dx}{dt}=-Bx^{2}\left(1+Cx+C_{2}x^{2}\right), (2.36)

with C2=β24β0C_{2}=\frac{\beta_{2}}{4\beta_{0}}.

A straightforward integration of this equation would be hard to invert, as shown in Ref. [128], it is more convenient to extend the approach of the previous section by using the Padé Approximant Approach (PAA) [130, 131, 132]. The Padé Approximant of a given quantity calculated perturbatively in QCD up to order nn, i.e. of the series:

S(x)=x(1+r1x+r2x2++rnxn)S(x)=x\left(1+r_{1}x+r_{2}x^{2}+\cdots+r_{n}x^{n}\right) (2.37)

is defined as the rational function

x[N/M]=x1+a1x++aNxN1+b1x++bMxM,x[N/M]=S+x𝒪(xN+M+1),x[N/M]=x\,\frac{1+a_{1}x+\ldots+a_{N}x^{N}}{1+b_{1}x+\ldots+b_{M}x^{M}},\quad x[N/M]=S+x\mathcal{O}\left(x^{N+M+1}\right), (2.38)

whose Taylor expansion up to order N+M=nN+M=n is identical to the original truncated series. The use of the PA makes the integration of Eq. (2.36) straightforward. PA’s may also be used either to predict the next term of a given perturbative expansion, called a Padé Approximant prediction (PAP), or to estimate the sum of the entire series, called Padé Summation. Features of the PA are described in Ref. [133].

The Padé Approximant (x2[1/1]x^{2}[1/1]) of the 3-loop β\beta-function is given by

βPA(x)=Bx21+[C(C2/C)]x1(C2/C)x,\beta_{\rm PA}(x)=-Bx^{2}\,\frac{1+\left[C-\left(C_{2}/C\right)\right]x}{1-\left(C_{2}/C\right)x}, (2.39)

which leads to the solution:

Bln(Q2/Λ2)=1xCln[1x+CC2C]B\ln\left(Q^{2}/\Lambda^{2}\right)=\frac{1}{x}-C\ln\left[\frac{1}{x}+C-\frac{C_{2}}{C}\right] (2.40)

and finally,

x(Q2)\displaystyle x\left(Q^{2}\right) =\displaystyle= 1C11(C2/C2)+W(z),\displaystyle-\frac{1}{C}\frac{1}{1-\left(C_{2}/C^{2}\right)+W(z)}, (2.41)
z\displaystyle z =\displaystyle= 1Cexp[1+(C2/C2)Bt/C],\displaystyle-\frac{1}{C}\exp\left[-1+\left(C_{2}/C^{2}\right)-Bt/C\right], (2.42)

the sign of CC determines the sign of zz and also the physically relevant branches of the Lambert function W(z)W(z): for C>0,z<0C>0,z<0 and the physical branch is W1(z)W_{-1}(z), taking real values in the range (,1)(-\infty,-1), while for C<0,z>0C<0,z>0 and the physical branch is given by the W0(z)W_{0}(z), taking real values in the range (0,)(0,\infty).

We notice that the only significant difference between the 3-loop solution and the 2-loop solution (2.32) is in the solution x(Q2)x\left(Q^{2}\right). This is because the difference in the definition of zz can be reabsorbed into an appropriate redefinition of the scale parameter:

Λ2Λ~2=Λ2eC2/(BC).\Lambda^{2}\longrightarrow\tilde{\Lambda}^{2}=\Lambda^{2}e^{C_{2}/\left(BC\right)}.

For orders up to β4\beta_{4}, an approximate analytical solution is obtained integrating Eq. (2.15):

lnμ2Λ2\displaystyle\ln\frac{\mu^{2}}{\Lambda^{2}} =\displaystyle= daβ(a)=1β0[1a+b1lna+a(b12+b2)+a2(b132b1b2+b32)\displaystyle\int\frac{da}{\beta(a)}~{}~{}=~{}~{}\frac{1}{\beta_{0}}\left[\frac{1}{a}+b_{1}\ln a+a\left(-b_{1}^{2}+b_{2}\right)+a^{2}\left(\frac{b_{1}^{3}}{2}-b_{1}b_{2}+\frac{b_{3}}{2}\right)\right. (2.43)
+a3(b143+b12b2b22323b1b3+b43)+O(a4)]+C\displaystyle\hskip 80.00012pt\hbox{}+a^{3}\left(-\frac{b_{1}^{4}}{3}+b_{1}^{2}b_{2}-\frac{b_{2}^{2}}{3}-\left.\frac{2}{3}b_{1}b_{3}+\frac{b_{4}}{3}\right)+O\left(a^{4}\right)\right]+C

where a=αs(μ)/(4π)a=\alpha_{s}(\mu)/(4\pi) and bNβN/β0b_{N}\equiv\beta_{N}/\beta_{0}, (N=1,,4)(N=1,\dots,4) and performing the inversion of the last formula by iteration as shown in Ref. [134], achieving the final result of the coupling at five-loop accuracy:

a\displaystyle a =\displaystyle= 1β0Lb1lnL(β0L)2+1(β0L)3[b12(ln2LlnL1)+b2]\displaystyle\frac{1}{\beta_{0}L}-\frac{b_{1}\ln L}{\left(\beta_{0}L\right)^{2}}+\frac{1}{\left(\beta_{0}L\right)^{3}}\left[b_{1}^{2}\left(\ln^{2}L-\ln L-1\right)+b_{2}\right] (2.44)
+1(β0L)4[b13(ln3L+52ln2L+2lnL12)3b1b2lnL+b32]\displaystyle+\frac{1}{\left(\beta_{0}L\right)^{4}}\left[b_{1}^{3}\left(-\ln^{3}L+\frac{5}{2}\ln^{2}L+2\ln L-\frac{1}{2}\right)-3b_{1}b_{2}\ln L+\frac{b_{3}}{2}\right]
+1(β0L)5[b14(ln4L133ln3L32ln2L+4lnL+76)+3b12b2(2ln2LlnL1)\displaystyle+\frac{1}{\left(\beta_{0}L\right)^{5}}\left[b_{1}^{4}\left(\ln^{4}L-\frac{13}{3}\ln^{3}L\right.-\frac{3}{2}\ln^{2}L+4\ln L+\frac{7}{6}\right)+3b_{1}^{2}b_{2}\left(2\ln^{2}L-\ln L-1\right)
b1b3(2lnL+16)+53b22+b43]+O(1L6).\displaystyle\hskip 60.00009pt-\left.b_{1}b_{3}\left(2\ln L+\frac{1}{6}\right)+\frac{5}{3}b_{2}^{2}+\frac{b_{4}}{3}\right]+O\left(\frac{1}{L^{6}}\right).

where L=ln(μ2/Λ2)L=\ln(\mu^{2}/\Lambda^{2}). The same definition of the Λ\Lambda scale given in Eq. (2.33) has been used for the MS¯\overline{\rm MS} scheme, which leads to setting the constant C=(b1/β0)ln(β0)C=\left(b_{1}/\beta_{0}\right)\ln(\beta_{0}).

Alternatively, we can relate the values of the coupling at two different scales by the 5-loop perturbative solution:

aQ2\displaystyle a_{Q_{2}} =\displaystyle= aQ1+β0ln(Q12Q22)aQ12+[β02ln2(Q12Q22)+β1ln(Q12Q22)]aQ13\displaystyle a_{Q_{1}}+\beta_{0}\ln\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)a_{Q_{1}}^{2}+\left[\beta_{0}^{2}\ln^{2}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\beta_{1}\ln\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)\right]a_{Q_{1}}^{3} (2.45)
+[β03ln3(Q12Q22)+52β0β1ln2(Q12Q22)+β2ln(Q12Q22)]aQ14\displaystyle+\left[\beta_{0}^{3}\ln^{3}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\frac{5}{2}\beta_{0}\beta_{1}\ln^{2}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\beta_{2}\ln\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)\right]a_{Q_{1}}^{4}
+[β04ln4(Q12Q22)+133β02β1ln3(Q12Q22)+32β12ln2(Q12Q22)+3β2β0ln2(Q12Q22)+β3ln(Q12Q22)]aQ15\displaystyle+\left[\beta_{0}^{4}\ln^{4}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\frac{13}{3}\beta_{0}^{2}\beta_{1}\ln^{3}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\frac{3}{2}\beta_{1}^{2}\ln^{2}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+3\beta_{2}\beta_{0}\ln^{2}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\beta_{3}\ln\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)\right]a_{Q_{1}}^{5}
+[β05ln5(Q12Q22)+7712β1β03ln4(Q12Q22)+(6β2β02+356β12β0)ln3(Q12Q22)\displaystyle+\left[\beta_{0}^{5}\ln^{5}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\frac{77}{12}\beta_{1}\beta_{0}^{3}\ln^{4}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\left(6\beta_{2}\beta_{0}^{2}+\frac{35}{6}\beta_{1}^{2}\beta_{0}\right)\ln^{3}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)\right.
+72(β3β0+β2β1)ln2(Q12Q22)+β4ln(Q12Q22)]aQ16+.\displaystyle\hskip 150.00023pt\left.+\frac{7}{2}\left(\beta_{3}\beta_{0}+\beta_{2}\beta_{1}\right)\ln^{2}\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)+\beta_{4}\ln\left(\frac{Q_{1}^{2}}{Q_{2}^{2}}\right)\right]a_{Q_{1}}^{6}+\cdots.

2.3   Renormalization scheme dependence

The βi\beta_{i} are the coefficients of the β\beta-function arising in the loop expansion, i.e. in powers of \hbar. Although the first two coefficients β0,β1\beta_{0},\beta_{1} are universally scheme-independent coefficients, depending only on the number of colors NcN_{c} and flavors NfN_{f}, the higher-order terms are, in contrast, scheme dependent. In particular, for the ’t Hooft scheme [135] the higher βi,i2\beta_{i},~{}i\geq 2 terms are set to zero, leading to the solution of Eq. (2.2.2) for the β\beta-function valid at all orders. Moreover, in all MS\rm MS-like schemes all the βi\beta_{i} coefficients are gauge independent, while other schemes, such as the momentum space subtraction (MOM) scheme [36, 6], depend on the particular gauge. Using the Landau gauge, the β\beta terms for the MOM scheme are given by [136]

β2=3040.48625.387Nf+19.3833Nf2\beta_{2}=3040.48-625.387N_{f}+19.3833N_{f}^{2}

and

β3=10054124423.3Nf+1625.4Nf227.493Nf3.\beta_{3}=100541-24423.3N_{f}+1625.4N_{f}^{2}-27.493N_{f}^{3}.

Results for the minimal MOM scheme and Landau gauge are given in Ref. [137]. The renormalization condition for the MOM scheme sets the virtual quark propagator to the same form as a free massless propagator. Different MOM schemes exist and the above values of β2\beta_{2} and β3\beta_{3} are determined with the MOM scheme defined by subtracting the 3-gluon vertex to a point with one null external momentum. This leads to a coupling that is not only RS dependent but also gauge dependent. The values of β2\beta_{2} and β3\beta_{3} given here are only valid in the Landau gauge. Values in the V\mathrm{V}-scheme defined by the static heavy-quark potential [138, 139, 140, 141, 142, 143, 144] can be found in Ref. [145]. They result in β2=4224.18746.01Nf+20.88Nf2\beta_{2}=4224.18-746.01N_{f}+20.88N_{f}^{2} and β3=4317512952Nf+707.0Nf2\beta_{3}=43175-12952N_{f}+707.0N_{f}^{2} respectively. We recall that the signs of the βi\beta_{i} control the running of αs\alpha_{s}. We have β0>0\beta_{0}>0 for Nf16,β1>0N_{f}\leq 16,\beta_{1}>0 for Nf8,β2>0N_{f}\leq 8,\beta_{2}>0 for Nf5N_{f}\leq 5 and β3\beta_{3} is always positive. Consequently, αs\alpha_{s} decreases at high momentum transfer, leading to the asymptotic freedom of pQCD. Note that, βi\beta_{i} are sometimes defined with an additional multiplying factor 1/(4π)i+11/(4\pi)^{i+1}. Different schemes are characterized by different βi,i2\beta_{i},i\geq 2 and lead to different definitions for the effective coupling.

The Λ\Lambda parameter represents the Landau ghost pole in the perturbative coupling in QCD. We recall that the Landau pole was initially identified in the context of Abelian QED. However, the presence of this pole does not affect QED. Given its value, Λ103040GeV\Lambda\sim 10^{30-40}\,{\rm GeV}, above the Planck scale [146], at which new physics is expected to occur in order to suppress the unphysical divergence. The QCD Λ\Lambda parameter in contrast is at low energies, its value depends on the RS, on the order of the β\beta-series, βi\beta_{i}, on the approximation of the coupling αs(μ)\alpha_{s}(\mu) at orders higher than β1\beta_{1} and on the number of flavors NfN_{f}. Although mass corrections due to light quarks at higher order in perturbative calculations introduce negligible terms, they actually indirectly affect αs\alpha_{s} through NfN_{f}. In fact, the number of active quark flavors runs with the scale QQ and a quark qq is considered active in loop integration if the scale QmqQ\geq m_{q}. Thus, in general, light quarks can be considered massless regardless of whether they are active or not, while αs\alpha_{s} varies smoothly when passing a quark threshold, rather than in discrete steps. The matching of the values of αs\alpha_{s} below and above a quark threshold makes Λ\Lambda depend on NfN_{f}. Matching requirements at leading order β0\beta_{0}, imply that:

αsNf1(Q=mq)=αsNf(Q=mq)\alpha_{s}^{N_{f}-1}\left(Q{=}m_{q}\right)=\alpha_{s}^{N_{f}}\left(Q{=}m_{q}\right)

and therefore that:

ΛNf=ΛNf1(ΛNf1mq)2/(332Nf)\Lambda^{N_{f}}=\Lambda^{N_{f}-1}\left(\frac{\Lambda^{N_{f}-1}}{m_{q}}\right)^{2/\left(33-2N_{f}\right)}

The formula with β1\beta_{1}, can be found in [147] and the four-loop matching in the MS¯\overline{\rm MS} RS is given in [148].

As shown in the previous section at the lowest order β0\beta_{0}, the Landau singularity is a simple pole on the positive real axis of the Q2Q^{2}-plane, whereas at higher order it acquires a more complicated structure. This pole is unphysical and is located on the positive real axis of the complex Q2Q^{2}-plane. This singularity of the coupling indicates that the perturbative regime of QCD breaks down and it may also suggest that a new mechanism takes over, such as the confinement. Thus, the value of Λ\Lambda is often associated with the confinement scale, or equivalently to the hadronic mass scale. An explicit relation between hadron masses and the Λ\Lambda scale has been obtained in the framework of holographic QCD [149]. Landau poles on the other hand, usually do not appear in nonperturbative approaches, such as AdS/QCD.

Different schemes are related perturbatively by:

αs(2)(Q)=αs(1)(Q)[1+v1αs(1)(Q)/(4π)]+𝒪(αs2)\alpha_{s}^{(2)}\left(Q\right)=\alpha_{s}^{(1)}\left(Q\right)\left[1+v_{1}\alpha_{s}^{(1)}\left(Q\right)/(4\pi)\right]+\mathcal{O}(\alpha_{s}^{2}) (2.46)

where v1v_{1} is the leading order difference between αs(Q)\alpha_{s}\left(Q\right) in the two schemes. In the case of the V-scheme and MS¯\overline{\rm MS} scheme we have: v1MS¯=319CA209TFnlv_{1}^{\overline{\mathrm{MS}}}=\frac{31}{9}C_{A}-\frac{20}{9}T_{F}n_{l}. Thus, the relation between Λ1\Lambda_{1} in a scheme 1 and Λ2\Lambda_{2} in a scheme 2 is, at the one-loop order, given by:

Λ2=Λ1ev12β0.\Lambda_{2}=\Lambda_{1}e^{\frac{v_{1}}{2\beta_{0}}}.

For example, the MS¯\overline{\rm MS} and V-scheme scale parameters are related by:

ΛV=ΛMS¯e9310Nf2(996Nf)\Lambda_{\rm V}=\Lambda_{\overline{\rm MS}}e^{\frac{93-10N_{f}}{2(99-6N_{f})}}

The relation is valid at each threshold, translating all values for the scale from one scheme to the other.

2.4   The Extended Renormalization Group Equations

Given that physical predictions cannot depend on the choice of the renormalization scale nor on the scheme, the same approach used for the renormalization scale based on the invariance under RGE is extended to scheme transformations. This approach leads to the Extended Renormalization Group Equations, which were introduced first by Stückelberg and Peterman [1], then discussed by Stevenson [8, 9, 11, 10] and also improved by Lu and Brodsky [150]. A physical quantity, RR, calculated at the NN-th order of accuracy is expressed as a truncated expansion in terms of a coupling constant αS(μ)\alpha_{S}(\mu) defined in the scheme S\rm S and at the scale μ\mu, such as

RN=r0αSp(μ)+r1(μ)αSp+1(μ)++rN(μ)αSp+N(μ).R_{N}=r_{0}\alpha_{S}^{p}(\mu)+r_{1}(\mu)\alpha_{S}^{p+1}(\mu)+\cdots+r_{N}(\mu)\alpha_{S}^{p+N}(\mu). (2.47)

At any finite order, the scale and scheme dependences of the coupling constant αS(μ)\alpha_{S}(\mu) and of the coefficient functions ri(μ)r_{i}(\mu) do not totally cancel, this leads to a residual dependence in the finite series and to the scale and scheme ambiguities.

In order to generalize the RGE approach, it is convenient to improve the notation by introducing the universal coupling function as the extension of an ordinary coupling constant to include the dependence on the scheme parameters {ci}\left\{c_{i}\right\}:

α=α(μ/Λ,{ci}).\alpha=\alpha\left(\mu/\Lambda,\left\{c_{i}\right\}\right). (2.48)

where Λ\Lambda is the standard two-loop MS¯\overline{\rm MS} scale parameter. The subtraction prescription is now characterized by an infinite set of continuous scheme parameters {ci}\left\{c_{i}\right\} and by the renormalization scale μ\mu. Stevenson [9] has shown that one can identify the beta-function coefficients of a given renormalization scheme with the scheme parameters. Considering that the first two coefficients of the β\beta-function are scheme independent, each scheme is identified by its {βi,i=2,3,}\left\{\beta_{i},~{}i=2,3,\ldots\right\} parameters.

More conveniently, let us define the rescaled coupling constant and the rescaled scale parameter as

a=β1β0α4π,τ=2β02β1log(μ/Λ).a=\frac{\beta_{1}}{\beta_{0}}\frac{\alpha}{4\pi},\quad\tau=\frac{2\beta_{0}^{2}}{\beta_{1}}\log(\mu/\Lambda). (2.49)

Then, the rescaled β\beta-function takes the canonical form:

β(a)=dadτ=a2(1+a+c2a2+c3a3+)\beta(a)=\frac{da}{d\tau}=-a^{2}\left(1+a+c_{2}a^{2}+c_{3}a^{3}+\cdots\right) (2.50)

with cn=βnβ0n1/β1nc_{n}=\beta_{n}\beta_{0}^{n-1}/\beta_{1}^{n} for n=2,3,n=2,3,\cdots.

The scheme and scale invariance of a given observable RR, can be expressed as:

δRδτ\displaystyle\frac{\delta R}{\delta\tau} =\displaystyle= βRa+Rτ=0\displaystyle\beta\frac{\partial R}{\partial a}+\frac{\partial R}{\partial\tau}=0
δRδcn\displaystyle\frac{\delta R}{\delta c_{n}} =\displaystyle= β(n)Ra+Rcn=0.\displaystyle\beta_{(n)}\frac{\partial R}{\partial a}+\frac{\partial R}{\partial c_{n}}=0. (2.51)

The fundamental beta function that appears in Eqs. (2.51) reads:

β(a,{ci})δaδτ=a2(1+a+c2a2+c3a3+)\beta\left(a,\left\{c_{i}\right\}\right)\equiv\frac{\delta a}{\delta\tau}=-a^{2}\left(1+a+c_{2}a^{2}+c_{3}a^{3}+\cdots\right) (2.52)

and the extended or scheme-parameter beta functions are defined as:

β(n)(a,{ci})δaδcn.\beta_{(n)}\left(a,\left\{c_{i}\right\}\right)\equiv\frac{\delta a}{\delta c_{n}}. (2.53)

The extended beta functions can be expressed in terms of the fundamental beta function. Since the (τ,{ci})(\tau,\{c_{i}\}) are independent variables, second partial derivatives respect the commutativity relation:

δ2aδτδcn=δ2aδcnδτ,\frac{\delta^{2}a}{\delta\tau\delta c_{n}}=\frac{\delta^{2}a}{\delta c_{n}\delta\tau}, (2.54)

which implies

δβ(n)δτ=δβδcn,\frac{\delta\beta_{(n)}}{\delta\tau}=\frac{\delta\beta}{\delta c_{n}}, (2.55)
ββ(n)=β(n)βan+2,\beta\beta_{(n)}^{\prime}=\beta_{(n)}\beta^{\prime}-a^{n+2}, (2.56)

where β(n)=β(n)/a\beta_{(n)}^{\prime}=\partial\beta_{(n)}/\partial a and β=β/a\beta^{\prime}=\partial\beta/\partial a. From here

β2(β(n)β)=an+2,\beta^{-2}\left(\frac{\beta_{(n)}}{\beta}\right)^{\prime}=-a^{n+2}, (2.57)
β(n)(a,{ci})=β(a,{ci})0a𝑑xxn+2β2(x,{ci}),\beta_{(n)}\left(a,\left\{c_{i}\right\}\right)=-\beta\left(a,\left\{c_{i}\right\}\right)\int_{0}^{a}dx\,\frac{x^{n+2}}{\beta^{2}\left(x,\left\{c_{i}\right\}\right)}, (2.58)

where the lower limit of the integral has been set to satisfy the boundary condition

β(n)O(an+1).\beta_{(n)}\sim O\left(a^{n+1}\right).

That is, a change in the scheme parameter cnc_{n} can only affect terms of order an+1a^{n+1} or higher in the evolution of the universal coupling function.

The extended renormalization group equations Eqs. (2.51) can be written in the form:

Rτ\displaystyle\frac{\partial R}{\partial\tau} =\displaystyle= βRa\displaystyle-\beta\frac{\partial R}{\partial a}
Rcn\displaystyle\frac{\partial R}{\partial c_{n}} =\displaystyle= β(n)Ra.\displaystyle-\beta_{(n)}\frac{\partial R}{\partial a}. (2.59)

Thus, provided we know the extended beta functions, we can determine any variation of the expansion coefficients of RR under scale-scheme transformations. In particular, we can evolve a given perturbative series into another, determining the expansion coefficients of the latter and vice versa. Thus, different schemes and scales can be related according to the extended renormalization group equations and the fundamental requirement of “renormalization scale and scheme invariance” is recovered via the extended renormalization group invariance of perturbative QCD. Unfortunately, these relations and in general all perturbative calculations are known only up to a certain level of accuracy and the truncated formulas are responsible for an important source of uncertainties: the scheme and scale ambiguities.

2.5   The running coupling constant αs(μ)\alpha_{s}(\mu)

The strong coupling αs\alpha_{s}, is a fundamental parameter of the SM theory and determines the strength of the interactions among quarks and gluons in quantum chromodynamics (QCD).

In order to understand hadronic interactions, it is necessary to determine the magnitude of the coupling and its behavior over a wide range of values, from low- to high-energy scales. Long and short distances are related to low and high energies respectively. In the high-energy region the strong coupling has an asymptotic behavior and QCD becomes perturbative, while in the region of low energies, e.g. at the proton-mass scale, the dynamics of QCD is affected by processes such as quark confinement, soft radiation and hadronization. In the first case experimental results can be matched with theoretical calculations and a precise determination of αs\alpha_{s} depends on both experimental accuracy and theoretical errors. In the latter case experimental results are difficult to achieve and theoretical predictions are affected by the confinement and hadronization mechanisms, which are rather model dependent. Various processes also involve a precise knowledge of the coupling in both the high and low momentum transfer regions and in some cases calculations must be improved with electroweak (EW) corrections. Thus, the determination of the QCD coupling over a wide range of energy scales is a crucial task in order to achieve results and to test QCD to the highest precision. Theoretical uncertainties in the value of αs(Q)\alpha_{s}(Q) contribute to the total theoretical uncertainty in the physics investigated at the Large Hadron Collider (LHC), such as the Higgs sector, e.g. Higgs production via gluon fusion [19]. The behavior of the perturbative coupling at low-momentum transfer is also fundamental for the scale of the proton mass, in order to understand hadronic structure, quark confinement and hadronization processes. IR effects, such as soft radiation and renormalon factorial growth, spoil the perturbative nature of QCD in the low-energy domain. Higher-twist effects can also play an important role. Processes involving the production of heavy quarks near threshold require knowledge of the QCD coupling at very low momentum scales. Even reactions at high energies may involve the integration of the behavior of the strong coupling over a large domain of momentum scales, including IR regions. Precision tests of the coupling are crucial also for other aspects of QCD that are still under continuous investigation, such as the hadron masses and their internal structure. In fact, the strong interaction is responsible for the mass of hadrons in the zero-mass limit of the uu, dd quarks.

The origin and phenomenology of the behavior of αs(μ)\alpha_{s}(\mu) at small distances, where asymptotic freedom appears, is well understood and explained in many textbooks on Quantum Field Theory and Particle Physics. Numerous reviews exist; see e.g. Refs. [151, 152]. However, standard explanations often create an apparent puzzle. Other questions remain even in this well understood regime: a significant issue is how to identify the scale QQ that controls a given hadronic process, especially if it depends on many physical scales.

