QCD Corrections to in Type-I THDM at Electron Positron Colliders
Abstract
We investigate in detail the charged Higgs production associated with a boson at electron-positron colliders within the framework of the Type-I two-Higgs-doublet model (THDM). We calculate the integrated cross section at the LO and analyze the dependence of the cross section on the THDM parameters and the colliding energy in a benchmark scenario of the input parameters of Higgs sector. The numerical results show that the integrated cross section is sensitive to the charged Higgs mass, especially in the vicinity of at a collider, and decreases consistently as the increment of in the low region. The peak in the colliding energy distribution of the cross section arises from the resonance of loop integrals and its position moves towards low colliding energy as the increment of . We also study the two-loop NLO QCD corrections to both the integrated cross section and the angular distribution of the charged Higgs boson, and find that the QCD relative correction is also sensitive to the charged Higgs mass and strongly depends on the final-state phase space. For , the QCD relative correction at a collider varies in the range of as increases from to .
- keywords
-
two-loop QCD correction, Type-I THDM, charged Higgs boson
I Introduction
In July 2012, the Higgs boson has been discovered by ATLAS and CMS collaborations at CERN Large Hadron Collider (LHC) Aad et al. (2012); Chatrchyan et al. (2012). In addition to measuring the Higgs boson precisely, great efforts have been made to search for exotic Higgs bosons in various scenarios beyond the standard model (SM). Among all the extensions of the SM, the two-Higgs-doublet model (THDM) is an appealing one. It provides rich phenomena such as charged Higgs bosons, explicit and spontaneous -violation, and the candidate for dark matter, since the Higgs sector of the THDM is composed of two complex scalar doublets Lee (1973); Gunion et al. (2000); Gunion and Haber (2003); Branco et al. (2012). In the THDM, there are five Higgs bosons: two neutral -even Higgs bosons and (), two charged Higgs bosons and a neutral -odd Higgs boson . Although both of the two -even scalars can be interpreted as the Higgs boson in the alignment limit, we assume that the lighter scalar is the Higgs boson in this paper. Since the flavor changing neutral currents (FCNCs) can be induced in THDM which have not been observed, an additional symmetry is imposed to eliminate FCNCs at the tree level. Depending on the types of the Yukawa interactions between fermions and the two Higgs doublets, one can introduce several different types of THDMs (Type-I, Type-II, lepton-specific, and flipped) Branco et al. (2012). The most investigated THDMs are the Type-I and Type-II THDMs. In the Type-I THDM, all the fermions only couple to one of the two Higgs doublets, while in the Type-II THDM, the up-type and down-type fermions couple to the two Higgs doublets, respectively.
The THDMs have been widely studied in many different aspects in previous works. Since it is possible to introduce the spontaneous -violation in THDM, it has been considered as a solution to the problem of baryogenesis in Refs.Turok and Zadrozny (1991); Davies et al. (1994); Fromme et al. (2006). In Refs.Deshpande and Ma (1978); Dolle and Su (2009), the neutral scalar in the inert THDM is interpreted as the candidate for dark matter. The existence of charged Higgs boson is one important characteristic of new physics beyond the SM. Therefore, searching for charged Higgs boson in various aspects is a high priority of experiments. At the LHC Run II, the charged Higgs boson has been probed in various channels, such as Sirunyan et al. (2017, 2020); Aaboud et al. (2018a, b).
To match the precise experimental data, the theoretical predictions on kinematic observables should be calculated with high precision. The renormalization of the THDM has been detailedly studied in different renormalization schemes in Refs.Santos and Barroso (1997); Degrande (2015); Krause et al. (2016); Denner et al. (2016); Altenkamp et al. (2017). The production mechanisms and decay modes of the charged Higgs boson have been investigated at one-loop level in the THDM. The Drell-Yan production of charged Higgs pair has been studied at NLO in Refs.Arhrib and Moultaka (1999); Heinemeyer and Schappacher (2016). The full one-loop contributions for the charged Higgs production associated with a vector boson were given in Refs.Zhu ; Arhrib et al. (2000); Kanemura (2000a); Heinemeyer and Schappacher (2016). The dominant decays of the charged Higgs boson into and have been studied in Refs.Drees and Roy (1991); Roy (1999), and the loop-induced decay modes and have also been investigated in Refs.Capdequi Peyranere et al. (1991); Kanemura (2000b); Hernandez-Sanchez et al. (2004); Arhrib et al. (2007). In this work, we focus on the associated production at electron-positron colliders in the Type-I THDM. This production channel is a loop-induced process at the lowest order due to the absence of tree-level and couplings, and has been investigated at one-loop level in the THDM as well as the minimal supersymmetric standard model Farris et al. (2004); Kanemura et al. (2011); Logan and Su (2002, 2003). In order to test the THDM via couplings precisely, we study in detail the two-loop QCD corrections to the process, provide the NLO QCD corrected integrated and differential cross sections, and discuss the dependence on the THDM parameters and the colliding energy.