3   The Principle of Maximum Conformality – PMC scale setting

The Principle of Maximum Conformality (PMC) [45, 46, 47, 49, 50] is the principle underlying BLM and it generalizes the BLM method to all possible applications and to all orders.

BLM scale setting is greatly inspired by QED. The standard Gell-Mann–Low scheme determines the correct renormalization scale identifying the scale with the virtuality of the exchanged photon [2]. For example, in electron–muon elastic scattering, the renormalization scale is given by the virtuality of the photon exchanged, i.e. the spacelike momentum transfer squared μr2=q2=t\mu_{r}^{2}=q^{2}=t. Thus,

α(t)=α(t0)1Π(t,t0),\alpha(t)={\alpha(t_{0})\over 1-\Pi(t,t_{0})}, (3.1)

where

Π(t,t0)=Π(t)Π(t0)1Π(t0)\Pi(t,t_{0})={\Pi(t)-\Pi(t_{0})\over 1-\Pi(t_{0})}

is the vacuum polarization (VP) function. From Eq. (3.1) it follows that the renormalization scale μR2=t\mu^{2}_{R}=t can be determined by the β0\beta_{0}-term at the lowest order. This scale is sufficient to sum all the vacuum polarization contributions into the dressed photon propagator, both proper and improper to all orders. Starting from a first evaluation of the physical observable that is obtained by calculating perturbative corrections in a given scheme (commonly used are MS\rm MS or MS¯\overline{\rm MS}) and at an initial renormalization scale μr=μrinit\mu_{r}=\mu_{r}^{\rm init}, one obtains the truncated expansion:

ρn=𝒞0αsp(μr)+i=1n𝒞i(μr)αsp+i(μr),(p0),\rho_{n}=\mathcal{C}_{0}\alpha_{s}^{p}\left(\mu_{r}\right)+\sum_{i=1}^{n}\mathcal{C}_{i}\left(\mu_{r}\right)\alpha_{s}^{p+i}\left(\mu_{r}\right),\quad(p\geq 0), (3.2)

where 𝒞0\mathcal{C}_{0} is the tree-level term, while 𝒞1,𝒞2,,𝒞n\mathcal{C}_{1},\mathcal{C}_{2},\dots,\mathcal{C}_{n} are the one-loop, two-loop and nn-loop corrections respectively and pp is the power of the coupling at tree-level. In order to improve the pQCD estimate of the observable, after the initial renormalization a change of scale using the RGE is performed according to the BLM scale setting.

Following the GM–L scheme in QED, the BLM scales can be determined at LO in perturbation theory by writing explicit contributions coming from the different NfN_{f} terms of the NLO coefficient in a physical observable as [15]:

ρ\displaystyle\rho =\displaystyle= C0αs,MS¯p(μr)[1+(ANf+B)αs,MS¯(μr)π]\displaystyle C_{0}\alpha_{s,\overline{\rm MS}}^{p}\left(\mu_{r}\right)\left[1+\left(AN_{f}+B\right)\frac{\alpha_{s,\overline{\rm MS}}\left(\mu_{r}\right)}{\pi}\right] (3.3)
=\displaystyle= C0αs,MS¯p(μr)[1+(32Aβ0+332A+B)αs,MS¯(μr)π],\displaystyle C_{0}\alpha_{s,\overline{\rm MS}}^{p}\left(\mu_{r}\right)\left[1+\left(-\frac{3}{2}A\beta_{0}+\frac{33}{2}A+B\right)\frac{\alpha_{s,\overline{\rm MS}}\left(\mu_{r}\right)}{\pi}\right],

where μr=μrinit\mu_{r}=\mu_{r}^{\rm init} stands for an initial renormalization scale, which practically can be taken as the typical momentum transfer of the process. The NfN_{f} term is due to the quark vacuum polarization. Calculations are in the MS¯\overline{\rm MS}-scheme.

At the NLO level, all NfN_{f} terms should be resummed into the coupling. Using the NLO αs\alpha_{s}-running formula:

αs,MS¯(μr)=αs,MS¯(μr)1+β04παs,MS¯(μr)ln(μrμr),\alpha_{s,\overline{\rm MS}}\left(\mu_{r}^{*}\right)=\frac{\alpha_{s,\overline{\rm MS}}\left(\mu_{r}\right)}{1+\frac{\beta_{0}}{4\pi}\alpha_{s,\overline{\rm MS}}\left(\mu_{r}\right)\ln\left(\frac{\mu_{r}^{*}}{\mu_{r}}\right)}, (3.4)

we obtain

ρ=C0αs,MS¯p(μr)[1+C1αs,MS¯(μr)π],\rho=C_{0}\alpha_{s,\overline{\rm MS}}^{p}\left(\mu_{r}^{*}\right)\left[1+C_{1}^{*}\frac{\alpha_{s,\overline{\rm MS}}\left(\mu_{r}^{*}\right)}{\pi}\right], (3.5)

where

μr=μrexp(3Ap)\mu_{r}^{*}=\mu_{r}\exp\left(\frac{3A}{p}\right)

is the BLM scale and

C1=332A+BC_{1}^{*}=\frac{33}{2}A+B

is the conformal coefficient, i.e. the NLO coefficient not depending on the RS and scale μ\mu. Both the effective BLM scale μr\mu_{r}^{*} and the coefficient C1C_{1}^{*} are NfN_{f} independent and conformal at LO. By including the term 33A/233A/2 into the scale we eliminate the β0\beta_{0} term of the NLO coefficient C1C_{1}, which is responsible for the running of the coupling constant, and the observable in the final results can be written in its maximal conformal form, Eq. (3.5).

The BLM method can be extended to higher orders in a systematic way by including the nfn_{f} terms arising at higher order into the BLM scales consistently. In order to extend the BLM beyond the NLO, the following points are considered essential:

  1. 1.

    All nfn_{f}-terms associated with the β\beta-function (i.e. NfN_{f} terms) and then with the renormalization of the coupling constant, must be absorbed into the effective coupling, while those nfn_{f}-terms that have no relation with UV-divergent diagrams (i.e. NFN_{F}-terms) should be identified and considered as part of the conformal coefficients. After BLM scale setting, the perturbative series for the physical observable becomes a conformal series, all nonconformal terms should be absorbed into the effective coupling in a consistent manner.

  2. 2.

    New NfN_{f}-terms (corresponding to new β0\beta_{0} coefficients) arise at each perturbative order, thus a new BLM scale that sums these terms consistently into the running coupling, should be introduced at each calculated perturbative order. In fact, there is no reason to use a unified effective scale for the entire perturbative series as shown in Refs. [153, 154].

  3. 3.

    The BLM scales themselves should be a RG-improved perturbative series [46]. The length of the perturbative series for each BLM scale depends on how many new NfN_{f}-terms (or βi\beta_{i}-terms) we have from the higher-order calculation and to what perturbative order we have performed.

Actually, the last point is not mandatory and needs clarification. In order to apply the BLM/PMC using perturbative scales, the argument of the coupling in the expansion of the BLM/PMC scale should be the physical scale of the process QQ, that may be either the center-of-mass energy s\sqrt{s} or even another variable, such as t,u,MH,\sqrt{t},\sqrt{u},M_{H},\dots, depending on the process. Setting the initial scale to the physical scale would greatly simplify the BLM/PMC procedure, preserving the original scale invariance of the observable and eliminating the initial scale dependence from the BLM/PMC scales. In case the BLM/PMC scales are not perturbatively calculated, as will be shown in Section 4.1, the initial scale can be treated as an arbitrary parameter.

In agreement with these indications, it is possible to achieve a scale setting method extendible iteratively to all orders, which leads to the correct coefficients 𝒞i(μBLM){\mathcal{C}_{i}}(\mu^{*}_{BLM}) for the final “maximally conformal” series:

ρn=𝒞0αsp(μBLM)+𝒞1(μBLM)αsp+1(μBLM)+𝒞2(μBLM)αsp+2(μBLM)+\rho_{n}=\mathcal{C}_{0}\alpha_{s}^{p}\left(\mu_{\rm BLM}^{*}\right)+{\mathcal{C}}_{1}\left(\mu_{\rm BLM}^{**}\right)\alpha_{s}^{p+1}\left(\mu_{\rm BLM}^{**}\right)+{\mathcal{C}}_{2}\left(\mu_{\rm BLM}^{***}\right)\alpha_{s}^{p+2}\left(\mu_{\rm BLM}^{***}\right)+\cdots (3.6)

where the BLM scales μBLM,μBLM,\mu^{*}_{\rm BLM},\mu^{**}_{\rm BLM},\dots are set by a recursive use of the RG equations in order to cancel all the NfN_{f} terms from the series. We remark that since the coefficients 𝒞i(μBLM){\mathcal{C}_{i}}(\mu^{*}_{BLM}) have been obtained cancelling all β\beta terms related to running of the coupling they actually are free of any scale and scheme dependence left. In other words, the 𝒞i(μBLM)𝒞i~{\mathcal{C}_{i}}(\mu^{*}_{BLM})\equiv\tilde{\mathcal{C}_{i}}, where the 𝒞i~\tilde{\mathcal{C}_{i}} are conformal coefficients not depending on the renormalization scale. Hence, the BLM approach leads to an maximally conformal observable, i.e. where all the renormalization scale and scheme dependence has been confined to the effective coupling and to its renormalization scale αs(μBLM)\alpha_{s}(\mu_{\rm BLM}).

Fundamental features of the BLM method:

  1. A)

    BLM scales at LO, are set simply by identifying the coefficient AA of the NfN_{f} term.

  2. B)

    Since all nfn_{f}-terms related to the running of the coupling are reabsorbed, scheme differences do not affect the results and the perturbative expansions in αs(μr)\alpha_{s}(\mu_{r}^{*}) in two different schemes, e.g. MS\rm MS and MS¯\overline{\rm MS}, are identical. We notice that nfn_{f}-terms related to the UV finite diagrams, may arise at every order in perturbation theory. These terms might be related either to the particular kinematics of the initial state or even to finite loop diagrams arising at higher orders, thus in both cases are insensitive to the UV cutoff or to the RS and cannot be considered as β\beta-terms. We label these terms as NFN_{F}-terms and they do not give contributions to the BLM scales.

  3. C)

    Using BLM scale setting, the perturbative expansion does not change across quark threshold, given that all vacuum-polarization effects due to a new quark are automatically absorbed into the effective coupling. This implies that in a process with fixed kinematic variables (e.g. a total cross-section), we can use a naive LO/NLO αs(μrBLM)\alpha_{s}(\mu_{r}^{\rm BLM})-running, with the number of active flavor NfN_{f} fixed to the value determined by the BLM scale, to perform the calculation [155].

  4. D)

    The BLM method preserves all the RG properties of existence and uniqueness, reflexivity, symmetry and transitivity. As shown in Refs. [156, 157, 158, 159], the RG invariance of the BLM leads to scheme-independent transformations that relate couplings in different schemes. These are known as commensurate scale relations (CSRs) and it has been shown that, even though the expansion coefficients under different renormalization schemes can be different, after a proper scale setting, one can determine a relation between the effective couplings leading to an invariant result for the calculated quantity. Using this approach it is also possible to extend conformal properties to renormalizable gauge theories, such as the generalized Crewther relation [160, 161, 162, 163].

  5. E)

    The BLM approach reduces to the GM–L scheme for QED in the Abelian limit Nc0N_{c}\rightarrow 0 [32]; the results are in perfect agreement.

  6. F)

    The elimination of the NfN_{f} term related to the β0\beta_{0} coefficient from the perturbative series eliminates the renormalon terms n!β0nαsn+1n!\beta^{n}_{0}\alpha^{n+1}_{s} over the entire range of the accessible physical energies and not only in the low-energy domain. The convergence of the resulting series is then greatly improved.

Several extended versions of the BLM approach beyond the NLO have been proposed in the literature, such as the dressed-skeleton expansion, the large-β0\beta_{0} expansion, the BLM expansion with an overall renormalization scale and the sequential BLM (seBLM), an extension to the sequential BLM (xBLM) in Refs. [164, 165, 159, 163, 153, 166, 167, 168]. These different extensions of the BLM are mostly partial or ad hoc improvements of the first LO-BLM [15] in some cases up to NNLO, in other cases using a rather effective approach, i.e. by introducing an overall effective BLM scale for the entire perturbative expansion. Results obtained with these approaches also did not respect the basic points (1–3). Most importantly, these methods lead to results that are still dependent on the initial renormalization scale. The fundamental feature of the BLM is to obtain results free of scale ambiguities and thus independent of the choice of initial renormalization scale. The first aim of the BLM scale is to eliminate the renormalization scale and scheme uncertainties; thus any extension of the BLM not respecting this basic requirement does not represent a real improvement of the standard conventional scale setting CSS method.

The reasons for the different extensions of the BLM method to higher orders were mainly two: firstly, it was not clear how to generalize this approach to all possible quantities, which translates into the question: what is the principle underlying the BLM method? And secondly, what is the correct procedure to identify and reabsorb the NfN_{f}-terms unambiguously order-by-order? A practical reason that renders the extension to higher orders not straightforward is the presence of finite UV corrections given by the three- and four-gluon vertices of the additional NFN_{F}-terms that are unrelated to the running of αs\alpha_{s}.

In its first formulation in Ref. [45] it was suggested to use a unique PMC scale at LO to reabsorb all β\beta contributions related to different skeleton-graph scales by properly weighting the two contributions, such as that of the tt-channel and ss-channel. This approach was more oriented towards a single PMC scale that reabsorbs all β\beta terms related to the running coupling. A multi-scale approach was later developed considering different scales arising at each order of accuracy including different β\beta coefficients according to the perturbative expansion. We remark that the PMC method preserves all the properties (A–F) of the BLM procedure and it extends these properties to all orders, eliminating the renormalization scale and scheme ambiguities. The PMC also generalizes this approach to all gauge theories. Firstly, this is crucial in order to apply the same method to all SM sectors. Secondly, in the perspective of a grand unified theory (GUT), only one scale-setting method may be applied for consistency, and this method must agree with the GM–L scheme and with the QED results.

In order to apply the PMC, is convenient to follow the flowchart shown in Fig. 3.1 and to write the observable of Eq. (3.2) with the explicit contributions of the nfn_{f} terms in the coefficients calculated at each order of accuracy:

Refer to caption
Figure 3.1: Flowchart for the PMC procedure.
ρ(Q)=i=1n(j=0i1ci,j(μr,Q)nfj)asp+i1(μr),\rho(Q)=\sum^{n}_{i=1}\left(\sum^{i-1}_{j=0}c_{i,j}(\mu_{r},Q)n_{f}^{j}\right)a_{s}^{p+i-1}(\mu_{r}), (3.7)

where QQ represents the kinematic scale or the physical scale of the measured observable and pp is the power of αs\alpha_{s} associated with the tree-level terms. In general, this procedure is always possible either for analytic or numerical (e.g. Monte Carlo) calculations, given that both strategies keep track of terms related to different color factors. The core of the PMC method, as was for BLM, is that all NfN_{f} terms related to the β\beta-function arising in a perturbative calculation must be summed, by proper definition of the renormalization scale μPMC\mu_{\rm PMC}, into the effective coupling αs\alpha_{s} by recursive use of the RGE. Essentially the difference between the two procedures is that while in BLM the scales are set iteratively order-by-order to remove all nfn_{f} terms, in the PMC the nfn_{f} terms are written first as β\beta terms and then reabsorbed into the effective coupling. The two procedures are related by the correspondence principle [46].

3.1   The multi-scale Principle of Maximum Conformality: PMCm

The PMCm method is based on a multi-scale application of the PMC. In this section we show how to implement this method to any order of accuracy. First, as shown in the flowchart in Fig. 3.1, for the pQCD approximant (3.7), it is convenient to transform the {nf}\{n_{f}\} series at each order into the {βi}\{\beta_{i}\} series. The QCD degeneracy relations [169] ensure the realizability of such a transformation. For example, Eq. (3.7) may be rewritten as [49, 50]

ρ(Q)\displaystyle\rho(Q) =\displaystyle= r1,0as(μr)+(r2,0+β0r2,1)as2(μr)\displaystyle r_{1,0}\,a_{s}(\mu_{r})+\bigg{(}r_{2,0}+\beta_{0}r_{2,1}\bigg{)}a_{s}^{2}(\mu_{r}) (3.8)
+(r3,0+β1r2,1+2β0r3,1+β02r3,2)as3(μr)\displaystyle+\bigg{(}r_{3,0}+\beta_{1}r_{2,1}+2\beta_{0}r_{3,1}+\beta_{0}^{2}r_{3,2}\bigg{)}a_{s}^{3}(\mu_{r})
+(r4,0+β2r2,1+2β1r3,1+52β1β0r3,2+3β0r4,1+3β02r4,2+β03r4,3)as4(μr)+,\displaystyle+\bigg{(}r_{4,0}+\beta_{2}r_{2,1}+2\beta_{1}r_{3,1}+\frac{5}{2}\beta_{1}\beta_{0}r_{3,2}+3\beta_{0}r_{4,1}+3\beta_{0}^{2}r_{4,2}+\beta_{0}^{3}r_{4,3}\bigg{)}a_{s}^{4}(\mu_{r})+\cdots,~{}

where ri,jr_{i,j} can be derived from ci,jc_{i,j}, ri,0r_{i,0} are conformal coefficients and ri,jr_{i,j} (j0)(j\neq 0) are nonconformal. For definiteness and without loss of generality, we have set p=1p=1 and n=4n=4 to illustrate the PMC procedures. Different types of {βi}\{\beta_{i}\}-terms can be absorbed into αs\alpha_{s} in an order-by-order manner by using the RGE, which leads to distinct PMC scales at each order:

ask(Qk)\displaystyle a^{k}_{s}(Q_{k}) \displaystyle\leftarrow ask(μr){1+kβ0rk+1,1rk,0as(μr)\displaystyle a^{k}_{s}(\mu_{r})\bigg{\{}1+k\beta_{0}\frac{r_{k+1,1}}{r_{k,0}}a_{s}(\mu_{r}) (3.9)
+k(β1rk+1,1rk,0+k+12β02rk+2,2rk,0)as2(μr)\displaystyle\hskip 40.00006pt+k\left(\beta_{1}\frac{r_{k+1,1}}{r_{k,0}}+\frac{k+1}{2}\beta_{0}^{2}\frac{r_{k+2,2}}{r_{k,0}}\right)a^{2}_{s}(\mu_{r})
+k(β2rk+1,1rk,0+2k+32β0β1rk+2,2rk,0+(k+1)(k+2)3!β03rk+3,3rk,0)as3(μr)+}.\displaystyle\hskip 40.00006pt+k\bigg{(}\beta_{2}\frac{r_{k+1,1}}{r_{k,0}}+\frac{2k+3}{2}\beta_{0}\beta_{1}\frac{r_{k+2,2}}{r_{k,0}}+\frac{(k+1)(k+2)}{3!}\beta_{0}^{3}\frac{r_{k+3,3}}{r_{k,0}}\bigg{)}a^{3}_{s}(\mu_{r})+\cdots\bigg{\}}.

The coefficients ri,jr_{i,j} are generally functions of μr\mu_{r}, which can be redefined as

ri,j=k=0jCjkr^ik,jklnk(μr2/Q2),\displaystyle r_{i,j}=\sum^{j}_{k=0}C^{k}_{j}{\hat{r}}_{i-k,j-k}{\rm ln}^{k}(\mu_{r}^{2}/Q^{2}), (3.10)

where the reduced coefficients r^i,j=ri,j|μr=Q{\hat{r}}_{i,j}=r_{i,j}|_{\mu_{r}=Q} (specifically, we have r^i,0=ri,0{\hat{r}}_{i,0}=r_{i,0}) and the combinatorial coefficients Cjk=j!/(k!(jk)!)C^{k}_{j}=j!/(k!(j-k)!). As discussed in the previous section, we set the renormalization scale μr\mu_{r} to the physical scale of the process QQ:

lnQ2Q12\displaystyle\ln\frac{Q^{2}}{Q_{1}^{2}} =\displaystyle= r^2,1r^1,0+β0(r^1,0r^3,2r^2,12r^1,02)as(Q)\displaystyle\frac{{\hat{r}}_{2,1}}{{\hat{r}}_{1,0}}+\beta_{0}\left(\frac{{\hat{r}}_{1,0}{\hat{r}}_{3,2}-{\hat{r}}_{2,1}^{2}}{{\hat{r}}_{1,0}^{2}}\right)a_{s}(Q) (3.11)
+[β1(3r^3,22r1,03r^2,122r^1,02)+β02(r^4,3r^1,02r^3,2r^2,1r^1,02+r^2,13r^1,03)]as2(Q)+,\displaystyle+\left[\beta_{1}\left(\frac{3{\hat{r}}_{3,2}}{2r_{1,0}}-\frac{3{\hat{r}}_{2,1}^{2}}{2{\hat{r}}_{1,0}^{2}}\right)+\beta_{0}^{2}\left(\frac{{\hat{r}}_{4,3}}{{\hat{r}}_{1,0}}-\frac{2{\hat{r}}_{3,2}{\hat{r}}_{2,1}}{{\hat{r}}_{1,0}^{2}}+\frac{{\hat{r}}_{2,1}^{3}}{{\hat{r}}_{1,0}^{3}}\right)\right]a^{2}_{s}(Q)+\cdots,
lnQ2Q22\displaystyle\ln\frac{Q^{2}}{Q_{2}^{2}} =\displaystyle= r^3,1r^2,0+3β0r^2,0r^4,2r^3,122r^2,02as(Q)+,\displaystyle\frac{{\hat{r}}_{3,1}}{{\hat{r}}_{2,0}}+3\beta_{0}\frac{{\hat{r}}_{2,0}{\hat{r}}_{4,2}-{\hat{r}}_{3,1}^{2}}{2{\hat{r}}_{2,0}^{2}}a_{s}(Q)+\cdots, (3.12)
lnQ2Q32\displaystyle\ln\frac{Q^{2}}{Q_{3}^{2}} =\displaystyle= r^4,1r^3,0+.\displaystyle\frac{{\hat{r}}_{4,1}}{{\hat{r}}_{3,0}}+\cdots. (3.13)

Note that the PMC scales are of a perturbative nature, which is also a sort of resummation and we need to know more loop terms to achieve more accurate predictions. The PMC resums all the known same type of {βi}\{\beta_{i}\}-terms to form precise PMC scales for each order. Thus, the precision of the PMC scale for the high-order terms decreases at higher and higher orders due to the less known {βi}\{\beta_{i}\}-terms in those higher-order terms. For example, Q1Q_{1} is determined up to next-to-next-to-leading-logarithm (N2LL) accuracy, Q2Q_{2} is determined up to NLL accuracy and Q3Q_{3} is determined at LL accuracy. Thus, the PMC scales at higher orders are of less accuracy due to more of the perturbative terms being unknown. This perturbative property of the PMC scale causes the first kind of residual scale dependence.

After fixing the magnitude of as(Qk)a_{s}(Q_{k}), we achieve a conformal series

ρ(Q)\displaystyle\rho(Q) =\displaystyle= i=14r^i,0asi(Qi)+.\displaystyle\sum_{i=1}^{4}{\hat{r}}_{i,0}a^{i}_{s}(Q_{i})+\cdots. (3.14)

The PMC scale for the highest-order term, e.g. Q4Q_{4} for the present case, is unfixed, since there is no {βi}\{\beta_{i}\}-term to determine its magnitude. This renders the last perturbative term unfixed and causes the second kind of residual scale dependence. Usually, the PMCm suggests setting Q4Q_{4} as the last determined scale Q3Q_{3}, which ensures the scheme independence of the prediction due to commensurate scale relations among the predictions under different renormalization schemes [159, 170]. The pQCD series (3.14) is renormalization scheme and scale independent and becomes more convergent due to the elimination of the β\beta terms including those related to the renormalon divergence. Thus, a more accurate pQCD prediction can be achieved by applying the PMCm. Two residual scale dependences are due to perturbative nature of either the pQCD approximant or the PMC scale, which is in principal different from the conventional arbitrary scale dependence. In practice, we have found that these two residual scale dependences are quite small even at low orders. This is due to a generally faster pQCD convergence after applying the PMCm. Some examples can be found in Ref. [171].

3.2   The single-scale Principle of Maximum Conformality: PMCs

In some cases, the perturbative series might have a weak convergence and the PMC scales might retain a comparatively larger residual scale dependence. To overcome this, a single-scale approach has been proposed, namely the PMCs, in order to suppress the residual scale dependence by directly fixing a single effective αs\alpha_{s}. Following the standard procedures of PMCs [172], the pQCD approximant (3.8) changes to the following conformal series,

ρ(Q)\displaystyle\rho(Q) =\displaystyle= i=14r^i,0asi(Q)+.\displaystyle\sum_{i=1}^{4}{\hat{r}}_{i,0}a^{i}_{s}(Q_{*})+\cdots. (3.15)

As in the previous section, we have set p=1p=1 and n=4n=4 for illustrating the procedure. The PMC scale QQ_{*} can be determined by requiring all the nonconformal terms to vanish, which can be fixed up to N2LL accuracy for p=1p=1 and n=4n=4, i.e. lnQ2/Q2\ln Q^{2}_{*}/Q^{2} can be expanded as a power series over as(Q)a_{s}(Q),

lnQ2Q2=T0+T1as(Q)+T2as2(Q)+,\displaystyle\ln\frac{Q^{2}_{*}}{Q^{2}}=T_{0}+T_{1}a_{s}(Q)+T_{2}a^{2}_{s}(Q)+\cdots, (3.16)

where the coefficients Ti(i=0,1,2)T_{i}~{}(i=0,1,2) are

T0\displaystyle T_{0} =r^2,1r^1,0,\displaystyle=-\frac{{\hat{r}}_{2,1}}{{\hat{r}}_{1,0}}, (3.17)
T1\displaystyle T_{1} =β0(r^2,12r^1,0r^3,2)r^1,02+2(r^2,0r^2,1r^1,0r^3,1)r^1,02,\displaystyle=\frac{\beta_{0}({\hat{r}}_{2,1}^{2}-{\hat{r}}_{1,0}{\hat{r}}_{3,2})}{{\hat{r}}_{1,0}^{2}}+\frac{2({\hat{r}}_{2,0}{\hat{r}}_{2,1}-{\hat{r}}_{1,0}{\hat{r}}_{3,1})}{{\hat{r}}_{1,0}^{2}},
and
T2\displaystyle T_{2} =3β1(r^2,12r^1,0r^3,2)2r^1,02\displaystyle=\frac{3\beta_{1}({\hat{r}}_{2,1}^{2}-{\hat{r}}_{1,0}{\hat{r}}_{3,2})}{2{\hat{r}}_{1,0}^{2}}
+4(r^1,0r^2,0r^3,1r^2,02r^2,1)+3(r^1,0r^2,1r^3,0r^1,02r^4,1)r^1,03\displaystyle\quad+\frac{4({\hat{r}}_{1,0}{\hat{r}}_{2,0}{\hat{r}}_{3,1}-{\hat{r}}_{2,0}^{2}{\hat{r}}_{2,1})+3({\hat{r}}_{1,0}{\hat{r}}_{2,1}{\hat{r}}_{3,0}-{\hat{r}}_{1,0}^{2}{\hat{r}}_{4,1})}{{\hat{r}}_{1,0}^{3}}
+β0(4r^2,1r^3,1r^1,03r^4,2r^1,02+2r^2,0r^3,2r^1,03r^2,0r^2,12)r^1,03\displaystyle\quad+\frac{\beta_{0}(4{\hat{r}}_{2,1}{\hat{r}}_{3,1}{\hat{r}}_{1,0}-3{\hat{r}}_{4,2}{\hat{r}}_{1,0}^{2}+2{\hat{r}}_{2,0}{\hat{r}}_{3,2}{\hat{r}}_{1,0}-3{\hat{r}}_{2,0}{\hat{r}}_{2,1}^{2})}{{\hat{r}}_{1,0}^{3}}
+β02(2r^1,0r^3,2r^2,1r^2,13r^1,02r^4,3)r^1,03.\displaystyle\quad+\frac{\beta_{0}^{2}(2{\hat{r}}_{1,0}{\hat{r}}_{3,2}{\hat{r}}_{2,1}-{\hat{r}}_{2,1}^{3}-{\hat{r}}_{1,0}^{2}{\hat{r}}_{4,3})}{{\hat{r}}_{1,0}^{3}}. (3.18)

Eq. (3.16) shows that the PMC scale QQ_{*} is also a power series over αs\alpha_{s}, which resums all the known {βi}\{\beta_{i}\}-terms and is explicitly independent of μr\mu_{r} at any fixed order, but depends only on the physical scale Q. It represents the correct momentum flow of the process and determines an overall effective αs\alpha_{s} value. Together with the μr\mu_{r}-independent conformal coefficients, the resultant PMC pQCD series is scheme and scale independent [58]. By using a single PMC scale determined with the highest accuracy from the known pQCD series, both the first and the second kind of residual scale dependence are suppressed.

4   Infinite-Order Scale Setting via the Principle of Maximum Conformality: PMC

In this section we introduce a parametrization of the observables that stems directly from the analysis of the perturbative QCD corrections and which reveals interesting properties, such as scale invariance, independently of the process or of the kinematics. We point out that this parametrization can be an intrinsic general property of gauge theories and we define this property intrinsic conformality (iCF333Here the conformality must be understood as RG invariance only.). We also show how this property directly indicates the correct renormalization scale μr\mu_{r} at each order of calculation and we define this new method PMC: Infinite-Order Scale Setting using the Principle of Maximum Conformality. We apply the iCF property and the PMC to the case of the thrust and CC-parameter distributions in e+e3e^{+}e^{-}\rightarrow 3\,jets and we display the results.

4.1   Intrinsic conformality (iCF)

In order to introduce intrinsic conformality (iCF), we consider the case of a normalized IR-safe single-variable distribution and write the explicit sum of pQCD contributions calculated up to NNLO at the initial renormalization scale μ0\mu_{0}:

1σ0Odσ(μ0)dO\displaystyle\frac{1}{\sigma_{0}}\!\frac{Od\sigma(\mu_{0})}{dO}\! =\displaystyle\!\!=\!\! {αs(μ0)2πOdAO(μ0)dO+(αs(μ0)2π)2OdBO(μ0)dO+(αs(μ0)2π)3OdCO(μ0)dO+𝒪(αs4)},\displaystyle\!\left\{\!\frac{\alpha_{s}(\mu_{0})}{2\pi}\frac{OdA_{\mathit{O}}(\mu_{0})}{dO}+\!\left(\!\frac{\alpha_{s}(\mu_{0})}{2\pi}\!\right)^{\!\!2}\!\!\frac{OdB_{\mathit{O}}(\mu_{0})}{dO}+\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{\!\!3}\!\frac{OdC_{\mathit{O}}(\mu_{0})}{dO}\!+\!{\cal O}(\alpha_{s}^{4})\right\}, (4.1)

where the σ0\sigma_{0} is a tree-level hadronic cross-section, AO,BO,COA_{O},B_{O},C_{O} are respectively the LO, NLO and NNLO coefficients, OO is the selected unintegrated variable. For the sake of simplicity, we shall refer to the perturbatively calculated differential coefficients as implicit coefficients and drop the derivative symbol, i.e.

AO(μ0)\displaystyle A_{O}(\mu_{0}) \displaystyle\equiv OdAO(μ0)dO,BO(μ0)OdBO(μ0)dO,\displaystyle\frac{OdA_{O}(\mu_{0})}{dO},\qquad B_{O}(\mu_{0})~{}~{}\equiv~{}~{}\frac{OdB_{O}(\mu_{0})}{dO},
CO(μ0)\displaystyle C_{O}(\mu_{0}) \displaystyle\equiv OdCO(μ0)dO.\displaystyle\frac{OdC_{O}(\mu_{0})}{dO}. (4.2)

We define here the intrinsic conformality as the property of a renormalizable SU(N)/U(1) gauge theory, such as QCD, which yields a particular structure of the perturbative corrections that can be made explicit by representing the perturbative coefficients using the following parametrization: 444We are neglecting here other running parameters, such as the mass terms.

AO(μ0)\displaystyle A_{O}(\mu_{0})\!\!\! =\displaystyle= A𝐶𝑜𝑛𝑓,\displaystyle\!\!\!A_{\mathit{Conf}},
BO(μ0)\displaystyle B_{O}(\mu_{0})\!\!\! =\displaystyle= B𝐶𝑜𝑛𝑓+12β0ln(μ02μI2)A𝐶𝑜𝑛𝑓,\displaystyle\!\!\!B_{\mathit{Conf}}+\frac{1}{2}\beta_{0}\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}}\right)A_{\mathit{Conf}},
CO(μ0)\displaystyle C_{O}(\mu_{0})\!\!\! =\displaystyle= C𝐶𝑜𝑛𝑓+β0ln(μ02μII2)B𝐶𝑜𝑛𝑓+14[β1+β02ln(μ02μI2)]ln(μ02μI2)A𝐶𝑜𝑛𝑓,\displaystyle\!\!\!C_{\mathit{Conf}}+\beta_{0}\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm II}^{2}}\right)B_{\mathit{Conf}}+\frac{1}{4}\left[\beta_{1}+\beta_{0}^{2}\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}}\right)\right]\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}}\right)A_{\mathit{Conf}}, (4.3)

where the A𝐶𝑜𝑛𝑓,B𝐶𝑜𝑛𝑓,C𝐶𝑜𝑛𝑓A_{\mathit{Conf}},B_{\mathit{Conf}},C_{\mathit{Conf}} are the scale-invariant Conformal Coefficients (i.e. the coefficients of each perturbative order not depending on the scale μ0\mu_{0}), while we define the μN\mu_{\rm N} as Intrinsic Conformal Scales and β0,β1\beta_{0},\beta_{1} are the first two coefficients of the β\beta-function. We recall that the implicit coefficients are defined at the scale μ0\mu_{0} and that they change according to the standard RG equations under a change of the renormalization scale according to:

AO(μr)\displaystyle A_{O}(\mu_{r})\!\!\! =\displaystyle= AO(μ0),\displaystyle\!\!\!A_{O}(\mu_{0}),
BO(μr)\displaystyle B_{O}(\mu_{r})\!\!\! =\displaystyle= BO(μ0)+12β0ln(μr2μ02)AO(μ0),\displaystyle\!\!\!B_{O}(\mu_{0})+\frac{1}{2}\beta_{0}\ln\left(\frac{\mu_{r}^{2}}{\mu_{0}^{2}}\right)A_{O}(\mu_{0}),
CO(μr)\displaystyle C_{O}(\mu_{r})\!\!\! =\displaystyle= CO(μ0)+β0ln(μr2μ02)BO(μ0)+14[β1+β02ln(μr2μ02)]ln(μr2μ02)AO(μ0).\displaystyle\!\!\!C_{O}(\mu_{0})+\beta_{0}\ln\left(\frac{\mu_{r}^{2}}{\mu_{0}^{2}}\right)B_{O}(\mu_{0})+\frac{1}{4}\left[\beta_{1}+\beta_{0}^{2}\ln\left(\frac{\mu_{r}^{2}}{\mu_{0}^{2}}\right)\right]\ln\left(\frac{\mu_{r}^{2}}{\mu_{0}^{2}}\right)A_{O}(\mu_{0}). (4.4)

It can be shown that the form of Eq. (4.3) is scale invariant and it is preserved under a change of the renormalization scale from μ0\mu_{0} to μr\mu_{r} by standard RG equations Eq. (4.4), i.e.:

AO(μr)\displaystyle A_{O}(\mu_{r})\!\!\! =\displaystyle= A𝐶𝑜𝑛𝑓,\displaystyle\!\!\!A_{\mathit{Conf}},
BO(μr)\displaystyle B_{O}(\mu_{r})\!\!\! =\displaystyle= B𝐶𝑜𝑛𝑓+12β0ln(μr2μI2)A𝐶𝑜𝑛𝑓,\displaystyle\!\!\!B_{\mathit{Conf}}+\frac{1}{2}\beta_{0}\ln\left(\frac{\mu_{r}^{2}}{\mu_{\rm I}^{2}}\right)A_{\mathit{Conf}},
CO(μr)\displaystyle C_{O}(\mu_{r})\!\!\! =\displaystyle= C𝐶𝑜𝑛𝑓+β0ln(μr2μII2)B𝐶𝑜𝑛𝑓+14[β1+β02ln(μr2μI2)]ln(μr2μI2)A𝐶𝑜𝑛𝑓.\displaystyle\!\!\!C_{\mathit{Conf}}+\beta_{0}\ln\left(\frac{\mu_{r}^{2}}{\mu_{\rm II}^{2}}\right)B_{\mathit{Conf}}+\frac{1}{4}\left[\beta_{1}+\beta_{0}^{2}\ln\left(\frac{\mu_{r}^{2}}{\mu_{\rm I}^{2}}\right)\right]\ln\left(\frac{\mu_{r}^{2}}{\mu_{\rm I}^{2}}\right)A_{\mathit{Conf}}. (4.5)

We note that the form of Eq. (4.3) is invariant and that the initial scale dependence is exactly removed by μr\mu_{r}. Extending this parametrization to all orders we achieve a scale-invariant quantity: the iCF parametrization is a sufficient condition in order to obtain a scale-invariant observable.

In order to show this property we collect together the terms identified by the same conformal coefficient, we call each set a conformal subset and extend the property to order nn:

σI\displaystyle\sigma_{\rm I} ={(αs(μ0)2π)+12β0ln(μ02μI2)(αs(μ0)2π)2\displaystyle=\left\{\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)+\frac{1}{2}\beta_{0}\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}}\right)\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{2}\right. (4.6)
+14[β1+β02ln(μ02μI2)]ln(μ02μI2)(αs(μ0)2π)3+}A𝐶𝑜𝑛𝑓\displaystyle\qquad\qquad+\left.\frac{1}{4}\!\left[\beta_{1}+\beta_{0}^{2}\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}}\right)\right]\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}}\right)\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{3}+\ldots\right\}A_{\mathit{Conf}}
σII\displaystyle\sigma_{\rm II} ={(αs(μ0)2π)2+β0ln(μ02μII2)(αs(μ0)2π)3+}B𝐶𝑜𝑛𝑓\displaystyle=\left\{\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{2}+\beta_{0}\ln\left(\frac{\mu_{0}^{2}}{\mu_{\rm II}^{2}}\right)\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{3}+\ldots\right\}B_{\mathit{Conf}}
σIII\displaystyle\sigma_{\rm III} ={(αs(μ0)2π)3+}C𝐶𝑜𝑛𝑓,\displaystyle=\left\{\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{3}+\ldots\right\}C_{\mathit{Conf}},
\displaystyle\vdots \displaystyle\qquad\qquad\boldsymbol{\cdot^{\textstyle\cdot^{\textstyle\cdot}}}
σn\displaystyle\sigma_{\rm n} ={(αs(μ0)2π)n}n𝐶𝑜𝑛𝑓.\displaystyle=\left\{\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{n}\right\}\mathcal{L}_{n\mathit{Conf}}.

In each subset we have only one intrinsic scale and only one conformal coefficient and the subsets are disjoint; thus, no mixing terms among the scales or the coefficients are introduced in this parametrization. Moreover, the structure of the subsets remains invariant under a global change of the renormalization scale, as shown from Eq. (4.5). The structure of each conformal set σI,σII,σIII,\sigma_{\rm I},\sigma_{\rm II},\sigma_{\rm III},\dots and consequently the iCF are preserved, also if we fix a different renormalization scale for each conformal subset, i.e.

(μ2μ2+β(αs)αs)σn=0.\displaystyle\left(\mu^{2}\frac{\partial}{\partial\mu^{2}}+\beta(\alpha_{s})\frac{\partial}{\partial\alpha_{s}}\right)\sigma_{\rm n}=0. (4.7)

We define here this property of Eq. 4.6 of separating an observable into the union of ordered scale-invariant disjoint subsets σI,σII,σIII,\sigma_{\rm I},\sigma_{\rm II},\sigma_{\rm III},\dots an ordered scale invariance.

In order to extend the iCF to all orders, we perform the nn\rightarrow\infty limit using the following strategy: we first perform a partial limit J/nJ_{/n}\rightarrow\infty including the higher-order corrections relative only to those β0,β1,β2,,βn2\beta_{0},\beta_{1},\beta_{2},\dots,\beta_{n-2} terms that have been determined already at order nn for each subset and we then perform the complementary n¯\bar{n} limit, which consists in including all the remaining higher-order terms. For the J/nJ_{/n} limit we have:

limJ/nσI\displaystyle\lim_{J_{/n}\rightarrow\infty}\sigma_{\rm I} \displaystyle\rightarrow (αs(μI)|n22π)A𝐶𝑜𝑛𝑓\displaystyle\left(\frac{\left.\alpha_{s}(\mu_{\rm I})\right|_{n-2}}{2\pi}\right)A_{\mathit{Conf}}
limJ/nσII\displaystyle\lim_{J_{/n}\rightarrow\infty}\sigma_{\rm II} \displaystyle\rightarrow (αs(μII)|n32π)2B𝐶𝑜𝑛𝑓\displaystyle\left(\frac{\left.\alpha_{s}(\mu_{\rm II})\right|_{n-3}}{2\pi}\right)^{2}B_{\mathit{Conf}}
limJ/nσIII\displaystyle\lim_{J_{/n}\rightarrow\infty}\sigma_{\rm III} \displaystyle\rightarrow (αs(μIII)|n42π)3C𝐶𝑜𝑛𝑓\displaystyle\left(\frac{\left.\alpha_{s}(\mu_{\rm III})\right|_{n-4}}{2\pi}\right)^{3}C_{\mathit{Conf}}
\displaystyle\vdots \displaystyle\qquad\vdots
limJ/nσn\displaystyle\lim_{J_{/n}\rightarrow\infty}\sigma_{\rm n} \displaystyle\equiv (αs(μ0)2π)nn𝐶𝑜𝑛𝑓,\displaystyle\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{n}\mathcal{L}_{n\mathit{Conf}}, (4.8)

where αs(μI)|n2\left.\alpha_{s}(\mu_{\rm I})\right|_{n-2} is the coupling calculated up to βn2\beta_{n-2} at the intrinsic scale μI\mu_{\rm I}. Given the particular ordering of the powers of the coupling, in each conformal subset we have the coefficients of the β0,,βnk1\beta_{0},\dots,\beta_{n-k-1} terms, where kk is the order of the conformal subset and the nn is the order of the highest subset with no β\beta terms. We note that the limit of each conformal subset is finite and scale invariant up to σn1\sigma_{n-1}. The remaining scale dependence is confined to the coupling of the nthn^{th} term. Any combination of the σI,,σn1\sigma_{\rm I},\dots,\sigma_{n-1} subsets is finite and scale invariant. We can now extend the iCF to all orders performing the n¯\bar{n} limit. In this limit we include all the remaining higher-order corrections. For the calculated conformal subsets this leads to defining the coupling at the same scales but including all the missing β\beta terms. Thus, each conformal subset remains scale invariant. We point out that we are not making any assumption on the convergence of the series for this limit. We thus have:

limn¯σI\displaystyle\lim_{\bar{n}\rightarrow\infty}\sigma_{\rm I} \displaystyle\rightarrow (αs(μI)2π)A𝐶𝑜𝑛𝑓\displaystyle\left(\frac{\alpha_{s}(\mu_{\rm I})}{2\pi}\right)A_{\mathit{Conf}}
limn¯σII\displaystyle\lim_{\bar{n}\rightarrow\infty}\sigma_{\rm II} \displaystyle\rightarrow (αs(μII)2π)2B𝐶𝑜𝑛𝑓\displaystyle\left(\frac{\alpha_{s}(\mu_{\rm II})}{2\pi}\right)^{2}B_{\mathit{Conf}}
limn¯σIII\displaystyle\lim_{\bar{n}\rightarrow\infty}\sigma_{\rm III} \displaystyle\rightarrow (αs(μIII)2π)3C𝐶𝑜𝑛𝑓\displaystyle\left(\frac{\alpha_{s}(\mu_{\rm III})}{2\pi}\right)^{3}C_{\mathit{Conf}}
\displaystyle\vdots \displaystyle\qquad\qquad\vdots
limn¯σn\displaystyle\hphantom{ConformalLimit}\lim_{\bar{n}\rightarrow\infty}\sigma_{\rm n} \displaystyle\equiv limn(αs(μ0)2π)nn𝐶𝑜𝑛𝑓Conformal Limit,\displaystyle\lim_{n\rightarrow\infty}\left(\frac{\alpha_{s}(\mu_{0})}{2\pi}\right)^{n}\!\!\mathcal{L}_{n\mathit{Conf}}\rightarrow\hbox{Conformal Limit}, (4.9)

where here now αs(μI)\alpha_{s}(\mu_{\rm I}) is the complete coupling determined at the same scale μI\mu_{\rm I}. Equation (4.9) shows that the entire renormalization scale dependence has been completely removed. In fact, neither the intrinsic scales μN\mu_{\rm N} nor the conformal coefficients A𝐶𝑜𝑛𝑓,B𝐶𝑜𝑛𝑓,C𝐶𝑜𝑛𝑓,,n𝐶𝑜𝑛𝑓,A_{\mathit{Conf}},B_{\mathit{Conf}},C_{\mathit{Conf}},\dots,\mathcal{L}_{n\mathit{Conf}},\dots depend on the particular choice of the initial scale. The only term with a residual μ0\mu_{0} dependence is the nn-th term, but this dependence cancels in the limit nn\rightarrow\infty. The scale dependence is totally confined to the coupling αs(μ0)\alpha_{s}(\mu_{0}) and its behavior does not depend on the particular choice of any scale μ0\mu_{0} in the perturbative region, i.e. limnαs(μ0)nan\lim_{n\rightarrow\infty}\alpha_{s}(\mu_{0})^{n}\sim a^{n} with a<1a<1. Hence, the limit of limnσn\lim_{n\rightarrow\infty}\sigma_{\rm n} depends only on the properties of the theory and not on the scale of the coupling in the perturbative regime. The proof given here shows that the iCF is sufficient to have a scale-invariant observable and it does not depend on the particular convergence of the series.

In order to show the necessary condition, we separate the two cases of a convergent series and an asymptotic expansion. For the first case the necessary condition stems directly from the uniqueness of the iCF form, since given a finite limit and the scale invariance, any other parametrization can be reduced to the iCF by means of appropriate transformations in agreement with the RG equations. For the second case, we have that an asymptotic expansion though not convergent, can be truncated at a certain order nn, which is the case of Eq. (4.6). Given the particular structure of the iCF we can perform the first partial limit J/nJ_{/n} and we would achieve a finite and scale-invariant prediction, σN1=Σi=1n1σi\sigma_{N-1}=\Sigma_{i=1}^{n-1}\sigma_{i}, for a truncated asymptotic expansion, as shown in Eq. (4.8). Given the truncation of the series in the region of maximum of convergence the nn-th term would be reduced to the lowest value and so the scale dependence of the observable would reach its minimum. Given the finite and scale-invariant limit σN1\sigma_{N-1} we conclude that the iCF is unique and thus necessary for an ordered scale-invariant truncated asymptotic expansion up to the nn-th order.

We point out that in general the iCF form is the most general and irreducible parametrization that leads to scale invariance; other parametrization are forbidden, since if we were to introduce more scales 555Here we refer to the form of Eq. (4.3). In principle, it is possible to write other parametrizations preserving the scale invariance, but these can be reduced to the iCF in agreement with the RG equations. into the logarithms of one subset, we would spoil the invariance under the RG transformation and we could not achieve Eq. (4.5), while on the other hand no scale dependence can be introduced into the intrinsic scales since it would remain in the observable already in the first partial limit J/nJ_{/n} and it could not be eliminated. The conformal coefficients are conformal at each order by definition; thus, they do not depend on the renormalization scale and they do not have a perturbative expansion. Hence the iCF is a necessary and sufficient condition for scale invariance.

4.2   Comments on the iCF and ordered scale invariance

The iCF parametrization can stem either from an inner property of the theory, the iCF, or from direct parametrization of the scale-invariant observable. In both cases the iCF parametrization makes the scale dependence of the observable explicit and it exactly preserves the scale invariance. The iCF parametrization is invariant with respect to the choice of initial scale μ0\mu_{0}, this implies that the same calculation performed choosing different arbitrary initial scales, μ0,μ0\mu_{0},\,\mu_{0}^{\prime} leads to the same result in the limit J/nJ_{/n}, a limit that is scale and scheme independent. The iCF is also strongly motivated by the renormalizability of QCD and by the uniqueness of the β\beta-function in a given scheme; i.e. two different βi,βi\beta_{i},\,\beta_{i}^{\prime} do not occur in a perturbative calculation at any order in one RS and the UV divergencies are cancelled by redefinition of the same parameters at lowest and higher orders. We remark that the conservation of the iCF form in one observable is strongly related to the validity of the RG transformations; we thus expect the iCF to be well preserved in the deep Euclidean region.

Once we have defined an observable in the iCF-form, we have not only the scale invariance of the entire observable, but also the ordered scale invariance (i.e. the scale invariance of each subset σn\sigma_{\rm n} or σN1\sigma_{\rm N-1}). The latter property is crucial in order to obtain scale-invariant observables independently from the particular kinematic region and independently from the starting order of the observable or the order of the truncation of the series. Since in general, a theory is blind with respect to the particular observable/process that we might investigate, the theory should preserve the ordered scale invariance in order to always define scale-invariant observables. Hence if the iCF is an inner property of the theory, it leads to implicit coefficients that are neither independent nor conformal. This is made explicit in Eq. (4.3), but it is hidden in the perturbative calculations in the case of the implicit coefficients. For instance, the presence of the iCF clearly reveals itself when a particular kinematic region is approached and the AOA_{O} becomes null. This would cause a breaking of the scale invariance since a residual initial scale dependence would remain in the observable in the higher-order coefficients. The presence of the iCF solves this issue by leading to the correct redefinition of all the coefficients at each order preserving the correct scale invariance exactly. Thus, in the case of a scale-invariant observable OO, defined according to the implicit form (Eq. (4.1)), by the coefficients {AO,BO,CO,..,OO,},\{A_{O},B_{O},C_{O},..,O_{O},\dots\}, it cannot simply undergo the change {0,BO,CO,..,OO,}\rightarrow\{0,B_{O},C_{O},..,O_{O},\dots\}, since this would break the scale invariance. In order to preserve the scale invariance, we must redefine the coefficients {A~O=0,B~O,C~O,..,O~O,}\{\tilde{A}_{O}=0,\tilde{B}_{O},\tilde{C}_{O},..,\tilde{O}_{O},\dots\} cancelling out all the initial scale dependence originating from the LO coefficient AOA_{O} at all orders. This is equivalent to subtracting out an entire invariant conformal subset σI\sigma_{\rm I} related to the coefficient A𝐶𝑜𝑛𝑓A_{\mathit{Conf}} from the scale-invariant observable OO. This mechanism is clear in the case of the explicit form of the iCF, Eq. (4.3), where, if A𝐶𝑜𝑛𝑓=0A_{\mathit{Conf}}=0, then the entire conformal subset is null and the scale invariance is preserved.