The rest of this paper is organized as follows. In Sec.II, we give a brief review of the Type-I THDM and provide the benchmark scenario that we adopt. The methods and details of our LO and NLO calculations are presented in Sec.III. In Sec.IV, the numerical results for both integrated and differential cross sections and some discussions are provided. Finally, a short summary is given in Sec.V.
II Two-Higgs-Doublet model
The Higgs sector of the THDM is composed of two complex scalar doublets and , which are both in the representation of the gauge group. In this paper, we consider only the -conserving THDM with a discrete symmetry of the form . Then the renormalizable and gauge invariant scalar potential is given by
(1) | ||||
Since the parameter has the mass-dimension , the terms of this kind only break the symmetry softly which can be retained. The parameters have to be real since the Lagrangian must be real. Though the parameters and can be complex, the imaginary parts of these two parameters would induce explicit violation that we do not consider in this paper. So we assume all the parameters in Eq.(1) are real. The minimization of the potential in Eq.(1) gives two minima and of the form
(2) |
where and are the vacuum expectation values of the neutral components of the two Higgs doublets and , respectively. With respect to the convention in Ref.Eriksson et al. (2010), we define and , where . Expanding at the minima, the two complex Higgs doublets can be expressed as
(3) |
The mass eigenstates of the Higgs fields are given by Branco et al. (2012); Altenkamp et al. (2017)
(4) |
(5) |
(6) |
where is the mixing angle of the neutral -even Higgs sector. After the spontaneous electroweak symmetry breaking, the charged and neutral Goldstone fields and are absorbed by the weak gauge bosons and , respectively. Thus, the THDM predicts the existence of five physical Higgs bosons: two neutral -even Higgs bosons and , one neutral -odd Higgs boson , and two charged Higgs bosons . The masses of these physical Higgs bosons are given by
(7) | ||||
where is the mass matrix of the neutral -even Higgs sector. The explicit form of is expressed as
(8) |
where
(9) |
The SM Higgs is the combination of the two neutral -even Higgs bosons as
(10) | ||||
Thus, the lighter neutral -even scalar can be identified as the SM-like Higgs boson in the so-called alignment limit of . In this paper, we consider the lighter -even Higgs as the SM-like Higgs boson discovered at the LHC.
The input parameters for the Higgs sector of the THDM are chosen as
(11) |
which are implemented as the “physical basis” in 2HDMC Eriksson et al. (2010). We adopt the following benchmark scenario,
(12) | ||||
which satisfies the theoretical constraints from perturbative unitarity Grinstein et al. (2016), stability of vacuum Nie and Sher (1999), and tree-level unitarity Akeroyd et al. (2000). The soft-breaking parameter is chosen as in order to satisfy the perturbative unitarity for . Considering the constraints from experiments at the LHC in Ref.Sirunyan et al. (2019), is very closed to . So we apply the alignment limit in our benchmark scenario to set . As to the and parameters, we concentrate on the region with low mass and small .
III Descriptions of calculation
In this section, we present in detail the calculation procedure for the process at one- and two-loop levels. The Feynman diagrams and the amplitudes are generated by FeynArts-3.11 Hahn (2001), using the Feynman rules of THDM in Ref.Altenkamp et al. (2017). The evaluation of Dirac trace and the contraction of Lorentz indices are performed by the FeynCalc-9.3 package Mertig et al. (1991); Shtabovenko et al. (2016). In order to reduce the Feynman integrals into the combinations of a small set of integrals called the master integrals (MIs), we utilize the KIRA-1.2 package Maierhöfer et al. (2018), which adopts the integration-by-parts (IBP) method with Laporta’s algorithm Laporta (2000). One can get the numerical results of amplitudes after evaluating the MIs which is the main difficulty of multi-loop calculation.