We stress that the conformal coefficients may acquire all possible values without breaking scale invariance, they contain the essential information on the physics of the process, while all the correlation factors can be reabsorbed into the renormalization scales as shown by the PMC method [46, 47, 49, 50]. Hence, if a theory has the property of ordered scale invariance, it exactly preserves the scale invariance of observables independently of the process, the kinematics and the starting order of the observable. We stress that if a theory has intrinsic conformality, all renormalized quantities, such as cross sections, can be parametrized with the iCF-form. This property should be preserved by the renormalization scheme or by the definition of IR safe quantities and it should also be preserved in observables defined in effective theories. The iCF shows that point (3) of the BLM/PMC approach (Section 3) can be improved by eliminating the perturbative expansion of the BLM/PMC scales, leading to a scale- and scheme-invariant result. We remark though that the perturbative corrections in the BLM or PMCm scales are suppressed in the perturbative region.

4.3   The PMC

We introduce here a new method for eliminating the scale-setting ambiguity in single variable scale-invariant distributions, which we call PMC. This method is based on the original PMC principle and agrees with all the different PMC formulations for the PMC scales at lowest order. The core of the PMC is essentially the same for all BLM-PMC prescriptions, i.e. the effective running-coupling value and hence its renormalization scale at each order is determined by the β0\beta_{0}-term of the next-higher order, or equivalently by the intrinsic conformal scale μN\mu_{\rm N}. The PMC preserves the iCF and thus the scale and scheme invariance, absorbing an infinite set of β\beta-terms to all orders.

This method differs from the other PMC prescriptions since, due to the presence of the intrinsic conformality, no perturbative correction in αs\alpha_{s} needs to be introduced at higher orders in the PMC scales. Given that all the β\beta-terms of a single conformal subset are included in the renormalization scale already with the definition at lowest order, no initial scale or scheme dependences are left due to the unknown β\beta-terms in each subset. The PMC scale of each subset can be unambiguously determined by β0\beta_{0}-term of each order, we stress that all logarithms of each subset have the same argument and all the differences arising at higher orders have to be included only in the conformal coefficients. Reabsorbing all the β\beta-terms into the scale also eliminates the n!β0nαsnn!\beta_{0}^{n}\alpha_{s}^{n} terms (related to renormalons [23]); thus, the precision is improved and the perturbative QCD predictions can be extended to a wider range of values. The initial scale dependence is totally confined in the unknown PMC scale of the last order of accuracy (i.e. up to NNLO case in the αs(μ0)3\alpha_{s}(\mu_{0})^{3}). Thus, if we fix the renormalization scale independently to the proper intrinsic scale for each subset μN\mu_{\rm N}, we end up with a perturbative sum of totally conformal contributions up to the order of accuracy:

1σ0Odσ(μI,μII,μIII)dO\displaystyle\frac{1}{\sigma_{0}}\frac{Od\sigma(\mu_{\rm I},\mu_{\rm II},\mu_{\rm III})}{dO} =\displaystyle= {αs(μI)2πOdA𝐶𝑜𝑛𝑓dO+(αs(μII)2π)2OdB𝐶𝑜𝑛𝑓dO\displaystyle\left\{\frac{\alpha_{s}(\mu_{\rm I})}{2\pi}\frac{OdA_{\mathit{Conf}}}{dO}+\left(\frac{\alpha_{s}(\mu_{\rm II})}{2\pi}\right)^{2}\frac{OdB_{\mathit{Conf}}}{dO}\right. (4.10)
+(αs(μIII)2π)3OdC𝐶𝑜𝑛𝑓dO}+𝒪(αs4).\displaystyle\hskip 100.00015pt+\left.\left(\frac{\alpha_{s}(\mu_{\rm III})}{2\pi}\right)^{3}\frac{OdC_{\mathit{Conf}}}{dO}\right\}+{\cal O}(\alpha_{s}^{4}).

At this order, the last scale is set to the physical scale QQ, i.e. μIII=μ0=Q\mu_{\rm III}=\mu_{0}=Q.

4.4   The iCF coefficients and scales: a new “How-To” method

We describe here how all the coefficients of Eq. (4.3) can be identified from either a numerical or analytical perturbative calculation. This method applies in general to any perturbative calculation once results for the different color factors are kept separate; however, we refer to the particular case of the NNLO thrust distribution results calculated in Refs. [173, 174] for the purpose. Since the leading order is already (A𝐶𝑜𝑛𝑓A_{\mathit{Conf}}) void of β\beta-terms, we start with NLO coefficients. A general numerical/analytical calculation keeps tracks of all the color factors and the respective coefficients:

BO(Nf)=CF[CABONc+CFBOCF+TFNfBONf]\displaystyle B_{O}(N_{f})=C_{F}\left[C_{A}B_{O}^{N_{c}}+C_{F}B_{O}^{C_{F}}+T_{F}N_{f}B_{O}^{N_{f}}\right] (4.11)

where CF=(Nc21)2NcC_{F}=\frac{\left(N_{c}^{2}-1\right)}{2N_{c}}, CA=NcC_{A}=N_{c} and TF=1/2.T_{F}=1/2. The dependence on NfN_{f} is made explicit here for sake of clarity. We can determine the conformal coefficient B𝐶𝑜𝑛𝑓B_{\mathit{Conf}} of the NLO order straightforwardly, by fixing the number of flavors NfN_{f} in order to kill the β0\beta_{0} term:

B𝐶𝑜𝑛𝑓\displaystyle B_{\mathit{Conf}} =\displaystyle= BO(Nf332),\displaystyle B_{O}\left(N_{f}\equiv\frac{33}{2}\right),
Bβ0logμ02μI2\displaystyle B_{\beta_{0}}\equiv\log\frac{\mu_{0}^{2}}{\mu_{\rm I}^{2}} =\displaystyle= 2BOB𝐶𝑜𝑛𝑓β0A𝐶𝑜𝑛𝑓.\displaystyle 2\frac{B_{O}-B_{\mathit{Conf}}}{\beta_{0}A_{\mathit{Conf}}}. (4.12)

We would achieve the same results in the usual PMC way; i.e. by identifying the NfN_{f} coefficient with the β0\beta_{0} term and then determining the conformal coefficient. Both methods are consistent and results for the intrinsic scales and the coefficients are in perfect agreement. At NNLO a general coefficient is composed of the contribution of six different color factors:

CO(Nf)\displaystyle C_{O}(N_{f}) =\displaystyle= CF4{Nc2CONc2+CONc0+1Nc2CO1Nc2\displaystyle\frac{C_{F}}{4}\left\{N_{c}^{2}C_{O}^{N_{c}^{2}}+C_{O}^{N_{c}^{0}}+\frac{1}{N_{c}^{2}}C_{O}^{\frac{1}{N_{c}^{2}}}\right. (4.13)
+NfNcCONfNc+NfNcCONf/Nc+Nf2CONf2}.\displaystyle\qquad\qquad+\left.N_{f}N_{c}\cdot C_{O}^{N_{f}N_{c}}+\frac{N_{f}}{N_{c}}C_{O}^{N_{f}/N_{c}}+N_{f}^{2}C_{O}^{N_{f}^{2}}\right\}.

In order to identify all the terms of Eq. (4.3), we notice first that the coefficients of the terms β02\beta_{0}^{2} and β1\beta_{1} are already given by the NLO coefficient Bβ0B_{\beta_{0}}; we thus need to determine only the β0\beta_{0} and the conformal C𝐶𝑜𝑛𝑓C_{\mathit{Conf}} terms. In order to determine the latter coefficients, we use the same procedure used for the NLO; i.e. we set the number of flavors Nf33/2N_{f}\equiv 33/2 in order to remove all the β0\beta_{0} terms. We then have

C𝐶𝑜𝑛𝑓\displaystyle C_{\mathit{Conf}} =\displaystyle= CO(Nf332)14β¯1Bβ0A𝐶𝑜𝑛𝑓,\displaystyle C_{O}\left(N_{f}\equiv\frac{33}{2}\right)-\frac{1}{4}\overline{\beta}_{1}B_{\beta_{0}}A_{\mathit{Conf}},
Cβ0log(μ02μII2)\displaystyle C_{\beta_{0}}\equiv\log\left(\frac{\mu_{0}^{2}}{\mu_{\rm II}^{2}}\right) =\displaystyle= 1β0B𝐶𝑜𝑛𝑓(COC𝐶𝑜𝑛𝑓14β02Bβ02A𝐶𝑜𝑛𝑓14β1Bβ0A𝐶𝑜𝑛𝑓),\displaystyle\frac{1}{\beta_{0}B_{\mathit{Conf}}}\left(C_{O}-C_{\mathit{Conf}}-\frac{1}{4}\beta_{0}^{2}B_{\beta_{0}}^{2}A_{\mathit{Conf}}-\frac{1}{4}\beta_{1}B_{\beta_{0}}A_{\mathit{Conf}}\!\right)\!, (4.14)

with β¯1β1(Nf=33/2)=107\overline{\beta}_{1}\equiv\beta_{1}(N_{f}=33/2)=-107. Up to accuracy 𝒪(αs5)\mathcal{O}(\alpha_{s}^{5}), we have:

D𝐶𝑜𝑛𝑓\displaystyle D_{\mathit{Conf}} =\displaystyle= DO(Nf332)18β¯2Bβ0A𝐶𝑜𝑛𝑓12β¯1Cβ0B𝐶𝑜𝑛𝑓,\displaystyle D_{O}\left(N_{f}\equiv\frac{33}{2}\right)-\frac{1}{8}\overline{\beta}_{2}B_{\beta_{0}}A_{\mathit{Conf}}-\frac{1}{2}\overline{\beta}_{1}C_{\beta_{0}}B_{\mathit{Conf}},
Dβ0log(μ02μIII2)\displaystyle D_{\beta_{0}}\equiv\log\left(\frac{\mu_{0}^{2}}{\mu_{\rm III}^{2}}\right) =\displaystyle= 23β0C𝐶𝑜𝑛𝑓[DOD𝐶𝑜𝑛𝑓18(β03Bβ03+52β0β1Bβ02+β2Bβ0)A𝐶𝑜𝑛𝑓\displaystyle\frac{2}{3\beta_{0}C_{\mathit{Conf}}}\left[D_{O}-D_{\mathit{Conf}}-\frac{1}{8}\left(\beta_{0}^{3}B_{\beta_{0}}^{3}+\frac{5}{2}\beta_{0}\beta_{1}B_{\beta_{0}}^{2}+\beta_{2}B_{\beta_{0}}\right)A_{\mathit{Conf}}\right. (4.15)
14(3β02Cβ02+2β1Cβ0)B𝐶𝑜𝑛𝑓],\displaystyle\hskip 140.00021pt-\left.\frac{1}{4}\left(3\beta_{0}^{2}C_{\beta_{0}}^{2}+2\beta_{1}C_{\beta_{0}}\right)B_{\mathit{Conf}}\right],

with β¯2β2(Nf=33/2).\overline{\beta}_{2}\equiv\beta_{2}(N_{f}=33/2).

This procedure may be extended to all orders and one may decide whether to cancel β0\beta_{0}, β1\beta_{1} or β2\beta_{2} by fixing the appropriate number of flavors. The results can be compared to exactly determine all the coefficients. We point out that extending the intrinsic conformality to all orders, we can at this stage predict the coefficients of all the color factors of the higher orders related to the β\beta-terms, except those related to the higher-order conformal coefficients and β0\beta_{0}-terms (e.g. at N4N^{4}LO, E𝐶𝑜𝑛𝑓E_{\mathit{Conf}} and Eβ0E_{\beta_{0}}).

4.5   Comment on the PMC/PMC scales

PMC scales stem directly from the renormalization of the UV-divergent diagrams. As pointed out in Sec. 2.1, the finite part of the divergent integrals contribute to the β\beta-terms. In fact, these coefficients derive from the UV-divergent diagrams connected with the running of the coupling constant and not from UV-finite diagrams. UV-finite NFN_{F} terms may arise but would not contribute to the β\beta-terms. These terms can be easily identified by the kinematic constraint at lowest order or by checking deviations of the nfn_{f} coefficients from the iCF form. In fact, only the NfN_{f} terms coming from UV-divergent diagrams, depending dynamically on the virtuality of the underlying quark and gluon subprocesses have to be considered as β\beta-terms and they would determine the intrinsic conformal scales. In general, each μN\mu_{\rm N} is an independent function of the physical scale of the process s\sqrt{s} (or t,u,\sqrt{t},\sqrt{u},\dots), of the selected variable OO and it varies with the number of colors NcN_{c} mainly due to the gggggg and gggggggg vertices. The latter terms arise at higher orders only in a non-Abelian theory, but they are not expected to spoil the iCF-form. We stress that iCF applies to scale-invariant single-variable differential distributions, in case one is interested in the renormalization of a particular diagram, e.g. the gggggg vertex, contributions from different β\beta-terms should be singled out in order to identify the respective intrinsic conformal scale consistently with the renormalization of the non-Abelian gggggg vertex, as shown in [175].

In the renormalization procedure of gauge theories, one first identifies the UV singularities of a scattering amplitude, which appear as poles using dimensional regularization. The UV-divergent contributions are absorbed into renormalization constants ZiZ_{i} adopting a particular scheme (e.g. the MS¯\overline{MS} scheme). This cancels the UV divergences and at the same time defines the finite part of the loop integral.

The finite parts, such as the finite βi\beta_{i} and γi\gamma_{i} contributions, are associated with the renormalization of the running coupling and running mass, respectively. The βi\beta_{i} terms can then be summed into the running coupling using the standard renormalization group equations; this is basically the core of the BLM-PMC scale-setting procedure that is analogous to the Gell-Mann–Low scheme in QED. This procedure eliminates the scale ambiguity and reabsorbs the scheme dependence at once into the effective running coupling up to the computed order. In addition, the factorial renormalon divergence is eliminated. One thus can use the same renormalization procedure for QED, QCD and EW in a grand unified theory,

Given the renormalizability of QCD, once the coupling is renormalized all the vertices are finite, but this does not cancel the contributions of the finite parts of the integrals, i.e. the βn\beta_{n} terms, which define the PMC scale for each vertex (3g3g, 4g4g, ccgccg, qqgqqg).

5   PMC results for thrust and the CC-parameter

The thrust distribution and event-shape variables are fundamental tools for probing the geometrical structure of a given process at colliders. Being observables that are exclusive enough with respect to the final state, they allow for a deeper geometrical analysis of the process and are also particularly suited for measurement of the strong coupling αs\alpha_{s} [176].

Given the high-precision data collected at LEP and SLAC [177, 178, 179, 180, 181], refined calculations are crucial in order to extract information to the highest possible precision. Although extensive studies on these observables have been produced during the last few decades, including higher-order corrections from next-to-leading order (NLO) calculations [182, 183, 184, 185, 186, 187] to the next-to-next-to-leading order(NNLO) [188, 189, 190, 173, 174] and including resummation of the large logarithms [31, 30], the theoretical predictions are still affected by significant theoretical uncertainties that are related to large renormalization scale ambiguities. In the particular case of the three-jet event-shape distributions the conventional practice of CSS leads to results that do not match the experimental data and the extracted values of αs\alpha_{s} deviate from the world average [126].

The thrust (TT) and CC-parameter (CC) are defined by

T=maxn(i|pin|i|pi|),\displaystyle T=\max\limits_{\vec{n}}\left(\frac{\sum_{i}|\vec{p}_{i}\cdot\vec{n}|}{\sum_{i}|\vec{p}_{i}|}\right), (5.1)
C=32i,j|pi||pj|sin2θij(i|pi|)2,\displaystyle C=\frac{3}{2}\frac{\sum_{i,j}|\vec{p_{i}}||\vec{p_{j}}|\sin^{2}\theta_{ij}}{\left(\sum_{i}|\vec{p_{i}}|\right)^{2}}, (5.2)

where the sum runs over all particles in the hadronic final state and pi\vec{p}_{i} denotes the three-momentum of particle ii. The unit vector n\vec{n} is varied to maximize thrust TT; the corresponding n\vec{n} is called the thrust axis and denoted by nT\vec{n}_{T}. The variable (1T)(1-T) is often used, which for the LO of 3-jet production is restricted to the range (0<1T<1/3)(0<1-T<1/3). We have a back-to-back or a spherically symmetric event at T=1T=1 and at T=2/3T=2/3 respectively. For the CC-parameter, θij\theta_{ij} is the angle between pi\vec{p_{i}} and pj\vec{p_{j}}. At LO for 3-jet production the CC-parameter is restricted by kinematics to the range 0C0.750\leq C\leq 0.75.

In general, a normalized IR-safe single-variable observable, such as the thrust distribution for e+e3e^{+}e^{-}\rightarrow 3\,jets [191, 192], is the sum of pQCD contributions calculated up to NNLO at the initial renormalization scale μ0=s=MZ\mu_{0}=\sqrt{s}=M_{Z}:

1σtotOdσ(μ0)dO\displaystyle\frac{1}{\sigma_{tot}}\!\frac{Od\sigma(\mu_{0})}{dO}\! =\displaystyle= {x0OdA¯O(μ0)dO+x02OdB¯O(μ0)dO+x03OdC¯O(μ0)dO+𝒪(αs4)},\displaystyle\left\{x_{0}\cdot\frac{Od\bar{A}_{\mathit{O}}(\mu_{0})}{dO}+x_{0}^{2}\cdot\frac{Od\bar{B}_{\mathit{O}}(\mu_{0})}{dO}+x_{0}^{3}\cdot\frac{Od\bar{C}_{\mathit{O}}(\mu_{0})}{dO}+{\cal O}(\alpha_{s}^{4})\right\}, (5.3)

where x(μ)αs(μ)/(2π)x(\mu)\equiv\alpha_{s}(\mu)/(2\pi), OO is the selected event-shape variable, σ\sigma the cross-section of the process,

σtot=σ0(1+x0Atot+x02Btot+𝒪(αs3))\sigma_{tot}=\sigma_{0}\Big{(}1+x_{0}A_{tot}+x_{0}^{2}B_{tot}+{\cal O}\big{(}\alpha_{s}^{3}\big{)}\Big{)}

is the total hadronic cross-section and A¯O,B¯O,C¯O\bar{A}_{O},\bar{B}_{O},\bar{C}_{O} are respectively the normalized LO, NLO and NNLO coefficients:

A¯O\displaystyle\bar{A}_{O} =\displaystyle= AO\displaystyle A_{O}
B¯O\displaystyle\bar{B}_{O} =\displaystyle= BOAtotAO\displaystyle B_{O}-A_{tot}A_{O} (5.4)
C¯O\displaystyle\bar{C}_{O} =\displaystyle= COAtotBO(BtotAtot2)AO,\displaystyle C_{O}-A_{tot}B_{O}-\left(B_{tot}-A_{tot}^{2}\right)A_{O},

where AO,BO,COA_{O},B_{O},C_{O} are the coefficients normalized to the tree-level cross-section σ0\sigma_{0} calculated by Monte Carlo (see e.g. the EERAD and Event2 codes [188, 189, 190, 173, 174]) and A𝑡𝑜𝑡,B𝑡𝑜𝑡A_{\mathit{tot}},B_{\mathit{tot}} are

A𝑡𝑜𝑡\displaystyle A_{\mathit{tot}} =\displaystyle= 32CF;\displaystyle\frac{3}{2}C_{F};
B𝑡𝑜𝑡\displaystyle B_{\mathit{tot}} =\displaystyle= CF4Nc+34CFβ02(118ζ(3))38CF2,\displaystyle\frac{C_{F}}{4}N_{c}+\frac{3}{4}C_{F}\frac{\beta_{0}}{2}\big{(}11-8\zeta(3)\big{)}-\frac{3}{8}C_{F}^{2}, (5.5)

where ζ\zeta is the Riemann zeta function.

In general, according to CSS the renormalization scale is set to μ0=s=MZ\mu_{0}=\sqrt{s}=M_{Z} and theoretical uncertainties are evaluated using standard criteria. In this case, we have used the definition of the parameter δ\delta given in Ref. [189]; we define the average error for the event-shape variable distributions as:

δ¯=1NiNmaxμ(σi(μ))minμ(σi(μ))2σi(μ=MZ),\bar{\delta}=\frac{1}{N}\sum_{i}^{N}\frac{{\rm max}_{\mu}\big{(}\sigma_{i}(\mu)\big{)}-{\rm min}_{\mu}\big{(}\sigma_{i}(\mu)\big{)}}{2\sigma_{i}(\mu{=}M_{Z})}, (5.6)

where ii is the index of the bin and NN is the total number of bins, the renormalization scale is varied in the range μ[MZ/2,2MZ]\mu\in[M_{Z}/2,2M_{Z}].

5.1   The PMC scales at LO and NLO

According to the PMC prescription, we fix the renormalization scale to μN\mu_{\rm N} at each order absorbing all the β\beta terms into the coupling. We notice a small mismatch between the zeroes of the conformal coefficient B𝐶𝑜𝑛𝑓B_{\mathit{Conf}} and those of the remaining β0\beta_{0} term in the numerator (the formula is shown in Eq. (4.14)). Due to our limited knowledge of the strong coupling at low energies, in order to avoid singularities in the NLO scale μII\mu_{\rm II}, we introduce a regularization that leads to a finite scale μ~II\tilde{\mu}_{\rm II} over the entire range of values of the variable (1T)(1-T). These singularities might be due either to the presence of UV finite NFN_{F} terms or to the logarithmic behavior of the conformal coefficients when low values of the variable 1T1-T are approached. Large logarithms arise from the IR-divergence cancellation procedure and they can be resummed in order to restore a predictive perturbative regime [25, 26, 27, 30, 31]. We point out that IR cancellation should not spoil the iCF property. Whether this is an actual deviation from the iCF-form must be investigated further. However, since the discrepancies between the coefficients are rather small, we introduce a regularization method based on redefinition of the norm of the coefficient B𝐶𝑜𝑛𝑓B_{\mathit{Conf}} in order to cancel out these singularities in the μII\mu_{\rm II} scale. This regularization is consistent with the PMC principle and up to the accuracy of the calculation it does not introduce any bias effect in the results or any ambiguity in the NLO-PMC scale. All differences introduced by the regularization would enter at N3LO\rm N^{3}LO accuracy and they may be reabsorbed later in the higher-order PMC scales. For the PMC scales, μN\mu_{\rm N} we thus obtain

μI\displaystyle\mu_{\rm I} =\displaystyle= sefsc12Bβ0(1T)<0.33,\displaystyle\sqrt{s}\cdot e^{f_{sc}-\frac{1}{2}B_{\beta_{0}}}\hskip 123.76965pt(1-T)<0.33, (5.7)
μ~II\displaystyle\tilde{\mu}_{\rm II} =\displaystyle= {sefsc12Cβ0B𝐶𝑜𝑛𝑓B𝐶𝑜𝑛𝑓+ηA𝑡𝑜𝑡A𝐶𝑜𝑛𝑓(1T)<0.33,sefsc12Cβ0(1T)>0.33,\displaystyle\left\{\begin{array}[]{lr}\sqrt{s}\cdot e^{f_{sc}-\frac{1}{2}C_{\beta_{0}}\cdot\frac{B_{\mathit{Conf}}}{B_{\mathit{Conf}}+\eta\cdot A_{\mathit{tot}}A_{\mathit{Conf}}}}\hskip 42.67912pt(1-T)<0.33,\\[3.22916pt] \sqrt{s}\cdot e^{f_{sc}-\frac{1}{2}C_{\beta_{0}}}\hskip 110.96556pt(1-T)>0.33,\end{array}\right. (5.10)

where s=MZ\sqrt{s}=M_{Z} and the third scale is set to μIII=μ0=s\mu_{\rm III}=\mu_{0}=\sqrt{s}. The renormalization scheme factor for the QCD results is set to fsc0f_{sc}\equiv 0. This scheme factor also reabsorbs the scheme difference into the renormalization scale and is related to the particular choice of the scale parameter Λ\Lambda as discussed in Section 2.3. The coefficients Bβ0,Cβ0B_{\beta_{0}},C_{\beta_{0}} are the coefficients related to the β0\beta_{0}-terms of the NL and NNL perturbative order of the thrust distribution respectively. They are determined from the calculated AO,BO,COA_{O},B_{O},C_{O} coefficients.

The η\eta parameter is a regularization term to cancel the singularities of the NLO scale, μII\mu_{\rm II}, in the range (1T)<0.33(1-T)<0.33, depending on non-matching zeroes between numerator and denominator in Cβ0C_{\beta_{0}}. In general, this term is not mandatory for applying the PMC, it is necessary only in case one is interested in applying the method over the entire range covered by thrust, or any other observable. Its value has been determined as η=3.51\eta=3.51 for the thrust distribution and it introduces no bias effects up to the accuracy of the calculations and the related errors are totally negligible up to this stage.

We point out that in the region (1T)>0.33(1-T)>0.33 we have a clear example of intrinsic conformality-iCF where the kinematic constraints set the A𝐶𝑜𝑛𝑓=0A_{\mathit{Conf}}=0. According to Eq. (4.5) setting the A𝐶𝑜𝑛𝑓=0A_{\mathit{Conf}}=0 the entire conformal subset σI\sigma_{\rm I} becomes null. In this case all the β\beta terms at NLO and NNLO disappear except the β0\beta_{0}-term at NNLO, which determines the μII\mu_{\rm II} scale. The surviving nfn_{f} terms at NLO or the nf2n_{f}^{2} at NNLO are related to the finite NFN_{F}-term at NLO and to the mixed NfNFN_{f}\cdot N_{F} term arising from BOβ0B_{O}\cdot\beta_{0} at NNLO. Using the parametrization with explicit nfn_{f} terms, we have for (1T)>0.33(1-T)>0.33:

AO\displaystyle A_{O} =\displaystyle= 0,\displaystyle 0,
BO\displaystyle B_{O} =\displaystyle= B0+B1NF,\displaystyle B_{0}+B_{1}\cdot N_{F},
CO\displaystyle C_{O} =\displaystyle= C0+C1nf+C2NfNF.\displaystyle C_{0}+C_{1}\cdot n_{f}+C_{2}\cdot N_{f}\cdot N_{F}. (5.11)

we can determine μ~II\tilde{\mu}_{\rm II} for the region (1T)>0.33(1-T)>0.33 as shown in Eq. (5.10):

Cβ0=(C1113CAB123B0)C_{\beta_{0}}=\left(\frac{C_{1}}{\frac{11}{3}C_{A}B_{1}-\frac{2}{3}B_{0}}\right) (5.12)

by identifying the β0\beta_{0}-term at NNLO. The LO and NLO PMC scales are shown in Fig. 5.1.

Refer to caption
Figure 5.1: The LO-PMC (solid red) and the NLO-PMC (dashed black) scales for thrust. [89]

We notice that the two PMC scales have similar behaviors in the range (1T)<0.33(1-T)<0.33 and that the LO-PMC scale agrees with the PMC scale used in Ref. [87]. Small numerical fluctuations are visible in the NLO-PMC scale as unphysical "kinks", due to the Monte-Carlo simulation results of Ref.[174]. In the range (1T)<0.33(1-T)<0.33, where they are more evident, an interpolation has been performed. However, for the case of thrust, these have a negligible effect on the final distribution and do not depend on the method used for setting the scale, e.g. the PMC, but on the method used for performing the calculations (i.e. the Monte-Carlo [174]).

The PMC method totally eliminates both the ambiguity in the choice of the renormalization scale and the scheme dependence to all orders in QCD.

5.2   NNLO thrust distribution results

We use here the results of Ref. [173, 174] and for the running coupling αs(Q)\alpha_{s}(Q) we use the RunDec program [193]. In order to normalize the thrust distribution consistently, we expand the denominator in α0αs(μ0)\alpha_{0}\equiv\alpha_{s}(\mu_{0}) while the numerator has the couplings renormalized at different PMC scales αIαs(μI)\alpha_{I}\equiv\alpha_{s}(\mu_{\rm I}) and αIIαs(μ~II)\alpha_{II}\equiv\alpha_{s}(\tilde{\mu}_{II}). We point out here that the proper normalization would be given by the integration of the total cross-section after renormalization with the PMC scales, nonetheless since the PMC prescription only involves absorption of higher-order terms into the scales, the difference would be within the accuracy of the calculations, i.e. 𝒪(αs4(μ0))\sim\mathcal{O}(\alpha_{s}^{4}(\mu_{0})). Equation (5.3) becomes:

1σtotOdσ(μI,μ~II,μ0)dO={σ¯I+σ¯II+σ¯III+𝒪(αs4)},\frac{1}{\sigma_{tot}}\,\frac{Od\sigma(\mu_{\rm I},\tilde{\mu}_{\rm II},\mu_{0})}{dO}=\left\{\overline{\sigma}_{\rm I}+\overline{\sigma}_{\rm II}+\overline{\sigma}_{\rm III}+{\cal O}(\alpha_{s}^{4})\right\}, (5.13)

where the σ¯N\overline{\sigma}_{N} are normalized subsets that are given by:

σ¯I\displaystyle\overline{\sigma}_{\rm I} =\displaystyle= A𝐶𝑜𝑛𝑓xI\displaystyle A_{\mathit{Conf}}\cdot x_{\rm I}
σ¯II\displaystyle\overline{\sigma}_{\rm II} =\displaystyle= (B𝐶𝑜𝑛𝑓+ηAtotA𝐶𝑜𝑛𝑓)xII2ηAtotA𝐶𝑜𝑛𝑓x02AtotA𝐶𝑜𝑛𝑓x0xI\displaystyle\big{(}B_{\mathit{Conf}}+\eta A_{\textrm{tot}}A_{\mathit{Conf}}\big{)}\cdot x_{\rm II}^{2}-\eta A_{\textrm{tot}}A_{\mathit{Conf}}\cdot x_{0}^{2}-A_{\textrm{tot}}A_{\mathit{Conf}}\cdot x_{0}x_{\rm I}
σ¯III\displaystyle\overline{\sigma}_{\rm III} =\displaystyle= (C𝐶𝑜𝑛𝑓AtotB𝐶𝑜𝑛𝑓(BtotAtot2)A𝐶𝑜𝑛𝑓)x03,\displaystyle\left(C_{\mathit{Conf}}-A_{\textrm{tot}}B_{\mathit{Conf}}\!-\!(B_{\textrm{tot}}-A_{\textrm{tot}}^{2})A_{\mathit{Conf}}\right)\cdot x_{0}^{3}, (5.14)

A𝐶𝑜𝑛𝑓,B𝐶𝑜𝑛𝑓,C𝐶𝑜𝑛𝑓A_{\mathit{Conf}},B_{\mathit{Conf}},C_{\mathit{Conf}} are the scale-invariant conformal coefficients (i.e. the coefficients of each perturbative order not depending on the scale μr\mu_{r}) while xI,xII,x0x_{\rm I},x_{\rm II},x_{0} are the couplings determined at the μI,μ~II,μ0\mu_{\rm I},\tilde{\mu}_{\rm II},\mu_{0} scales respectively.

Normalized subsets for the region (1T)>0.33(1-T)>0.33 can be achieved simply by setting A𝐶𝑜𝑛𝑓0A_{\mathit{Conf}}\equiv 0 in the Eq. (5.14). Within the numerical precision of these calculations there is no evidence of the presence of spurious terms, such as any further UV-finite NFN_{F} term up to NNLO [194], besides the kinematic term at lowest order in the multi-jet region. These terms, if there are any, must remain rather small over the entire range of the thrust variable in comparison with the β\beta term or even be compatible with numerical fluctuations. Moreover, we notice a small rather constant difference between the iCF-predicted and the calculated coefficient for the Nf2N_{f}^{2} color factor of Ref. [173], which might be due to an nf2n_{f}^{2} UV-finite coefficient or possibly to statistics. This small difference must be included in the conformal coefficient but it has a completely negligible impact on the total thrust distribution.

Refer to caption
Figure 5.2: The thrust distribution under the PMC at NLO (dot-dashed blue) and at NNLO (solid red) [89]. The experimental data points are taken from the ALEPH, DELPHI, OPAL, L3, SLD experiments [177, 178, 179, 180, 181]. The shaded area shows theoretical errors for the PMC predictions at NLO and at NNLO.

In Fig. 5.2 we show the thrust distribution at NLO and at NNLO with the use of the PMC method. Theoretical errors for the thrust distribution at NLO and at NNLO are also shown (the shaded area). Conformal quantities are not affected by a change of renormalization scale. Thus, the errors shown give an evaluation of the level of conformality achieved up to the order of accuracy and they have been calculated using standard criteria, i.e. varying the remaining initial scale value in the range s/2μ02s\sqrt{s}/2\leq\mu_{0}\leq 2\sqrt{s}.

We recall that the distributions are calculated by a Monte-Carlo simulation, considering 50 equidistant bins in the range 0<1T<0.50<1-T<0.5, thus the peak in the NNLO-PMC distribution appears more as a broken-straight line also due to the linear interpolation used in the figure to join the points.

Using the same definition of the parameter δ¯\bar{\delta} given in Eq. (5.6), we have in the interval 0<(1T)<0.330<(1-T)<0.33 an average error of δ¯3.54%\bar{\delta}\simeq 3.54\% and 1.77%1.77\% for the thrust at NLO and at NNLO respectively. A greater improvement has been obtained over the entire range of reliable results for thrust distribution, i.e. 0<(1T)<0.420<(1-T)<0.42, from δ¯7.36%\bar{\delta}\simeq 7.36\% to 1.95%1.95\% from NLO to the NNLO accuracy with the PMC.

Refer to caption
Figure 5.3: The thrust distribution at NNLO under the Conventional (dashed black), the PMC(μ\muLO) (dot-dashed blue) and the PMC (solid red) [89]. The experimental data points are taken from the ALEPH, DELPHI, OPAL, L3, SLD experiments [177, 178, 179, 180, 181]. The shaded areas show theoretical error predictions at NNLO, which have been calculated varying the remaining initial scale value in the range s/2μ02s\sqrt{s}/2\leq\mu_{0}\leq 2\sqrt{s}.

In Fig. 5.3 a direct comparison of the PMC with the CSS results (obtained in [173] and [189, 190]) is shown. In addition, we show the results of the first PMC approach used in [87], which we indicate as PMC(μ\muLO) extended to NNLO accuracy. In this approach the last unknown PMC scale μ\muNLO of the NLO was set to the last known PMC scale μ\muLO of the LO, while the NNLO scale μ\muNNLOμ0\equiv\mu_{0} was left unset and varied in the range s/2μ02s\sqrt{s}/2\leq\mu_{0}\leq 2\sqrt{s}.

Average errors calculated in different regions of the spectrum are reported in Table 5.1. The PMC(μLO)(\mu_{LO}) cannot be defined in the range 0.33<1T<0.420.33<1-T<0.42 since AConf=0A_{\rm Conf}=0; thus the third and fifth row in Table 5.1 are blank. In fact, this further analysis was performed in order to show that the procedure of setting the last unknown scale to the last known one leads to stable and precise results and is consistent with the proper PMC method over a wide range of accessible values of the (1T)(1-T) variable.

δ¯[%]\bar{\delta}[\%] Conv. PMC(μ\muLO) PMC
0.10<(1T)<0.330.10<(1-T)<0.33 6.03 1.41 1.31
0.21<(1T)<0.330.21<(1-T)<0.33 6.97 2.19 0.98
0.33<(1T)<0.420.33<(1-T)<0.42 8.46 2.61
0.00<(1T)<0.330.00<(1-T)<0.33 5.34 1.33 1.77
0.00<(1T)<0.420.00<(1-T)<0.42 6.00 1.95
Table 5.1: The average error δ¯\bar{\delta} for the NNLO thrust distribution under Conventional, PMC(μ\muLO) and PMC scale settings calculated in different ranges of values of the (1T)(1-T) variable.

From the comparison with the CSS, we notice that the PMC prescription significantly improves the theoretical predictions. Moreover, results are in remarkable agreement with the experimental data over a wider range of values (0.0151T0.33)(0.015\leq 1-T\leq 0.33) and they show an improvement of the PMC(μLO)(\mu_{LO}) results when the two-jet and multi-jet regions are approached, i.e. the region of the peak and the region (1T)>0.33(1-T)>0.33 respectively. The use of the PMC approach in perturbative QCD thrust calculations restores the correct behavior of the thrust distribution in the region (1T)>0.33(1-T)>0.33 and this is a clear effect of the iCF property. Comparison with experimental data has been improved over the entire spectrum and the introduction of the N3LO\rm N^{3}LO correction would improve this comparison especially in the multi-jet (1T)>0.33(1-T)>0.33 region. In the PMC method theoretical errors are given by the unknown intrinsic conformal scale of the last order of accuracy. We expect this scale not to be significantly different from that of the previous orders. In this particular case, as shown in Eq. (5.14), we also have a dependence on the initial scale αs(μ0)\alpha_{s}(\mu_{0}) left due to the normalization and to the regularization terms. These errors represent 12.5% and 1.5% respectively of the full theoretical errors in the range 0<(1T)<0.420<(1-T)<0.42 and they could be improved by means of a correct normalization.

5.3   NNLO CC-parameter distribution results

The same analysis applies straightforwardly to the CC-parameter distribution including the regularizing η\eta parameter, which has been set to the same value 3.513.51. The same scales of Eq. (5.7) and Eq. (5.10) apply to the CC-parameter distribution in the region 0<C<0.750<C<0.75 and in the region 0.75<C<10.75<C<1. In fact, due to kinematic constraints that set the A𝐶𝑜𝑛𝑓=0A_{\mathit{Conf}}=0, we also have the same iCF effect for the CC-parameter.

Results for the CC-parameter scales are shown in Fig. 5.4. We notice the effect of the iCF intrinsic conformality on the LO-PMC scale, which terminates at the kinematic boundary C=0.75C=0.75. In fact, at this boundary AConf=0A_{\rm Conf}=0 and thus sets the whole conformal subset σI=0\sigma_{\rm I}=0. The NLO-PMC scale has two distinct domains separated by the kinematic constraint C=0.75C=0.75. In the range 0<C<0.750<C<0.75, the NLO-PMC scale has the same physical behavior as the LO-PMC scale, in the range 0.75<C<0.970.75<C<0.97 is almost constant, due to a similar behavior of the numerator and denominator in Eq.5.12 and at C=0.97C=0.97 it goes through a saturation effect given by the C1C_{1} coefficient in Eq.5.12, which becomes null. Results for the CC-parameter distributions are shown in Fig. 5.5.

Refer to caption
Figure 5.4: The LO-PMC (solid red) and the NLO-PMC (dashed black) scales for the CC-parameter [89].
Refer to caption
Figure 5.5: The CC-parameter distribution under the PMC at NLO (dot-dashed blue) and at NNLO (dashed red) [89]. The blue points indicate the NNLO-PMC thrust distribution obtained with μIII=μ0=MZ\mu_{\rm III}=\mu_{0}=M_{Z}. The experimental data points (green) are taken from the ALEPH experiment [177]. The dashed lines of the NNLO distribution show fits of the theoretical calculations with interpolating functions for the values of the remaining initial scale μ0=2MZ\mu_{0}=2M_{Z} and MZ/2M_{Z}/2. The shaded area shows theoretical errors for the PMC predictions at NLO and at NNLO calculated varying the remaining initial scale value in the range s/2μ02s\sqrt{s}/2\leq\mu_{0}\leq 2\sqrt{s}.

Theoretical errors have been calculated, as in the previous case, using standard criteria and results indicate an average error over the entire spectrum 0<C<10<C<1 of the CC-parameter distribution at NLO and at NNLO of δ¯7.26%\bar{\delta}\simeq 7.26\% and 2.43%2.43\% respectively.

δ¯\bar{\delta}\;[%] Conv. PMC(μ\muLO) PMC
0.00<(C)<0.750.00<(C)<0.75 4.77 0.85 2.43
0.75<(C)<1.000.75<(C)<1.00 11.51 3.68 2.42
0.00<(C)<1.000.00<(C)<1.00 6.47 1.55 2.43
Table 5.2: The average error δ¯\bar{\delta} for the NNLO CC-parameter distribution under Conventional, PMC(μ\muLO) and PMC scale settings calculated in different ranges of values of the (C)(C) variable.

A comparison of average errors according to the different methods is displayed in Table 5.2. Results show that the PMC improves the NNLO QCD predictions for the CC-parameter distribution over the entire spectrum.

Refer to caption
Figure 5.6: The NNLO CC-parameter distribution under CSS (dashed black), the PMC(μ\muLO) (dot-dashed blue) and the PMC (solid red) [89]. The experimental data points (Black) are taken from the ALEPH experiment [177]. The shaded area shows theoretical error predictions at NNLO calculated varying the remaining initial scale value in the range s/2μ02s\sqrt{s}/2\leq\mu_{0}\leq 2\sqrt{s}.

A comparison of the distributions calculated with the CSS, the PMC(μ\muLO) [88] and the PMC is shown in Fig. 5.6. The results for the PMC display remarkable agreement with the experimental data away from the regions C<0.05C<0.05 and C0.75C\simeq 0.75. The errors due to the normalization and to the regularization terms (Eq. (5.14)) are respectively 8.8%8.8\% and 0.7%0.7\% of the full theoretical errors.

The perturbative calculations could be further improved using a correct normalization and also by introducing the resummation of the large-logarithm technique in order to extend the perturbative regime and to eliminate the unphysical spike at C=0.75C=0.75, which is due to enhanced logarithmic terms at the kinematic boundary.

5.4   The thrust distribution in the QCD conformal window and in QED

For the first time, we employ the perturbative regime of the quantum chromodynamics (pQCD) infrared conformal window as a laboratory to investigate in a controllable manner (near) conformal properties of physically relevant quantities, such as the thrust distribution in electron–positron annihilation processes [90]. The conformal window of pQCD has a long and noble history conveniently summarized and generalized to arbitrary representations in Ref. [195]. Several lattice gauge theory applications and results have been summarized in a recent report on the subject in Ref. [196].

5.4.1   The thrust distribution according to NfN_{f}

It would be highly desirable to compare the PMC and CSS methods along the entire renormalization group flow from the highest energies down to zero energy. This is precluded in standard QCD with a number of active flavors less than six because the theory becomes strongly coupled at low energies. We therefore employ the perturbative regime of the conformal window (Sec. 2.2.2) which allows us to arrive at arbitrary low energies and obtain the corresponding results for the SU(33) case at the cost of increasing the number of active flavors. Here we are able to deduce the full solution at NNLO in the strong coupling. In this section we shall consider the region of flavors and colors near the upper bound of the conformal window, i.e. Nf11/2NcN_{f}\sim 11/2N_{c}, where the IR fixed point can be reliably accessed in perturbation theory and we compare the two renormalization scale setting methods, the CSS and the PMC.

Results for the thrust distribution calculated using the NNLO solution for the coupling αs(μ)\alpha_{s}(\mu), at different values of the number of flavors, NfN_{f}, is shown in Fig. 5.7.

Refer to caption
Figure 5.7: Thrust distributions for different values of NfN_{f}, using the PMC (solid line) and the CSS (dashed line) [90]. Shaded colored areas show error bars for each curve respectively. The experimental data points are taken from the ALEPH, DELPHI, OPAL, L3, SLD experiments [177, 178, 179, 180, 181].

A direct comparison between PMC (solid line) and CSS (dashed line) is shown at different values of the number of flavors. We notice that, despite the phase transition (i.e. the transition from an infrared finite coupling to an infrared divergent coupling), the curves given by the PMC at different NfN_{f}, preserve with continuity the same characteristics of the conformal distribution setting NfN_{f} outside the conformal window of pQCD.

Technically, this is explained by the fact that the PMC reabsorbs all the NfN_{f} terms into the running coupling and the PMC scales are both above 2 GeV in almost the entire range of the distribution; in particular, for values in the range 0.015<1T<0.420.015<1-T<0.42, and thus the PMC thrust distribution is affected by the change of behavior of the coupling only in the first two bins at 1T01-T\sim 0. However, this region is a multi-scale region and is not only affected by nonperturbative effects, but also by the presence of large-logarithms deriving from incomplete IR cancellation. On the other hand, the CSS distribution is more sensitive to the decreasing of NfN_{f}.

In fact, the position of the thrust distribution peak is well preserved varying NfN_{f} in and outside the conformal window using the PMC, while there is a constant shift towards lower values using the CSS. These trends are shown in Fig. 5.8. We notice that in the central range, 2<Nf<152<N_{f}<15, the position of the peak is exactly preserved using the PMC and overlaps with the position of the peak shown by the experimental data. According to our analysis for the case PMC, in the range Nf<2N_{f}<2, the number of bins is insufficient to resolve the peak, although the behavior of the curve is consistent with the presence of a peak in the same position, while for Nf0N_{f}\rightarrow 0, the peak is no longer visible. Theoretical uncertainties on the position of the peak have been calculated using standard criteria, i.e. by varying the remaining initial scale value in the range MZ/2μ02MZM_{Z}/2\leq\mu_{0}\leq 2M_{Z} and considering the lowest uncertainty given by the half the spacing between two adjacent bins.

Refer to caption
Figure 5.8: Comparison of the position of the peak for the thrust distribution using the CSS and the PMC vs the number of flavors, NfN_{f}. Dashed lines indicate the particular trend in each graph [90].

Using the definition given in Eq. (5.6), we have determined the average error, δ¯\bar{\delta}, calculated in the interval 0.005<(1T)<0.40.005<(1-T)<0.4 of thrust and results for CSS and PMC are shown in Fig. 5.9. We notice that the PMC in the perturbative and IR conformal window, i.e. 12<Nf<N¯f12<N_{f}<\bar{N}_{f}, which is the region where αs(μ)<1\alpha_{s}(\mu)<1 in the entire range of the renormalization scale values, from 0 up to \infty, the average error given by PMC tends to zero (0.230.26%\sim 0.23-0.26\%) while the error given by the CSS tends to remain constant (0.850.89%0.85-0.89\%). Comparison of the two methods shows that, outside the conformal window, Nf<34Nc313Nc23N_{f}<\frac{34N_{c}^{3}}{13N_{c}^{2}-3}, the PMC leads to higher precision.

Refer to caption
Figure 5.9: Comparison of the average theoretical error, δ¯\bar{\delta}, calculated using standard criteria in the range: 0.005<(1T)<0.40.005<(1-T)<0.4, using the CSS and the PMC for the thrust distribution vs. the number of flavors, NfN_{f} [90].

From our analysis in section 2.2.2, it follows that the IR fixed point at Nf=112NcN_{f}=\frac{11}{2}N_{c} is not phenomenologically accessible, assuming a measured value of the coupling at a certain scale, since in the region N¯f<Nf<112Nc\bar{N}_{f}<N_{f}<\frac{11}{2}N_{c}, the coupling would no longer have the UV asymptotically free behavior. Thus, in order to have a phenomenological application of the Banks-Zaks variable, (i.e. Δf=33/2Nfϵ\Delta_{f}=33/2-N_{f}\simeq\epsilon, as shown in Refs.[197, 198]), one should be able to reach any small value of the coupling above the Planck scale in QCD, since the IR fixed point is interacting. Moreover, a conformal result is just characterized by the absence of running of the coupling and not necessarily by a null value. A straight conformal solution of Eqs. 2.31 and 2.32, is given by: z=0z=0, W=0W=0 with x(μ)=x=x0x(\mu)=x^{\ast}=x_{0} for any scale 0<μ<0<\mu<\infty. Our application of the Banks-Zaks results shows that the conformal limit of the thrust distribution is less sensitive to the method adopted to set the renormalization scale, CSS or PMC, once the same phenomenological value of the coupling has been determined at a certain scale. We suggest that a more suitable variable for the expansion would be given by Δ~=N¯fNfϵ\tilde{\Delta}=\bar{N}_{f}-N_{f}\simeq\epsilon, where N¯f\bar{N}_{f} is the maximum allowed value of NfN_{f} to be asymptotically free, introduced by the phenomenological value of the coupling at an initial scale μ0\mu_{0}.

5.4.2   The thrust distribution in the Abelian limit Nc0N_{c}\rightarrow 0

We consider now the thrust distribution in U(1) Abelian QED, which rather than being infrared interacting is infrared free. We obtain the QED thrust distribution performing the Nc0N_{c}\rightarrow 0 limit of the QCD thrust at NNLO according to [32, 145]. In the zero number of colors limit the gauge group color factors are fixed by NA=1,N_{A}=1, CF=1,C_{F}=1, TR=1,T_{R}=1, CA=0,C_{A}=0, Nc=0,N_{c}=0, Nf=NlN_{f}=N_{l}, where NlN_{l} is the number of active leptons, while the β\beta-terms and the coupling rescale as βn/CFn+1\beta_{n}/C_{F}^{n+1} and αsCF\alpha_{s}\cdot C_{F} respectively. In particular, β0=43Nl\beta_{0}=-\frac{4}{3}N_{l} and β1=4Nl\beta_{1}=-4N_{l} using the normalization of Eq. (2.15). According to this rescaling of the color factors, we have determined the QED thrust and the QED PMC scales. For the QED coupling, we have used the analytic formula for the effective fine structure constant in the MS¯\overline{\textrm{MS}} scheme:

α(Q)=α(1eΠMS¯(Q2)),{\alpha(Q)}={\alpha\over{\left(1-\Re e\Pi^{\overline{\textrm{MS}}}(Q^{2})\right)}}, (5.15)

with α1α(0)1=137.036\alpha^{-1}\equiv\alpha(0)^{-1}=137.036 and the vacuum polarization function (Π\Pi) calculated perturbatively to two loops, including contributions from leptons, quarks and the WW boson. The QED PMC scales have the same form of Eqs. (5.7) and (5.10) with the factor for the MS¯\overline{\textrm{MS}} scheme set to fsc5/6f_{sc}\equiv 5/6 and the η\eta regularization parameter introduced to cancel singularities in the NLO PMC scale μII\mu_{\rm II} in the Nc0N_{c}\rightarrow 0 limit tends to the same QCD value, η=3.51\eta=3.51. A direct comparison between QED and QCD PMC scales is shown in Fig. 5.10.

Refer to caption
Figure 5.10: PMC scales for the thrust distribution: LO-QCD scale (solid red); LO-QED scale (solid blue);NLO-QCD scale (dashed red); NLO-QED scale (dashed blue) [90].

We note that in the QED limit the PMC scales have analogous dynamical behavior to those calculated in QCD; differences arise mainly owing to the MS¯\overline{\textrm{MS}} scheme factor reabsorption, the effects of the NcN_{c} number of colors at NLO are very small. Thus, we notice that perfect consistency is shown from QCD to QED using the PMC method. The normalized QED thrust distribution is shown in Fig. 5.11. We note that the curve is peaked at the origin, T=1T=1, which suggests that the three-jet event in QED occurs with a rather back-to-back symmetry. Results for the CSS and the PMC methods in QED are O(α)O(\alpha) and show very small differences, given the good convergence of the theory.

Refer to caption
Figure 5.11: Thrust distributions in the QED limit at NNLO using the PMC (solid red) and the CSS (dashed blue) [90].

5.5   A novel method for the precise determination of the strong coupling and its behavior

In this section we present a novel method for precisely determining the running QCD coupling constant αs(Q)\alpha_{s}(Q) over a wide range of QQ from event-shape variables for electron–positron annihilation measured at a single center-of-mass energy s\sqrt{s}, based on PMC scale setting. In particular, we display the results obtained in Refs. [88, 87] using the approach of a single PMC scale at LO and NLO, i.e. the PMC(μ\muLO) of the previous section.

The precise determination of the strong coupling αs(Q)\alpha_{s}(Q) is one of the crucial tests of QCD. The dependence of αs(Q)\alpha_{s}(Q) on the renormalization scale QQ obtained from many different physical processes shows consistency with QCD predictions and asymptotic freedom. The Particle Data Group (PDG) currently gives the world average: αs(MZ)=0.1179±0.0009\alpha_{s}(M_{Z})=0.1179\pm 0.0009 [126] in the MS¯\overline{\rm{MS}} renormalization scheme.

Particularly suitable to the determination of the strong coupling is the process e+e3e^{+}e^{-}\rightarrow 3\,jets since its leading order is 𝒪(αs)\mathcal{O}(\alpha_{s}) [176]. Currently, theoretical calculations for event shapes are based on CSS. By using conventional scale setting, only one value of αs\alpha_{s} at the scale s\sqrt{s} can be extracted and the main source of the uncertainty is given by the choice of the renormalization scale. Several values for the strong coupling have been extracted from several processes, e.g. αs(MZ)=0.1224±0.0035\alpha_{s}(M_{Z})=0.1224\pm 0.0035 [199] is obtained by using perturbative corrections and resummation of the large logarithms in the NNLO+NLL accuracy predictions. Other evaluations improving the resummation calculations up to N3LL give a result of αs(MZ)=0.1135±0.0011\alpha_{s}(M_{Z})=0.1135\pm 0.0011 [31] from thrust and αs(MZ)=0.1123±0.0015\alpha_{s}(M_{Z})=0.1123\pm 0.0015 [200] from the CC-parameter. Nonperturbative corrections for hadronization effects have also been included in Ref. [201], but as pointed out in Ref. [202], the systematics of the theoretical uncertainties introduced by hadronization effects are not well understood.

In this section we show that by using the PMC, it is possible to eliminate the renormalization scale ambiguities and obtain consistent results for the strong coupling using the precise experimental data of event-shape variable distributions. We notice that improved event-shape distributions have been obtained in Refs. [44, 194, 203] using BLM and soft and collinear effective theory (SCET).

5.5.1   Running behavior

Given that the PMC scale (PMC(μ\muLO)) is not a single-valued but rather a monotonically increasing function of s\sqrt{s} and of the selected observable,(as shown in Figs. 5.1 and 5.4), it is possible to determine the strong coupling at different scales from one single experiment at one single center-of-mass energy. The dependence of the scale on the observable reflects the dynamics of the underlying gluon and quark subprocess. This dynamics also varies the number of active flavors NfN_{f}. Considering that PMC scales for QCD and QED show the same behavior and that their relation at LO is only given by a RS redefinition term, QQCD2/QQED2=e5/3Q^{2}_{\rm QCD}/Q^{2}_{\rm QED}=e^{-5/3}, this approach may also be extended to QED.

Refer to caption
Figure 5.12: The coupling constant αs(Q)\alpha_{s}(Q) extracted by comparing PMC predictions with the ALEPH data [177] at a single energy of s=MZ\sqrt{s}=M_{Z} from the CC-parameter distributions in the MS¯\overline{\rm MS} scheme (from Ref. [88]). The error bars are the squared averages of the experimental and theoretical errors. The three lines are the world average evaluated from αs(MZ)=0.1179±0.009\alpha_{s}(M_{Z})=0.1179\pm 0.009 [126].

We extract αs\alpha_{s} at different scales bin-by-bin from the comparison of PMC predictions for (1T1-T) and CC differential distributions with measurements at s=MZ\sqrt{s}=M_{Z}. The extracted αs\alpha_{s} values from the CC-parameter distribution are shown in Fig. 5.12. We note that the αs\alpha_{s} values extracted in the scale range of 33 GeV<Q<11\,<Q<11 GeV are in excellent agreement with those evaluated from the world average αs(MZ)\alpha_{s}(M_{Z}) [126]. Given that the PMC scale setting eliminates the scale uncertainties, the corresponding extracted αs\alpha_{s} values are not plagued by ambiguities in the choice of μr\mu_{r}. The extracted αs\alpha_{s} values from the thrust observable using PMC are shown in Fig. 5.13. There is good agreement also in this case in the range 3.53.5 GeV <μr<16<\mu_{r}<16 GeV (corresponding to the range 0.05<(1T)<0.290.05<(1-T)<0.29). These extracted values of αs\alpha_{s} are also in good agreement with the world average value of the PDG [126].

Refer to caption
Figure 5.13: The extracted αs\alpha_{s} from the comparison of PMC predictions with ALEPH data at s=MZ\sqrt{s}=M_{Z} (from Ref. [87]). The error bars are from the experimental data. The three lines are the world average evaluated from αs(MZ)=0.1179±0.0009\alpha_{s}(M_{Z})=0.1179\pm 0.0009 [126].

Thus, PMC scale setting provides a remarkable way to verify the running of αs(Q)\alpha_{s}(Q) from event shapes measured at a single energy of s\sqrt{s}. Analogously in QED, the running of the QED coupling α(Q)\alpha(Q) can be measured at a single energy of s\sqrt{s} (see e.g. [204]).

The differential distributions of event shapes are afflicted with large logarithms especially in the two-jet region. Thus, the comparison of QCD predictions with experimental data and thus the extracted αs\alpha_{s} values are restricted to the region where the theory is able to describe the data well. Choosing a different area of distributions leads to the different values of αs\alpha_{s}.

5.5.2   αs(MZ)\alpha_{s}(M_{Z}) from a χ2\chi^{2} fit

In order to obtain a reliable αs\alpha_{s} at the scale of the Z0Z^{0} mass, we determine αs(MZ)\alpha_{s}(M_{Z}) from the fit of the PMC predictions to measurements. In particular, we perform the fit by minimizing the χ2\chi^{2} respect to the αs(MZ)\alpha_{s}(M_{Z}) parameter. The variable χ2\chi^{2} is defined as:

χ2=i(yiexptyithσi)2,\chi^{2}=\sum_{i}\left(\frac{\langle y\rangle^{\rm{expt}}_{i}-\langle y\rangle^{\rm{th}}_{i}}{\sigma_{i}}\right)^{2},

where yiexpt\langle y\rangle^{\rm{expt}}_{i} is the value of the experimental data, σi\sigma_{i} is the corresponding experimental uncertainty and yith\langle y\rangle^{\rm{th}}_{i} is the theoretical prediction. The fit for thrust and the CC-parameter leads to the following results:

αs(MZ)\displaystyle\alpha_{s}(M_{Z}) =\displaystyle= 0.1185±0.0011(expt)±0.0005(th)\displaystyle 0.1185\pm 0.0011(\rm expt)\pm 0.0005(\rm th) (5.16)
=\displaystyle= 0.1185±0.0012,\displaystyle 0.1185\pm 0.0012,

with χ2/\chi^{2}/d.o.f. =27.3/20=27.3/20 for the thrust mean value and

αs(MZ)\displaystyle\alpha_{s}(M_{Z}) =\displaystyle= 0.11930.0010+0.0009(expt)0.0016+0.0019(th)\displaystyle 0.1193^{+0.0009}_{-0.0010}(\rm expt)\,^{+0.0019}_{-0.0016}(\rm th) (5.17)
=\displaystyle= 0.11930.0019+0.0021,\displaystyle 0.1193^{+0.0021}_{-0.0019},

with χ2/\chi^{2}/d.o.f. =43.9/20=43.9/20 for the CC-parameter mean value, where the first error is the experimental uncertainty and the second the theoretical uncertainty. Both results are consistent with the world average αs(MZ)=0.1179±0.0009\alpha_{s}(M_{Z})=0.1179\pm 0.0009 [126].

The precision of the extracted αs\alpha_{s} has been greatly improved by using the PMC: the dominant μr\mu_{r} scale uncertainties are eliminated and the convergence of pQCD series is greatly improved. In particular, a strikingly much faster pQCD convergence is obtained for the mean thrust value [87], theoretical uncertainties are even smaller than the experimental uncertainties. We remark that these results for αs(MZ)\alpha_{s}(M_{Z}) are among the most precise determinations of the strong coupling at the Z0Z^{0} mass from event-shape variables.

6   Comparison of the PMCm, PMCs and PMC\infty for fully integrated fundamental quantities

In this section, we show predictions for three important quantities: Re+eR_{e^{+}e^{-}}, RτR_{\tau} and Γ(Hbb¯)\Gamma(H\to b\bar{b}), which have been calculated up to four-loop QCD corrections using alternative PMC scale-setting procedures. Numerical results for the conventional, PMCm, PMCs and PMC approaches will be presented. For self-consistency, the same loop αs\alpha_{s}-running behavior will be adopted for calculating the same loop perturbative series. The QCD asymptotic scale (ΛQCD\Lambda_{\rm QCD}) is then fixed by using αs(MZ)=0.1179\alpha_{s}(M_{Z})=0.1179 [126] and for a four-loop prediction, we obtain ΛQCDnf=4=291.7\Lambda_{\rm{QCD}}^{n_{f}=4}=291.7 MeV and ΛQCDnf=5=207.2\Lambda_{\rm{QCD}}^{n_{f}=5}=207.2 MeV in a conventional MS¯\overline{\rm MS} renormalization scheme.

6.1   The pQCD predictions for Re+eR_{e^{+}e^{-}}, RτR_{\tau} and Γ(Hbb¯)\Gamma(H\to b\bar{b})

The annihilation of electrons and positrons into hadrons provides one of the most important platforms for determining the running behavior of the QCD coupling. The RR ratio is defined as

Re+e(Q)\displaystyle R_{e^{+}e^{-}}(Q) =\displaystyle= σ(e+ehadrons)σ(e+eμ+μ)\displaystyle\frac{\sigma\left(e^{+}e^{-}\rightarrow{\rm hadrons}\right)}{\sigma\left(e^{+}e^{-}\rightarrow\mu^{+}\mu^{-}\right)} (6.1)
=\displaystyle= 3qeq2[1+R(Q)],\displaystyle 3\sum_{q}e_{q}^{2}\left[1+R(Q)\right],

where Q=sQ=\sqrt{s}, corresponding to the electron–positron collision energy in the center-of-mass frame. The pQCD series for R(Q)R(Q), up to (n+1)(n{+}1)-loop QCD corrections, is

Rn(Q)=i=0n𝒞i(Q,μr)asi+1(μr).R_{n}(Q)=\sum_{i=0}^{n}{\cal C}_{i}(Q,\mu_{r})a_{s}^{i+1}(\mu_{r}).

The perturbative coefficients 𝒞i(Q,μr){\cal C}_{i}(Q,\mu_{r}) in the MS¯\overline{\rm MS} scheme up to four-loop level have been calculated in Refs. [205, 162, 206, 207]. As a reference point, at s=31.6\sqrt{s}=31.6 GeV, we have 311Re+eexpt=1.0527±0.0050\frac{3}{11}R_{e^{+}e^{-}}^{\rm expt}=1.0527\pm 0.0050 [208].

Another useful ratio is RτR_{\tau} for τ\tau-lepton decays into hadrons, defined as

Rτ(Mτ)\displaystyle R_{\tau}(M_{\tau}) =\displaystyle= Γ(τντ+hadrons)Γ(τντ+ν¯l+l)\displaystyle\frac{\Gamma\left(\tau\rightarrow\nu_{\tau}+{\rm hadrons}\right)}{\Gamma\left(\tau\rightarrow\nu_{\tau}+\bar{\nu}_{l}+l\right)} (6.2)
=\displaystyle= 3|Vud|2SEW[1+R^(Mτ)+δEW+δ2+δNP],\displaystyle 3|V_{ud}|^{2}S_{\rm EW}\left[1+\hat{R}(M_{\tau})+\delta_{\rm EW}^{{}^{\prime}}+\delta_{2}+\delta_{\rm NP}\right],

where Vud=0.97373±0.00031V_{ud}=0.97373\pm 0.00031 [126] is the relevant Cabibbo–Kobayashi–Maskawa matrix element, SEW=1.0198±0.0006S_{\rm EW}=1.0198\pm 0.0006 and δEW=0.001\delta_{\rm EW}^{{}^{\prime}}=0.001 for the electroweak corrections, δ2=(4.4±2.0)×104\delta_{2}=(-4.4\pm 2.0)\times 10^{-4} for light-quark mass effects666An improved determination of the msm_{s} and |Vus||V_{\rm us}| from the τ\tau-decay has been recently shown in Ref. [209]., δNP=(4.8±1.7)×103\delta_{\rm NP}=(-4.8\pm 1.7)\times 10^{-3} for the nonperturbative effects and Mτ=1.777M_{\tau}=1.777 GeV [210, 211, 205]. The pQCD series for R^(Mτ)\hat{R}(M_{\tau}) up to (n+1)(n{+}1)-loop QCD corrections is

R^n(Mτ)=i=0n𝒞^i(Mτ,μr)asi+1(μr),\hat{R}_{n}(M_{\tau})=\sum_{i=0}^{n}\hat{\cal C}_{i}(M_{\tau},\mu_{r})a_{s}^{i+1}(\mu_{r}),

where the perturbative coefficients 𝒞^i(Mτ,μr)\hat{\cal C}_{i}(M_{\tau},\mu_{r}) up to four-loop QCD corrections can be derived by using the relation between Rτ(Mτ)R_{\tau}(M_{\tau}) and Re+e(s)R_{e^{+}e^{-}}(\sqrt{s}) [212].

The decay width for Higgs boson decay into a bottom and anti-bottom pair, Hbb¯H\to b\bar{b}, can be written as

Γ(Hbb¯)=3GFMHmb2(MH)42π[1+R~(MH)],\displaystyle\Gamma(H\to b\bar{b})=\frac{3G_{F}M_{H}m_{b}^{2}(M_{H})}{4\sqrt{2}\pi}[1+\tilde{R}(M_{H})], (6.3)

where the Fermi constant GF=1.16638×105GeV2G_{F}=1.16638\times 10^{-5}\,\rm{GeV}^{-2}, the Higgs mass MH=125.1M_{H}=125.1 GeV and the bb-quark MS¯\overline{\rm{MS}} running mass is mb(MH)=2.78m_{b}(M_{H})=2.78 GeV [69]. The pQCD series for R~(MH)\tilde{R}(M_{H}) up to (n+1)(n{+}1)-loop QCD corrections is

R~n(MH)=i=0n𝒞~i(MH,μr)asi+1(μr).\tilde{R}_{n}(M_{H})=\sum_{i=0}^{n}\tilde{\cal C}_{i}(M_{H},\mu_{r})a_{s}^{i+1}(\mu_{r}).

The perturbative coefficients 𝒞~i(MH,μr)\tilde{\cal C}_{i}(M_{H},\mu_{r}) up to four-loop QCD corrections have been calculated in Ref. [213]. In the following, we give the properties for the pQCD series of Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}) using each scale-setting approach. As for the leading-order ratios with n=0n=0, we have no information to set the renormalization scale for all the scale-setting approaches; and for convenience, we directly set it to be QQ, MτM_{\tau}, or MHM_{H}, respectively, which gives R0=0.04428R_{0}=0.04428, R^0=0.0891\hat{R}_{0}=0.0891 and R~0=0.2034\tilde{R}_{0}=0.2034. We point out that all four-loop calculations for these observables derive from analytic properties (of the Adler function) rather than from a direct multi-loop calculation of the perturbative series.

6.2   Properties using the conventional scale-setting approach

   𝒞1(μr){\cal C}_{1}(\mu_{r})    𝒞2(μr){\cal C}_{2}(\mu_{r})    𝒞3(μr){\cal C}_{3}(\mu_{r})    𝒞4(μr){\cal C}_{4}(\mu_{r})
R(Q)R(Q) 44 22.5542.51+42.5122.55^{+42.51}_{-42.51} 819.50266.26+1145.54-819.50^{+1145.54}_{-266.26} 205917812.4+31244.8-20591^{+31244.8}_{-7812.4}
R^(Mτ)\hat{R}(M_{\tau}) 44 83.2441.39+49.9183.24^{+49.91}_{-41.39} 1687.421588.77+3054.601687.42^{+3054.60}_{-1588.77} 32532.137674.5+139210.032532.1^{+139210.0}_{-37674.5}
R~(MH)\tilde{R}(M_{H}) 22.666722.6667 466.347240.907+240.907466.347^{+240.907}_{-240.907} 2672.498567.51+13688.402672.49^{+13688.40}_{-8567.51} 21139127651.9+358424-211391^{+358424}_{-27651.9}
Table 6.1: The scale-dependent coefficients 𝒞i(μr){\cal C}_{i}(\mu_{r}) of the conventional series for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}), respectively. The central values are for μr=Q\mu_{r}=Q, MτM_{\tau}, or MHM_{H}, respectively. The errors are evaluated by taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] for R^n(Mτ)\hat{R}_{n}(M_{\tau}) and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] for R~n(MH)\tilde{R}_{n}(M_{H}), respectively.

We present the perturbative coefficients 𝒞i(μr){\cal C}_{i}(\mu_{r}) in Table 6.1, where the errors are evaluated by taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] for R(Q=31.6R(Q{=}31.6 GeV), μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] for R^(Mτ)\hat{R}(M_{\tau}) and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] for R~(MH)\tilde{R}(M_{H}), respectively. Table 6.1 shows that those coefficients are highly scale dependent.

n=1n=1 n=2n=2 n=3n=3 κ1\kappa_{1} κ2\kappa_{2} κ3\kappa_{3}
Rn|Conv.R_{n}|_{\rm Conv.} 0.047530.00138+0.000440.04753^{+0.00044}_{-0.00138} 0.046380.00070+0.000120.04638^{+0.00012}_{-0.00070} 0.046080.00009+0.000150.04608^{+0.00015}_{-0.00009} 7.3+9.22.9%7.3^{-2.9}_{+9.2}\% 2.41.4+0.7%\hphantom{0}2.4^{+0.7}_{-1.4}\% 0.6+0.10.0%0.6^{-0.0}_{+0.1}\%
R^n|Conv.\hat{R}_{n}|_{\rm Conv.} 0.15220.0295+0.04820.1522^{+0.0482}_{-0.0295} 0.18260.0268+0.03600.1826^{+0.0360}_{-0.0268} 0.19800.0194+0.01700.1980^{+0.0170}_{-0.0194} 70.8+2.3+2.6%70.8^{+2.6}_{+2.3}\% 20.0+7.010.9%20.0^{-10.9}_{+7.0}\% 8.4+6.26.8%8.4^{-6.8}_{+6.2}\%
R~n|Conv.\tilde{R}_{n}|_{\rm Conv.} 0.24040.0075+0.00740.2404^{+0.0074}_{-0.0075} 0.24230.0007+0.00020.2423^{+0.0002}_{-0.0007} 0.24090.0007+0.00150.2409^{+0.0015}_{-0.0007} 18.2+6.78.0%18.2^{-8.0}_{+6.7}\% 0.8+2.9+1.3%\hphantom{0}0.8^{+1.3}_{+2.9}\% 0.6+0.00.6%0.6^{-0.6}_{+0.0}\%
Table 6.2: Results for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}), R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the conventional scale-setting approach. The central values are obtained by setting μr\mu_{r} as QQ, MτM_{\tau}, or MHM_{H}, respectively. The errors are evaluated by taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] for R^n(Mτ)\hat{R}_{n}(M_{\tau}) and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] for R~n(MH)\tilde{R}_{n}(M_{H}).

We present the results of Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the conventional scale-setting approach in Table 6.2, where the errors are evaluated by taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] for R3(Q=31.6R_{3}(Q{=}31.6 GeV), μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] for R^3(Mτ)\hat{R}_{3}(M_{\tau}) and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] for R~3(MH)\tilde{R}_{3}(M_{H}), respectively. For self-consistency, we adopt the (n+1)(n{+}1)th-loop αs\alpha_{s}-running behavior in deriving (n+1)(n{+}1)th-loop prediction for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}). We define the ratio

κn=|nn1n1|,\kappa_{n}=\left|\frac{{\cal R}_{n}-{\cal R}_{n-1}}{{\cal R}_{n-1}}\right|,

where {\cal R} stands for RR, R^\hat{R} and R~\tilde{R}, respectively. It shows how the “known” prediction n1{\cal R}_{n-1} is affected by the one-order-higher terms. Table 6.2 shows that generally we have κ1>κ2>κ3\kappa_{1}>\kappa_{2}>\kappa_{3} for all those quantities, consistently with the perturbative nature of the series and indicates that one can obtain more precise predictions by including more loop terms. To show the perturbative nature more explicitly, we present the magnitudes of each loop term for the four-loop approximants R3(Q=31.6GeV)R_{3}(Q=31.6\,\mathrm{GeV}), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) in Table 6.3, which displays the relative importance among the LO, NLO, N2LO and N3LO terms, which for those approximants are

1:+0.0630.100+0.099:0.0260.016+0.034:0.007+0.009+0.001,\displaystyle 1:+0.063^{+0.099}_{-0.100}:-0.026^{+0.034}_{-0.016}:-0.007^{+0.001}_{+0.009}, (6.4)
1:+0.531+0.1010.108:+0.275+0.1520.205:+0.135+0.1590.186,\displaystyle 1:+0.531^{-0.108}_{+0.101}:+0.275^{-0.205}_{+0.152}:+0.135^{-0.186}_{+0.159}, (6.5)
1:+0.1840.085+0.071:+0.0090.035+0.039:0.0070.002+0.010,\displaystyle 1:+0.184^{+0.071}_{-0.085}:+0.009^{+0.039}_{-0.035}:-0.007^{+0.010}_{-0.002}, (6.6)

where the central values are for μr=Q\mu_{r}=Q, μr=Mτ\mu_{r}=M_{\tau} and μr=MH\mu_{r}=M_{H}; and the errors are for μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q], μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}], respectively. Consistently with Table 6.2, the scale dependence for each loop term is large, but due to the cancellation of scale dependence among different orders, the net scale dependence is small, e.g. ()0.2%+0.3%\left({}^{+0.3\%}_{-0.2\%}\right), ()9.8%+8.6%\left({}^{+8.6\%}_{-9.8\%}\right) and ()0.3%+0.6%\left({}^{+0.6\%}_{-0.3\%}\right) for R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}), respectively. Note that due to the usual renormalon divergence and a larger αs\alpha_{s} value at a smaller scale MτM_{\tau}, i.e. αs(Mτ)0.33\alpha_{s}(M_{\tau})\sim 0.33, the net scale dependence of the four-loop prediction R^3(Mτ)\hat{R}_{3}(M_{\tau}) is still sizable.

LO\rm LO NLO\rm NLO N2LO\rm N^{2}LO N3LO\rm N^{3}LO Total\rm Total
R3|Conv.R_{3}|_{\rm Conv.} 0.04473+0.005120.004990.04473^{-0.00499}_{+0.00512} 0.002820.00468+0.003600.00282^{+0.00360}_{-0.00468} 0.001150.00095+0.00147-0.00115^{+0.00147}_{-0.00095} 0.00032+0.00042+0.00007-0.00032^{+0.00007}_{+0.00042} 0.046080.00009+0.000150.04608^{+0.00015}_{-0.00009}
R^3|Conv.\hat{R}_{3}|_{\rm Conv.} 0.10200.0261+0.04710.1020^{+0.0471}_{-0.0261} 0.05420.0062+0.00890.0542^{+0.0089}_{-0.0062} 0.0280+0.00440.0176\hphantom{-}0.0280^{-0.0176}_{+0.0044} 0.0138+0.00850.0214\hphantom{-}0.0138^{-0.0214}_{+0.0085} 0.19800.0194+0.01700.1980^{+0.0170}_{-0.0194}
R~3|Conv.\tilde{R}_{3}|_{\rm Conv.} 0.2030+0.02260.01750.2030^{-0.0175}_{+0.0226} 0.03740.0151+0.01000.0374^{+0.0100}_{-0.0151} 0.00190.0077+0.0070\hphantom{-}0.0019^{+0.0070}_{-0.0077} 0.00140.0005+0.0020-0.0014^{+0.0020}_{-0.0005} 0.24090.0007+0.00150.2409^{+0.0015}_{-0.0007}
Table 6.3: The value of each loop term (LO, NLO, N2LO and N3LO) for the four-loop QCD predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the conventional scale-setting approach. The errors are evaluated by taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] for R3(Q=31.6R_{3}(Q{=}31.6 GeV), μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] for R^3(Mτ)\hat{R}_{3}(M_{\tau}) and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] for R~3(MH)\tilde{R}_{3}(M_{H}).

In the pQCD calculation, it is helpful to give a reliable prediction of the uncalculated higher-order terms. The Padé approximant approach (PAA) [130, 131, 132], shown in Eq. (2.38), provides an effective method to estimate the (n+1)(n{+}1)th-order coefficient from a given nn-th-order series.777Another method, which uses the scale-invariant conformal series together with the Bayesian model [214, 215, 216] to provide probabilistic estimates of the unknown higher-orders terms has been proposed [217] In practice, it has been found that the PAA becomes more effective when more loop terms are known. For a pQCD approximant, ρ(Q)=c1as+c2as2+c3as3+c4as4+\rho(Q)=c_{1}a_{s}+c_{2}a_{s}^{2}+c_{3}a_{s}^{3}+c_{4}a_{s}^{4}+\cdots, the predicted N3LO and the N4LO terms are

ρ[1/1]N3LO\displaystyle\rho^{\rm N^{3}LO}_{[1/1]} =\displaystyle= c32c2as4,\displaystyle\frac{c_{3}^{2}}{c_{2}}a_{s}^{4}, (6.7)
ρ[0/2]N3LO\displaystyle\rho^{\rm N^{3}LO}_{[0/2]} =\displaystyle= 2c1c2c3c23c12as4,\displaystyle\frac{2c_{1}c_{2}c_{3}-c_{2}^{3}}{c_{1}^{2}}a_{s}^{4}, (6.8)
ρ[1/2]N4LO\displaystyle\rho^{\rm N^{4}LO}_{[1/2]} =\displaystyle= 2c2c3c4c33c1c42c22c1c3as5,\displaystyle\frac{2c_{2}c_{3}c_{4}-c_{3}^{3}-c_{1}c_{4}^{2}}{c_{2}^{2}-c_{1}c_{3}}a_{s}^{5}, (6.9)
ρ[2/1]N4LO\displaystyle\rho^{\rm N^{4}LO}_{[2/1]} =\displaystyle= c42c3as5,\displaystyle\frac{c_{4}^{2}}{c_{3}}a_{s}^{5}, (6.10)
ρ[0/3]N4LO\displaystyle\rho^{\rm N^{4}LO}_{[0/3]} =\displaystyle= c243c1c22c3+2c12c2c4+c12c32c13as5.\displaystyle\frac{c_{2}^{4}-3c_{1}c_{2}^{2}c_{3}+2c_{1}^{2}c_{2}c_{4}+c_{1}^{2}c_{3}^{2}}{c_{1}^{3}}a_{s}^{5}. (6.11)
N3LO\rm N^{3}LO N4LO\rm N^{4}LO
R3|Conv.R_{3}|_{\rm Conv.} [1/1][1/1]: 0.000470.00284+0.000000.00047^{+0.00000}_{-0.00284} [1/2][1/2]: 0.000020.00010+0.00011-0.00002^{+0.00011}_{-0.00010}
[2/1][2/1]: 0.000090.00000+0.00028-0.00009^{+0.00028}_{-0.00000}
R^3|Conv.\hat{R}_{3}|_{\rm Conv.} [1/1][1/1]: 0.01450.0128+0.00740.0145^{+0.0074}_{-0.0128} [1/2][1/2]: +0.00610.0142+0.0093+0.0061^{+0.0093}_{-0.0142}
[2/1][2/1]: +0.00680.0011+0.0085+0.0068^{+0.0085}_{-0.0011}
R~3|Conv.\tilde{R}_{3}|_{\rm Conv.} [1/1][1/1]: 0.00010.0000+0.00160.0001^{+0.0016}_{-0.0000} [1/2][1/2]: 0.00060.0000+0.0005-0.0006^{+0.0005}_{-0.0000}
[2/1][2/1]: +0.00100.0016+0.0000+0.0010^{+0.0000}_{-0.0016}
Table 6.4: The preferable diagonal-type PAA predictions of the N3LO\rm N^{3}LO and N4LO\rm N^{4}LO terms of R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the conventional scale-setting approach. The uncertainties correspond to taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q], μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}], respectively.

Table 6.4 displays the preferable diagonal-type PAA predictions [133] of the N3LO\rm N^{3}LO and N4LO\rm N^{4}LO terms of R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the conventional scale-setting approach. Owing to the large scale dependence of each loop term, the PAA predictions show large scale dependence. The two allowable diagonal-type PAA predictions for N4LO\rm N^{4}LO terms are consistent with each order within errors. Comparing Table 6.4 with Table 6.3, we note that the values of the predicted N3LO\rm N^{3}LO terms agree with their exact values within errors. Thus, by employing a perturbative series with enough higher-order terms, the PAA prediction can be made reliable.

6.3   Properties using the PMCm approach

Following the standard PMCm procedures, the nonconformal {βi}\{\beta_{i}\}-terms are eliminated by using the RGE recursively, determining the effective αs\alpha_{s} at each perturbative order and resulting in the renormalon-free and scheme-independent conformal series (3.14).

r^1,0{\hat{r}}_{1,0} r^2,0{\hat{r}}_{2,0} r^3,0{\hat{r}}_{3,0} r^4,0{\hat{r}}_{4,0}
R(Q)R(Q) 44 29.4429.44 64.25-64.25 2812.74-2812.74
R^(Mτ)\hat{R}(M_{\tau}) 44 34.3334.33 219.89219.89 1741.151741.15
R~(MH)\tilde{R}(M_{H}) 22.666722.6667 216.356216.356 8708.09-8708.09 110597-110597
Table 6.5: Conformal coefficients r^i,0{\hat{r}}_{i,0} for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}), respectively.

We present the conformal coefficients r^i,0{\hat{r}}_{i,0} in Table 6.5. The PMC scales are of a perturbative nature, which leads to the first kind of residual scale dependence for PMCm predictions. If the pQCD approximants are known up to four-loop QCD corrections, three PMC scales (Q1Q_{1}, Q2Q_{2} and Q3Q_{3}) can be determined up to N2LL\rm N^{2}LL, NLL\rm NLL and LL\rm LL order, which are {\{41.19, 36.85, 168.68}\} GeV for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), {\{1.26, 0.98, 0.36}\} GeV for R^n(Mτ)\hat{R}_{n}(M_{\tau}) and {\{62.03, 40.76, 52.76}\} GeV for R~n(MH)\tilde{R}_{n}(M_{H}), accordingly. There is no {βi}\{\beta_{i}\}-term to set the scale Q4Q_{4}, the PMCm prediction has the second kind of residual scale dependence. As mentioned in Sec. 3.1, there is the second kind of residual scale dependence for PMCm series and, for convenience, we set Q4=Q3Q_{4}=Q_{3} as the default choice of Q4Q_{4}. A discussion of the magnitude of the second kind of residual scale dependence by taking some other typical choices of Q4Q_{4} will be presented at the end of this subsection. For the scales ΛQCD\gg\Lambda_{\rm QCD}, we adopt the usual approximate four-loop analytic solution of the RGE to derive the value of αs\alpha_{s}. Due to the sizable difference between the approximate analytic solution and the exact numerical solution of the RGE at scales below a few GeV [52, 126], we adopt the exact numerical solution of the RGE to evaluate Rτ(Mτ)R_{\tau}(M_{\tau}) at 1.26 GeV and 0.98 GeV. For the scales close to ΛQCD\Lambda_{\rm QCD}, various low-energy models have been suggested in the literature; a detailed comparison of various low-energy models can be found in Ref. [99]. For definiteness, we shall adopt the Massive Perturbation Theory (MPT) model [218] to evaluate Rτ(Mτ)R_{\tau}(M_{\tau}) at Q3=0.36Q_{3}=0.36 GeV, which gives αs|MPTξ=10±2(0.36)=0.5590.032+0.042\alpha_{s}|_{\rm MPT}^{\xi=10\pm 2}(0.36)=0.559^{+0.042}_{-0.032}, where ξ\xi is the parameter in the MPT model.

n=1n=1 n=2n=2 n=3n=3 κ1\kappa_{1} κ2\kappa_{2} κ3\kappa_{3}
Rn|PMCmR_{n}|_{\rm PMCm} 0.047350.04735 0.046400.04640 0.046100.04610 6.9%6.9\% 2.0%2.0\% 0.6%0.6\%
R^n|PMCm\hat{R}_{n}|_{\rm PMCm} 0.21330.2133 0.19910.1991 0.20870.2087 139.4%139.4\% 6.7%6.7\% 4.8%4.8\%
R~n|PMCm\tilde{R}_{n}|_{\rm PMCm} 0.24810.2481 0.24020.2402 0.24000.2400 22.0%22.0\% 3.2%3.2\% 0.1%0.1\%
Table 6.6: Results for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}), R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the PMCm scale-setting approach. The renormalization scale is set as QQ, MτM_{\tau} and MHM_{H}, respectively.

We present the results of Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the PMCm scale-setting approach in Table 6.6. Table 6.6 shows that the PMCm predictions generally have behavior close to the central predictions under conventional scale-setting procedures, especially when more loop terms are known. This is due to the fact that when the renormalization scale of the conventional series is set as the one to eliminate the large logarithms, the divergent renormalon terms may also be simultaneously removed, since the {βi}\{\beta_{i}\}-terms are always accompanied by the logarithm terms.

LO\rm LO NLO\rm NLO N2LO\rm N^{2}LO N3LO\rm N^{3}LO Total\rm Total
R3|PMCmR_{3}|_{\rm PMCm} 0.042670.00001+0.000030.04267^{+0.00003}_{-0.00001} 0.003490.00004+0.000030.00349^{+0.00003}_{-0.00004} 0.00004-0.00004 0.00002-0.00002 0.046100.00005+0.000060.04610^{+0.00006}_{-0.00005}
R^3|PMCm\hat{R}_{3}|_{\rm PMCm} 0.12720.0090+0.00620.1272^{+0.0062}_{-0.0090} 0.05530.0178+0.01470.0553^{+0.0147}_{-0.0178} 0.0194\hphantom{-}0.0194 0.0068\hphantom{-}0.0068 0.20870.0268+0.02090.2087^{+0.0209}_{-0.0268}
R~3|PMCm\tilde{R}_{3}|_{\rm PMCm} 0.22580.2258 0.02470.0001+0.00010.0247^{+0.0001}_{-0.0001} 0.0093-0.0093 0.0012-0.0012 0.24000.0001+0.00010.2400^{+0.0001}_{-0.0001}
Table 6.7: The value of each loop term (LO, NLO, N2LO and N3LO) for the four-loop QCD predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the PMCm scale-setting approach. The uncertainties correspond to taking μr[Q/2,2Q]\mu_{r}\in[Q/2,2Q] for R3(Q=31.6R_{3}(Q{=}31.6 GeV), μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}] for R^3(Mτ)\hat{R}_{3}(M_{\tau}) and μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] for R~3(MH)\tilde{R}_{3}(M_{H}).

We present the value of each loop term for the four-loop predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the PMCm scale-setting approach in Table 6.7.888By using the RGE recursively, one can obtain correct αs\alpha_{s} value and achieve good matching of αs\alpha_{s} to its coefficients at the same perturbative order. However, such treatment is a sort of {βi}\{\beta_{i}\}-resummation and the resultant PMC series is no longer the usual fixed-order series. Thus, the function of Table 6.7 is to show its own perturbative behavior. The relative importance among the LO terms, the NLO terms, the N2LO terms and the N3LO terms for those approximants are

1:+0.0818:0.0009:0.0005(μr=Q),\displaystyle 1:+0.0818:-0.0009:-0.0005\quad(\mu_{r}{=}Q), (6.12)
1:+0.4347:+0.1525:+0.0535(μr=Mτ),\displaystyle 1:+0.4347:+0.1525:+0.0535\quad(\mu_{r}{=}M_{\tau}), (6.13)
1:+0.1094:0.0412:0.0053(μr=MH).\displaystyle 1:+0.1094:-0.0412:-0.0053\quad(\mu_{r}{=}M_{H}). (6.14)

Table 6.7 shows that there are residual scale dependences of R3(Q=31.6R_{3}(Q{=}31.6 GeV) and R~3(MH)\tilde{R}_{3}(M_{H}) for the LO and NLO terms that are quite small (e.g. the errors are only about ±0.1%\pm 0.1\% of the LO terms and ±0.04%\pm 0.04\% of the NLO terms, respectively), which are smaller than the corresponding ones under the conventional scale-setting approach. However, the residual scale dependence of R^3(Mτ)\hat{R}_{3}(M_{\tau}) is sizable – approximately ±10%\pm 10\% – and is comparable to the conventional scale dependence. Here the large residual scale dependence of R^3(Mτ)\hat{R}_{3}(M_{\tau}) is reasonable, caused by the poor pQCD convergence for the PMC scales at higher orders and the uncertainties of the αs\alpha_{s}-running behavior in the low-energy region.

Q4=2Q3Q_{4}=2Q_{3} Q4=12Q3Q_{4}=\frac{1}{2}Q_{3} Q4=Q1Q_{4}=Q_{1} Q4=Q2Q_{4}=Q_{2} Q4=(Q1+Q2+Q3)/3Q_{4}={(Q_{1}+Q_{2}+Q_{3})}/{3}
R3|PMCmR_{3}|_{\rm PMCm} 0.046110.04611 0.046100.04610 0.046080.04608 0.046080.04608          0.046100.04610
R^3|PMCm\hat{R}_{3}|_{\rm PMCm} 0.20540.0036+0.00540.2054^{+0.0054}_{-0.0036} 0.21050.0053+0.00880.2105^{+0.0088}_{-0.0053} 0.20330.0032+0.00490.2033^{+0.0049}_{-0.0032} 0.20410.0033+0.00500.2041^{+0.0050}_{-0.0033}          0.20460.0034+0.00510.2046^{+0.0051}_{-0.0034}
R~3|PMCm\tilde{R}_{3}|_{\rm PMCm} 0.24040.2404 0.23920.2392 0.24010.2401 0.23980.2398          0.24000.2400
Table 6.8: The four-loop pQCD predictions of R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) under some other choices of the undetermined Q4Q_{4} as an estimate of the second kind of residual scale dependence under the PMCm scale-setting approach. The uncertainty for R^3(Mτ)\hat{R}_{3}(M_{\tau}) is obtained by changing the MPT-model parameter ξ=10±2\xi=10\pm 2.

As a final remark, we discuss the possible magnitudes of the second kind of residual scale dependence under the PMCm scale-setting approach by taking some other typical choices of Q4Q_{4}, e.g. 2Q32Q_{3}, Q3/2Q_{3}/2, Q1Q_{1}, Q2Q_{2} and (Q1+Q2+Q3)/3(Q_{1}+Q_{2}+Q_{3})/3, which also ensures the scheme independence of the PMCm series. Table 6.8 shows that the second kind of residual scale dependence of R3R_{3}, R^3\hat{R}_{3} and R~3\tilde{R}_{3} are (0.00002+0.00001)(^{+0.00001}_{-0.00002}), (0.0054+0.0018)(^{+0.0018}_{-0.0054}) and (0.0008+0.0004)(^{+0.0004}_{-0.0008}), respectively. It shows that those choices will change the magnitudes of R3R_{3}, R^3\hat{R}_{3} and R~3\tilde{R}_{3} at the default choice of Q4=Q3Q_{4}=Q_{3} by about ±0.04%\pm 0.04\%, ±2.6%\pm 2.6\% and ±0.3%\pm 0.3\%, respectively. For R^3(Mτ)\hat{R}_{3}(M_{\tau}), another uncertainty caused by the MPT model parameter ξ=10±2\xi=10\pm 2 is R^3|PMCm=0.20870.0046+0.0072\hat{R}_{3}|_{\rm PMCm}=0.2087^{+0.0072}_{-0.0046} in the choice of Q4=Q3Q_{4}=Q_{3}. We also give the numerical results under some other choices of the undetermined Q4Q_{4} in Table 6.8. It shows that the MPT model parameter ξ=10±2\xi=10\pm 2 will lead to 3%\sim 3\% uncertainties. These uncertainties caused by the small scale Q3Q_{3} and undetermined scale Q4Q_{4} indicate that we still need a more appropriate scale-setting approach to suppress the theoretical uncertainties.

6.4   Properties using the PMCs approach

The PMCs approach provides a method to suppress the residual scale dependence. Applying the standard PMCs scale-setting procedures, we obtain an overall effective αs\alpha_{s} and, accordingly, an overall effective scale (QQ_{*}) for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}), respectively. If they are known up to two-loop, three-loop and four-loop levels, the PMC scale QQ_{*} can be determined up to LL{\rm LL}, NLL{\rm NLL} and N2LL{\rm N^{2}LL} accuracies, respectively. That is, for n=1,2,3n=1,2,3, we have

Q|e+e\displaystyle Q_{*}|_{e^{+}e^{-}} =\displaystyle= {35.36,39.49,40.12}GeV,\displaystyle\{35.36,39.49,40.12\}\,{\rm GeV}, (6.15)
Q|τ\displaystyle Q_{*}|_{\tau}\quad =\displaystyle= {0.90,1.06,1.07}GeV,\displaystyle\{\hphantom{0}0.90,\hphantom{0}1.06,\hphantom{0}1.07\}\,{\rm GeV}, (6.16)
Q|Hbb¯\displaystyle Q_{*}|_{H\to b\bar{b}} =\displaystyle= {60.94,56.51,58.80}GeV.\displaystyle\{60.94,56.51,58.80\}\,{\rm GeV}. (6.17)

Their magnitudes become more precise as one includes more loop terms and the difference between the two nearby values becomes smaller and smaller when more loop terms are included, e.g. the N2LL{\rm N^{2}LL} scales only shift by about 2%4%2\%-4\% to the NLL{\rm NLL} ones. Since these PMCs scales are numerically sizable, one avoids confronting the possibly small scale problem at certain perturbative orders of the multi-scale-setting approaches, such as PMCm, PMC.

n=1n=1 n=2n=2 n=3n=3 κ1\kappa_{1} κ2\kappa_{2} κ3\kappa_{3}
Rn|PMCsR_{n}|_{\rm PMCs} 0.047350.04735 0.046290.04629 0.046130.04613 6.9%6.9\% 2.2%2.2\% 0.3%0.3\%
R^n|PMCs\hat{R}_{n}|_{\rm PMCs} 0.21360.2136 0.19970.1997 0.20640.2064 139.7%139.7\% 6.5%6.5\% 3.4%3.4\%
R~n|PMCs\tilde{R}_{n}|_{\rm PMCs} 0.24810.2481 0.24240.2424 0.23980.2398 22.0%22.0\% 2.3%2.3\% 1.1%1.1\%
Table 6.9: Results for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}), R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the PMCs scale-setting approach, which are independent of any choice of renormalization scale.
LO\rm LO NLO\rm NLO N2LO\rm N^{2}LO N3LO\rm N^{3}LO Total\rm Total
R3|PMCsR_{3}|_{\rm PMCs} 0.042870.04287 0.003380.00338 0.00008-0.00008 0.00004-0.00004 0.046130.04613
R^3|PMCs\hat{R}_{3}|_{\rm PMCs} 0.14650.1465 0.04600.0460 0.0108\hphantom{-}0.0108 0.0031\hphantom{-}0.0031 0.20640.2064
R~3|PMCs\tilde{R}_{3}|_{\rm PMCs} 0.22780.2278 0.02190.0219 0.0088-0.0088 0.0011-0.0011 0.23980.2398
Table 6.10: The value of each loop term (LO, NLO, N2LO or N3LO) for the four-loop predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}), R~3(MH)\tilde{R}_{3}(M_{H}) using the PMCs scale-setting approach.

We present the results of Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the PMCs scale-setting approach in Table 6.9. We also present the value of each loop term for the four-loop approximants R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the PMCs scale-setting approach in Table 6.10. The relative importance among the LO terms, the NLO terms, the N2LO terms and the N3LO terms for those approximants are

1:+0.0788:0.0019:0.0009,\displaystyle 1:+0.0788:-0.0019:-0.0009, (6.18)
1:+0.3140:+0.0737:+0.0212,\displaystyle 1:+0.3140:+0.0737:+0.0212, (6.19)
1:+0.0961:0.0386:0.0048.\displaystyle 1:+0.0961:-0.0386:-0.0048. (6.20)

These are comparable to the convergent behavior of the PMCm series and are more convergent than conventional predictions. Moreover, the sizable residual scale dependence of R^3(Mτ)\hat{R}_{3}(M_{\tau}) appearing in Table 6.7 has been eliminated by using the PMCs procedure. Thus, the PMCs approach, which requires a much simpler analysis, can be adopted as a reliable substitute for the basic PMCm approach to setting the renormalization scales for high-energy processes, with small residual scale dependence. As a conservative estimate of the first kind of residual scale dependence, we take the magnitude of its last known term as the unknown N3LL term, e.g. ±(|S2as2(Q)|)\pm(|S_{2}a^{2}_{s}(Q_{*})|) for R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ){\hat{R}}_{3}(M_{\tau}) and R~3(MH){\tilde{R}}_{3}(M_{H}). We then obtain

R3(Q=31.6GeV)\displaystyle R_{3}(Q{=}31.6\,{\rm GeV}) =\displaystyle= 0.04613±0.00014,\displaystyle 0.04613\pm 0.00014, (6.21)
R^3(Mτ)\displaystyle{\hat{R}}_{3}(M_{\tau}) =\displaystyle= 0.20640.0024+0.0026,\displaystyle 0.2064^{+0.0026}_{-0.0024}, (6.22)
R~3(MH)\displaystyle{\tilde{R}}_{3}(M_{H}) =\displaystyle= 0.23980.0016+0.0014,\displaystyle 0.2398^{+0.0014}_{-0.0016}, (6.23)

which show that the first kind of residual scale dependence is about ±0.3%\pm 0.3\%, ±1.3%\pm 1.3\% and ±0.7%\pm 0.7\%.999Since the N2LL accuracy lnQ2/Q2\ln Q^{2}_{*}/Q^{2}-series of these pQCD approximants already show good perturbative behavior, it is found that by using the PAA predicted N3LL term, e.g. ±|S22/S1as3(Q)|\pm|{S^{2}_{2}}/{S_{1}}a^{3}_{s}(Q_{*})|, to perform the estimates, one can obtain smaller first kind of residual scale dependence than the ones listed in Eqs. (6.21), 6.22, 6.23), which are ±0.00002\pm 0.00002, ()0.0003+0.0002\left({}^{+0.0002}_{-0.0003}\right) and ()0.0009+0.0007\left({}^{+0.0007}_{-0.0009}\right), respectively. More explicitly, we show the conservative estimate of the first kind of residual scale dependence under the PMCs in Figs. 6.1, (6.2) and (6.3), respectively.

Refer to caption
Figure 6.1: The renormalization scale dependence of the four-loop prediction R3(Q=31.6R_{3}(Q{=}31.6 GeV) using the conventional and PMCs scale-setting procedures (from Ref. [62]). The band represents a conservative estimate (6.21) of the first kind of residual scale dependence under the PMCs.
Refer to caption
Figure 6.2: The renormalization scale dependence of the four-loop prediction R^3(Mτ)\hat{R}_{3}(M_{\tau}) using the conventional and PMCs scale-setting procedures (from Ref. [62]). The band represents a conservative estimate (6.22) of the first kind of residual scale dependence under the PMCs.
Refer to caption
Figure 6.3: The renormalization scale dependence of the four-loop prediction R~3(MH)\tilde{R}_{3}(M_{H}) using the conventional and PMCs scale-setting procedures (from Ref. [62]). The band represents a conservative estimate (6.23) of the first kind of residual scale dependence under the PMCs.

The conformal PMCs series is scheme and scale independent; it thus provides a reliable basis for estimating the effect of unknown higher-order contributions. At present, there is no way to use a series with different effective αs(Qi)\alpha_{s}(Q_{i}) at different orders; and if there were any, its effectiveness would also be greatly affected by the possibly large residual scale dependence. Thus, we shall not use PMCm, PMC series to estimate the contribution of the unknown terms. As for the PMCs series, with an overall effective αs(Q)\alpha_{s}(Q_{*}), we may directly use the PAA [219].

       r^4,0{\hat{r}}_{4,0}        r^5,0{\hat{r}}_{5,0}
R(Q)R(Q) [0/2]: 2541.35-2541.35 [0/3]: 181893-181893
R^(Mτ)\hat{R}(M_{\tau}) [0/2]: 1245.3\hphantom{-}1245.3 [0/3]: 15088.6\hphantom{-}15088.6
R~(MH)\tilde{R}(M_{H}) [0/2]: 185951-185951 [0/3]: 3802450\hphantom{-}3802450
Table 6.11: The preferable [0/n10/n{-}1]-type PAA predictions of the conformal coefficients r^4,0{\hat{r}}_{4,0} and r^5,0{\hat{r}}_{5,0} for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}), respectively.
        N3LO\rm N^{3}LO         N4LO\rm N^{4}LO
R3|PMCsR_{3}|_{\rm PMCs} [0/2]: 0.00003-0.00003 [0/3]: 0.000003-0.000003
R^3|PMCs\hat{R}_{3}|_{\rm PMCs} [0/2]: +0.0022+0.0022 [0/3]: +0.0010+0.0010
R~3|PMCs\tilde{R}_{3}|_{\rm PMCs} [0/2]: 0.0019-0.0019 [0/3]: +0.0004+0.0004
Table 6.12: The preferable [0/n10/n{-}1]-type PAA predictions of the N3LO\rm N^{3}LO and N4LO\rm N^{4}LO terms of R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the PMCs scale-setting approach.

We present the preferable [0/n10/n{-}1]-type PAA predictions for the PMCs series of R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) in Table 6.11 and 6.12. Table 6.11 shows the predicted N3LO\rm N^{3}LO and N4LO\rm N^{4}LO conformal coefficients r^4,0{\hat{r}}_{4,0} and r^5,0{\hat{r}}_{5,0}. Note that the predicted r^4,0{\hat{r}}_{4,0} values are close to the exact values shown in Table 6.5 and those known conformal coefficients do not change when more loop terms are known. To obtain the final numerical result, we need to combine the coefficients r^4,0{\hat{r}}_{4,0} and r^5,0{\hat{r}}_{5,0} with the effective αs(Q)\alpha_{s}(Q_{*}) at corresponding orders. Table 6.12 displays the numerical results, these values will be very slightly changed for a more accurate QQ_{*}, since the N2LL{\rm N^{2}LL} accuracy QQ_{*} is already changed from that at NLL{\rm NLL} by less than 5%\sim 5\%.

6.5   Properties using the PMC approach

Given the unique form of intrinsic conformality iCF, any other attempt (such as PMCa [107]) to write the perturbative expansion in a scale-invariant form would lead to the iCF (as shown in Ref. [62]).

Following the standard PMC procedures, we calculate Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections. The perturbative coefficients (𝒞i=(1,,4),IC\mathcal{C}_{i=(1,\cdots,4),\rm{IC}}) are exactly the same as those of the PMCm and PMCs conformal coefficients (r^i=(1,,4),0{\hat{r}}_{i=(1,\cdots,4),0}). As shown by Eqs. (4.12), (4.14), (4.15), the PMC scales are definite and have no perturbative nature, they are free of renormalization scale ambiguities and do not have the first kind of residual scale dependence. Using the four-loop QCD corrections, we can determine their first three scales, i.e.

{μI,μII,μIII}|e+e\displaystyle\{\mu_{\rm I},\mu_{\rm II},\mu_{\rm III}\}|_{e^{+}e^{-}} =\displaystyle= {35.36,71.11,0.003}GeV,\displaystyle\{35.36,71.11,0.003\}\,{\rm GeV}, (6.24)
{μI,μII,μIII}|τ\displaystyle\{\mu_{\rm I},\mu_{\rm II},\mu_{\rm III}\}|_{\tau} =\displaystyle= {0.90,1.16,1.82}GeV,\displaystyle\{0.90,1.16,1.82\}\,{\rm GeV}, (6.25)
{μI,μII,μIII}|Hbb¯\displaystyle\{\mu_{\rm I},\mu_{\rm II},\mu_{\rm III}\}|_{H\to b\bar{b}} =\displaystyle= {60.94,41.24,46.44}GeV.\displaystyle\{60.94,41.24,46.44\}\,{\rm GeV}. (6.26)

For the case R3(Q=31.6R_{3}(Q{=}31.6 GeV), its third scale μIII=0.003\mu_{\rm III}=0.003 GeV is quite small and we adopt the above-mentioned MPT model to estimate its contribution, which gives αs|MPT(0.003)=0.606\alpha_{s}|_{\rm MPT}(0.003)=0.606. As mentioned in Sec. 4, the fourth scale μIV\mu_{\rm IV} is fixed to the initial scale μr\mu_{r}, i.e. the kinematic scale of the process, and varied in the range μIV[μr/2,2μr]\mu_{\rm IV}\in[\mu_{r}/2,2\mu_{r}] to ascertain the level of conformality achieved by the series. In fact, the last PMC scale is entangled with the missing higher-order contributions. This is referred as second kind of residual scale dependence.

n=1n=1 n=2n=2 n=3n=3 κ1\kappa_{1} κ2\kappa_{2} κ3\kappa_{3}
Rn|PMCR_{n}|_{\rm PMC_{\infty}} 0.047500.04750 0.046520.04652 0.039370.03937 7.3%7.3\% 2.1%2.1\% 15.4%15.4\%
R^n|PMC\hat{R}_{n}|_{\rm PMC_{\infty}} 0.18050.1805 0.21120.2112 0.21990.2199 102.6%102.6\% 17.0%17.0\% 4.1%4.1\%
R~n|PMC\tilde{R}_{n}|_{\rm PMC_{\infty}} 0.24380.2438 0.24480.2448 0.24050.2405 19.9%19.9\% 0.4%0.4\% 1.8%1.8\%
Table 6.13: Results for Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}), R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the PMC scale-setting approach. For each case, the undetermined PMC scale of the highest order terms is set as QQ, MτM_{\tau} and MHM_{H}, respectively.
LO\rm LO NLO\rm NLO N2LO\rm N^{2}LO N3LO\rm N^{3}LO Total\rm Total
R3|PMCR_{3}|_{\rm PMC_{\infty}} 0.043830.04383 0.002800.00280 0.00722-0.00722 0.000040.00004+0.00001-0.00004^{+0.00001}_{-0.00004} 0.039370.00004+0.000010.03937^{+0.00001}_{-0.00004}
R^3|PMC\hat{R}_{3}|_{\rm PMC_{\infty}} 0.17610.1761 0.03960.0396 0.0035\hphantom{-}0.0035 0.00070.0005+0.0034\hphantom{-}0.0007^{+0.0034}_{-0.0005} 0.21990.0005+0.00340.2199^{+0.0034}_{-0.0005}
R~3|PMC\tilde{R}_{3}|_{\rm PMC_{\infty}} 0.22650.2265 0.02460.0246 0.0099-0.0099 0.00070.0004+0.0002-0.0007^{+0.0002}_{-0.0004} 0.24050.0004+0.00020.2405^{+0.0002}_{-0.0004}
Table 6.14: The value of each loop term (LO, NLO, N2LO or N3LO) for the four-loop predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the PMC scale-setting approach. The errors, representing the second kind of residual scale dependence, are estimated by varying the undetermined PMC scale μIV\mu_{\rm IV} within the range [Q/2,2Q][Q/2,2Q] for R3(Q=31.6R_{3}(Q{=}31.6 GeV), [1GeV,2Mτ][1\,{\rm GeV},2M_{\tau}] for R^3(Mτ)\hat{R}_{3}(M_{\tau}) and [MH/2,2MH][M_{H}/2,2M_{H}] for R~3(MH)\tilde{R}_{3}(M_{H}).

We present the results of Rn(Q=31.6R_{n}(Q{=}31.6 GeV), R^n(Mτ)\hat{R}_{n}(M_{\tau}) and R~n(MH)\tilde{R}_{n}(M_{H}) up to four-loop QCD corrections using the PMC scale-setting approach in Table 6.13. For the cases Rn(Q=31.6R_{n}(Q{=}31.6 GeV) and R~n(MH)\tilde{R}_{n}(M_{H}), we have κ2<κ3\kappa_{2}<\kappa_{3}, indicating that the second kind of residual scale dependence is sizable for these two quantities, which largely affects the magnitude of the lower-order series. When one has enough higher-order terms, the residual scale dependence is highly suppressed owing to the more convergent renormalon-free series. For example, we present the value of each loop term (LO, NLO, N2LO or N3LO) for the four-loop predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) in Table 6.14. At the four-loop level, the PMC series already exhibits convergent behavior. As shown in Table 6.14, the relative importance among the LO terms, the NLO terms, the N2LO terms and the N3LO terms for those approximants are

1:+0.0639:0.1647:0.0009(μr=Q),\displaystyle 1:+0.0639:-0.1647:-0.0009\quad(\mu_{r}{=}Q), (6.27)
1:+0.2249:+0.0199:+0.0040(μr=Mτ),\displaystyle 1:+0.2249:+0.0199:+0.0040\quad(\mu_{r}{=}M_{\tau}), (6.28)
1:+0.1086:0.0437:0.0031(μr=MH).\displaystyle 1:+0.1086:-0.0437:-0.0031\quad(\mu_{r}{=}M_{H}). (6.29)

This perturbative behavior is similar to the predictions of PMCm and PMCs, except for R3(Q=31.6R_{3}(Q{=}31.6 GeV), which due to a much smaller scale μIII\mu_{\rm III} leads to quite large N2LO terms.

6.6   A comparison of the renormalization scale dependence of the various PMC approaches

We present the renormalization scale (μr\mu_{r}) dependence of the four-loop predictions R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) and R~3(MH)\tilde{R}_{3}(M_{H}) using the conventional, PMCm, PMCs and PMC scale-setting procedures in Figs. 6.4, 6.5 and 6.6, respectively. In these figures, we show the second kind of residual scale dependence of R3(Q=31.6R_{3}(Q{=}31.6 GeV), R^3(Mτ)\hat{R}_{3}(M_{\tau}) under the PMCm and PMC scale-setting procedures with the shaded bands. The green/lighter bands are obtained by changing the undetermined Q4Q_{4} to 2Q32Q_{3}, Q3/2Q_{3}/2, Q1Q_{1}, Q2Q_{2} and (Q1+Q2+Q3)/3(Q_{1}+Q_{2}+Q_{3})/3. And the red/darker bands are obtained by varying the undetermined PMC scale μIV\mu_{\rm IV} within the range [Q/2,2Q][Q/2,2Q] for R3(Q=31.6R_{3}(Q{=}31.6 GeV), [1GeV,2Mτ][1\,{\rm GeV},2M_{\tau}] for R^3(Mτ)\hat{R}_{3}(M_{\tau}) and [MH/2,2MH][M_{H}/2,2M_{H}] for R~3(MH)\tilde{R}_{3}(M_{H}).

Refer to caption
Figure 6.4: The renormalization scale dependence of the four-loop prediction R3(Q=31.6R_{3}(Q{=}31.6 GeV) using the conventional PMCm, PMCs and PMC scale-setting procedures (from Ref. [62]). The green/lighter band represents the second kind of residual scale dependence under the PMCm, which is obtained by changing the undetermined Q4Q_{4} to be 2Q32Q_{3}, Q3/2Q_{3}/2, Q1Q_{1}, Q2Q_{2} and (Q1+Q2+Q3)/3(Q_{1}+Q_{2}+Q_{3})/3, respectively. The red/darker band represents the second kind of residual scale dependence under the PMC, which is obtained by varying the undetermined PMC scale μIV\mu_{\rm IV} within the range [Q/2,2Q][Q/2,2Q]. The experimental result Rexpt(Q=31.6GeV)=0.0527±0.0050R^{\rm expt}(Q=31.6\,{\rm GeV})=0.0527\pm 0.0050 is extracted from 311Re+eexpt=1.0527±0.0050\frac{3}{11}R_{e^{+}e^{-}}^{\rm expt}=1.0527\pm 0.0050 [208].
Refer to caption
Figure 6.5: The renormalization scale dependence of the four-loop prediction R^3(Mτ)\hat{R}_{3}(M_{\tau}) using the conventional, PMCm, PMCs and PMC scale-setting procedures (from Ref. [62]). The green/lighter band represents the second kind of residual scale dependence under the PMCm, which is obtained by changing the undetermined Q4Q_{4} to be 2Q32Q_{3}, Q3/2Q_{3}/2, Q1Q_{1}, Q2Q_{2} and (Q1+Q2+Q3)/3(Q_{1}+Q_{2}+Q_{3})/3, respectively. The red/darker band represents the second kind of residual scale dependence under the PMC, which is obtained by varying the undetermined PMC scale μIV\mu_{\rm IV} within the region of [1GeV,2Mτ][1\,{\rm GeV},2M_{\tau}]. The experimental result R^expt(Mτ)=0.20220.0038+0.0038\hat{R}^{\rm expt}(M_{\tau})=0.2022^{+0.0038}_{-0.0038} is extracted from Rτexpt(Mτ)=3.475±0.011R_{\tau}^{\rm expt}(M_{\tau})=3.475\pm 0.011 [220].

Figure 6.4 shows that the theoretical predictions are smaller than the experimental result. This is reasonable since we have adopted the world average αs(MZ)=0.1179\alpha_{s}(M_{Z})=0.1179 [126] to set ΛQCD\Lambda_{\rm QCD} for all these observables and, if we adopt a strong coupling αs(MZ)\alpha_{s}(M_{Z}) fixed by using the e+ee^{+}e^{-} annihilation data alone, we obtain consistent predictions in agreement with the data. For example, using αs(MZ)=0.1224\alpha_{s}(M_{Z})=0.1224 [199], fixed by using the hadronic event shapes in e+ee^{+}e^{-} annihilation to set ΛQCD\Lambda_{\rm QCD}, we obtain a larger R3(Q=31.6R_{3}(Q{=}31.6 GeV), e.g. R3(Q=31.6GeV)=0.04826R_{3}(Q=31.6\,{\rm GeV})=0.04826 for the PMCs approach, which is consistent with Rexpt(Q=31.6GeV)=0.0527±0.0050R^{\rm expt}(Q=31.6\,{\rm GeV})=0.0527\pm 0.0050 within errors. It has been noticed that the second kind of residual scale dependence of R3(Q=31.6R_{3}(Q{=}31.6 GeV) under the PMCm and PMC scale-setting procedure are both very small, since the order αs4\alpha_{s}^{4} correction is highly suppressed in R3(Q=31.6R_{3}(Q{=}31.6 GeV). These figures show that by including enough higher-order terms, the following hold.

Refer to caption
Figure 6.6: The renormalization scale dependence of the four-loop prediction R~3(MH)\tilde{R}_{3}(M_{H}) using the conventional, PMCm, PMCs and PMC scale-setting procedures (from Ref. [62]). The green/lighter band represents the second kind of residual scale dependence under the PMCm, which is obtained by changing the undetermined Q4Q_{4} to be 2Q32Q_{3}, Q3/2Q_{3}/2, Q1Q_{1}, Q2Q_{2} and (Q1+Q2+Q3)/3(Q_{1}+Q_{2}+Q_{3})/3, respectively. The red/darker band represents the second kind of residual scale dependence under the PMC, which is obtained by varying the undetermined PMC scale μIV\mu_{\rm IV} within the region of [MH/2,2MH][M_{H}/2,2M_{H}].
  • 1.

    The renormalization scale dependence of the conventional prediction depends strongly on the convergence of the perturbative series and the cancellation of scale dependence among different orders. For a numerically strongly convergent series, such as R3(Q=31.6R_{3}(Q{=}31.6 GeV) and R~3(MH)\tilde{R}_{3}(M_{H}), the net scale dependence is only parts per thousand for a wide range of scale choices. For a less convergent series, such as R^3(Mτ)\hat{R}_{3}(M_{\tau}), the net renormalization scale uncertainty is sizable, which is up to 18%\sim 18\% for μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}], 24%\sim 24\% for μr[1GeV,3Mτ]\mu_{r}\in[1\,{\rm GeV},3M_{\tau}] and 28%\sim 28\% for μr[1GeV,5Mτ]\mu_{r}\in[1\,{\rm GeV},5M_{\tau}];

  • 2.

    The PMCm predictions have two kinds of residual scale dependence due to unknown terms. The second residual scale dependence can be greatly suppressed by the extra requirement of conformal invariance; the first property then dominates the net residual scale dependence. For numerically convergent series, such as R3(Q=31.6R_{3}(Q{=}31.6 GeV) and R~3(MH)\tilde{R}_{3}(M_{H}), the residual scale dependences are small, i.e. less than four parts per thousand.101010As a comparison, the conventional scale dependence of R~3(MH)\tilde{R}_{3}(M_{H}) is about nine parts per thousand for μr[MH/2,2MH]\mu_{r}\in[M_{H}/2,2M_{H}] For a less convergent series, such as R^3(Mτ)\hat{R}_{3}(M_{\tau}), due to the large residual scale dependence of the NLO terms, its net residual scale dependence is sizable, it is 23%\sim 23\% for μr[1GeV,2Mτ]\mu_{r}\in[1\,{\rm GeV},2M_{\tau}], 28%\sim 28\% for μr[1GeV,3Mτ]\mu_{r}\in[1\,{\rm GeV},3M_{\tau}] and 32%\sim 32\% for μr[1GeV,5Mτ]\mu_{r}\in[1\,{\rm GeV},5M_{\tau}]. Although in some special cases, such as R^3(Mτ)\hat{R}_{3}(M_{\tau}), the residual scale dependence may be comparable to the conventional prediction, the PMCm series has no renormalon divergence and it generally has a better pQCD convergence. For the case of R3(Q=31.6R_{3}(Q{=}31.6 GeV) and R~3(MH)\tilde{R}_{3}(M_{H}), the PMCm predictions show weaker dependence on μr\mu_{r} and its prediction can be more accurate than conventional pQCD predictions;

  • 3.

    The PMC predictions only have the second kind of residual scale dependence, which are suppressed for the present four-loop predictions. The magnitude of the residual scale dependence depends on the convergence of the resultant series and for the present processes, the second kind of residual scale dependence are only about parts per thousand to a few percent. Due to the application of  “intrinsic conformality” or equivalently the requirement of scale invariance at each order, the PMC scales determined are not of a perturbative nature, but they can be very small in certain cases. For the case R3(Q=31.6R_{3}(Q{=}31.6 GeV), we obtain a much smaller scale μIII=0.003\mu_{\rm III}=0.003 GeV, which is unreasonable and indicates that the PMC approach may not be applicable for this process. To obtain a numerical estimate, we have adopted the MPT model to calculate the magnitude of αs\alpha_{s} at such a small scale; Fig. 6.4 shows that the MPT prediction deviates from other approaches by about 15%15\%. By including the uncertainty from the MPT model parameter ξ=10±2\xi=10\pm 2, the PMC{\rm PMC}_{\infty} prediction still deviates from other approaches by about 11%11\%;

  • 4.

    The PMCs predictions for the dependence of observables on the renormalization scale are flat lines. The first kind of residual scale dependence of the PMCs predictions only affects the precision of the magnitude of effective αs\alpha_{s} and the PMCs predictions are exactly independent of the choice of μr\mu_{r} at any fixed order.

7   Summary

The Principle of Maximum Conformality (PMC) provides a rigorous first-principles method to eliminate conventional renormalization scheme and scale ambiguities for high-momentum-transfer processes. Its predictions have a solid theoretical foundation, satisfying renormalization group invariance and all other self-consistency conditions derived from the renormalization group. The PMC has now been successfully applied to many high-energy processes.

In this review, we have presented a new scale-setting procedure, namely PMC, which stems from the general PMC and preserves a particular property that we have defined as intrinsic conformality (iCF). The iCF is a particular parametrization of the perturbative series that exactly preserves the scale invariance of an observable perturbatively. We point out that this is a unique property of the perturbative expansion, any other attempt (such as PMCa [107]) to write the perturbative expansion in a scale-invariant form would lead to the iCF (as shown in Ref. [62]).

The PMC solves the conventional renormalization scale ambiguity in QCD, it preserves not only the iCF but also all the features of the PMC approach and leads to a final conformal series at any order of the perturbative calculation. In fact, the final series is given by perturbative conformal coefficients with the couplings determined at conformal renormalization scales. The PMC scale setting agrees with the Gell-Man–Low scheme and can be considered the non-Abelian analog of Serber–Uehling [221, 222] scale setting, which is essential in precision tests of QED and atomic physics.

Given the iCF form, a new “How-To” method for identifying conformal coefficients and scale has been developed and can be applied to either numerical or analytical calculations. The PMC has been applied to the NNLO thrust and CC-parameter distributions and the results show perfect agreement with the experimental data.

The evaluation of theoretical errors using standard criteria demonstrate that the PMC significantly improves the theoretical predictions over the entire spectra of the shape variables order-by-order and both the IR conformal and QED limits of thrust respect the theoretical consistency requirements. Moreover, the position of the thrust peak is in perfect agreement with experiment and is preserved on varying NfN_{f}. Hence, even though for the thrust distribution the peak stems directly from resummation (or partial resummation) of the large logarithms in the low-momentum region, its correct position is fixed by the PMC scale and can be considered a “conformal” property, given its independence from the NfN_{f} or βi\beta_{i} terms.

Unlike the previous BLM/PMC approaches, the PMC scales are not perturbatively calculated but are conformal functions of the physical scale(s) of the process and any other unintegrated momentum or variable, e.g. the event-shape variable (1T)(1-T) or CC. The PMC is totally independent of the initial scale μ0\mu_{0} used for renormalization in perturbative calculations and it preserves the scale invariance at all stages of calculation, independently of the kinematic boundary conditions, of the starting order of the observable or of the order of the truncated expansion. Moreover, this property leads to the possibility of determining the entire coupling from a single experiment at a single center-of-mass energy (this new method is in progress and will soon appear.)

The iCF improves the general BLM/PMC procedure and point “3” of Section 3. In the same section, we suggested that an improvement and simplification of the perturbatively calculated BLM/PMC scales, would be achieved by setting the renormalization scale μr\mu_{r} directly to the physical scale QQ of the process, before applying the BLM/PMC procedure. This would remove the initial scale μ0\mu_{0} dependence from the perturbatively calculated BLM/PMC scales.

We stress that, in contrast with the other PMCm and PMCs approaches, the PMC preserves the iCF; scales are thus set straightforwardly in kinematic regions where constraints cancel the effects of the lower-order conformal coefficients. These effects are particularly visible in the case of event-shape variables in the multi-jet region. For this case we have shown only fixed-order calculation results and other effects due to factorization, such as large logarithms coming from soft and collinear configurations, have not been included. The iCF effects in these kinematic regions are neglected by the PMCm and PMCs approach, unless an ad hoc prescription is introduced.

Another application of the PMC is presented in Ref. [72] and shows an improvement of the results on Γ(Hgg)\Gamma(H\to gg) with respect to the CSS also in this process. From the detailed comparison shown in Sec. 6, it follows that, though the application of the PMC improves the theoretical predictions also for the Re+eR_{e^{+}e^{-}}, RτR_{\tau}, Γ(Hbb¯)\Gamma(H\to b\bar{b}) with respect to the CSS, the PMCs leads to more stable results for these quantities.

In general:

  • \circ

    The PMCs approach determines an overall effective αs\alpha_{s} by eliminating all the RG-dependent nonconformal {βi}\{\beta_{i}\}-terms; this results in a single effective scale which effectively replaces the individual PMC scales of PMCm approach in the sense of a mean-value theorem. The PMCs prediction is renormalization scale-and-scheme independent up to any fixed order. The first kind of residual scale dependence is highly suppressed, since the PMC scale at all known orders is determined at the same highest-order accuracy. There is no second kind of residual scale dependence. The PMCs prediction also avoids the small-scale problem, which sometimes emerges in multi-scale approaches.

  • \circ

    The PMC approach fixes the PMC scales at each order by using the property of intrinsic conformality, which ensures scale invariance of the pQCD series at each order. The resulting PMC scales have no ambiguities, are not of a perturbative nature and thus avoid the first kind of residual scale dependence. Since the last effective scale of the highest-order perturbative term is set to the kinematic scale or physical scale of the process, the PMC prediction still has a reduced second kind of residual scale dependence. However, when more loop terms are included and scales are not in the nonperturbative regime, all PMC’s lead to similar results.

The PMCs approach is close to the PMCm approach in achieving the goals of the PMC by inheriting most of the features of the PMCm approach. It works remarkably well with fully integrated quantities, but has some difficulties in application to differential distributions, besides the fact that this approach may have an effect of averaging the differences of the PMC scales arising at each order, which might be significant to achieve a given precision at a certain level of accuracy. However, given the small differences that we have found in the first two consecutive PMC scales for thrust and CC-parameter, i.e. μI,μII\mu_{I},\mu_{II}, in the LO allowed kinematic region, i.e. 0<(1T)<1/30<(1-T)<1/3 and in 0<C<0.750<C<0.75, we may argue that in the same accessible kinematic domain two consecutive PMC scales have such small differences that a single-scale approach, such as the PMCs, would be justified, leading to analogously precise predictions.

We recall that only the nfn_{f} terms related to the UV-divergent diagrams (i.e. the NfN_{f} terms) must be reabsorbed into the PMC scales. Thus, PMC perfectly agrees with the PMCm when an observable has a manifestly iCF form. We remark that the iCF underlies scale invariance perturbatively, i.e. the ordered scale invariance. We also remark that PMC agrees with the single-scale approach PMCs in the case of an observable with a particular iCF form with all scales equal, i.e. μI=μII=μIII==μN\mu_{\rm I}=\mu_{\rm II}=\mu_{\rm III}=\cdots=\mu_{\rm N}. In this sense, the PMCm and PMCs may be considered more as “optimization procedures” that follow the purpose of the maximal conformal series by transforming the original perturbative series into an iCF-like final series by using the PMC scales. In contrast, the PMC does not indicate any particular value of the renormalization scale to be used, but indicates the final limit obtained by each conformal subset and then by the perturbative expansion, once all the terms related to each conformal subset are resummed.

The PMC is RG invariant at each order of accuracy, which means we may perform a change of scale at any stage and reobtain the initial perturbative quantity. In this sense PMC is not to be understood as an “optimization procedure”, but as an explicit RG-invariant form to parametrize a perturbative quantity that leads to the conformal limit faster. By setting the renormalization scale of each subset to the corresponding PMC scale, one simply cancels the infinite series of β\beta terms, leading to the same conformal result as the original series. Given that both scales and coefficients are conformal in the PMC, the scheme and scale dependence is also completely removed in the perturbative series up to infinity.

It was pointed out in Sec. 6 that the PMC scale might become quite small at a certain order for the case of fully integrated quantities, whose calculations were carried out using the analyticity property of the Adler function (this seems not to occur in direct multi-loop calculations, e.g. for the shape variables); and that the PMC retains the second kind of scale dependence. We stress that the last scale in the PMC controls the level of convergence and the conformality of the perturbative series and is thus entangled with the theoretical error of a given prediction. According to the PMC procedure the last scale must be set to the invariant physical scale of the process, given by s,MH,\sqrt{s},M_{H},\dots. In this review we have shown that the usual PMC practice of setting the last scale equal to the last unknown scale is also consistent for the PMC and leads to precise and stable results. Improvements to these points are currently under investigation.

We finally remark that the evaluation of the theoretical errors using standard criteria shows that the PMC significantly improves the precision of pQCD calculations and eliminates the scheme and scale ambiguities. An improved and more reliable analysis of theoretical errors might be obtained by using a statistical approach for evaluating the contributions of the uncalculated higher-order terms, as suggested in Refs. [214, 215, 216] and recently applied with the PMCs in Refs. [81, 217]. This implementation would lead to a more rigorous method to evaluate errors, also giving indications on the possible range of values for the last unknown PMC scale.

Acknowledgements

We thank Francesco Sannino, André Hoang for useful discussions. XGW is supported in part by the Natural Science Foundation of China under Grant No.12175025 and No.12147102. SQW is supported in part by the Natural Science Foundation of China under Grant No.12265011. SJB is supported in part by the Department of Energy Contract No. DE-AC02-76SF00515. SLAC-PUB-17737.

References