In this paper, we calculate the MIs by using the ordinary differential equations (ODEs) method Liu et al. (2018). A -loop Feynman integral can be expressed as
(13) |
where and are the denominators of Feynman propagators in which the are the linear combinations of loop momenta and external momenta. The physical results of the Feynman integrals are obtained by taking , i.e.,
(14) |
One can construct the ODEs with respect to ,
(15) |
where is a complete set of MIs. The boundary conditions of these ODEs are chosen at which are the simple vacuum integrals. The analytical expressions for the vacuum integrals up to three-loop order can be found in Refs.Davydychev and Tausk (1993); Broadhurst (1999); Kniehl et al. (2017). To solve these ODEs numerically, we utilize the odeint package Ahnert et al. (2011) to evaluate ODEs from an initial point to a target point . To perform the asymptotic expansion in the domain nearby , we transform the coefficient matrix into normalized fuchsian form with the help of the epsilon package Prausa (2017).
III.1 Calculation at LO
For the LO calculation, we adopt the ’t Hooft-Feynman gauge with the on-shell renormalization scheme at one-loop order mentioned in Ref.Altenkamp et al. (2017). Some representative Feynman diagrams for the process at the LO are shown in Fig.1, where and in the loops represent the Higgs/Goldstone and weak gauge bosons, respectively. Due to the tiny mass of electron, the contribution from the diagrams involving Higgs Yukawa coupling to electron is ignored. The diagrams with mixing are also not shown in Fig.1, because these diagrams can induce a factor of via Dirac equation. The last diagram in Fig.1 is a vertex counterterm diagram induced by the renormalization constant at one-loop level. In the on-shell renormalization scheme, the renormalization constant is given by
(16) |
where is the transition of at , and means to take the real parts of the loop integrals in the transition. It is worth mentioning that Eq.(16) is valid at both and .












III.2 Calculation at QCD NLO
There are two-loop and counterterm diagrams for at the QCD NLO. Some representative ones of them are depicted in Fig.2. At the QCD NLO, all the two-loop Feynman diagrams are generated from the LO quark triangle loop diagram (i.e., the first diagram in Fig.1). The cross in the quark loop diagrams represents the renormalization constant of quark mass at , while the circle cross displayed in the last counterterm diagram represents the renormalization constant at . The quark mass renormalization constant used in NLO QCD calculation is given by Bernreuther et al. (2005)
(17) |
where and . The lowest order for is one-loop order, therefore, the renormalization should be dealt with carefully in NLO QCD calculation. As shown in Figs.1 and 2, the wave-function renormalization constant is involved in both NLO and LO amplitudes. Since the self-energy is nonzero at , i.e. is nonzero at , the contribution from the last diagram in Fig.2 should be included in NLO QCD calculation. The typical Feynman diagrams for transition at are shown in Fig.3. After taking into account all the contributions at in -dimensional spacetime, both and singularities are all canceled.








IV Numerical results and discussion
Besides the input parameters for the Higgs sector of the THDM specified in benchmark scenario in Eq.(12), the following SM input parameters are adopted in our numerical calculation Tanabashi et al. (2018):
(18) | ||||
where is the Fermi constant. The fine structure constant is fixed by
(19) |
IV.1 LO
In Fig.4, we display the LO cross section for as a function of and in the benchmark scenario in Eq.(12) at (left) and (right), respectively. From the left plot, we can see clearly that the LO cross section for production at a collider peaks at due to the resonance effect of loop integrals. The cross section is sensitive to the mass of charged Higgs boson, it can exceed in the vicinity of at small . In the region of , the cross section increases slowly as the increment of , while it drops rapidly when . As the increment of from to , the cross section decreases consistently due to the decline of the Yukawa coupling strength in the low region. Comparing the two plots in Fig.4, we can see that the cross section at is much smaller than that at because of the -channel suppression. As the increasing of the colliding energy from to , the peak position of the cross section as a function of moves towards high and the dependence of the cross section is reduced significantly. Moreover, the production cross section at also decreases quickly as the increment of in the plotted region.


In Fig.5, we present the dependence of the LO cross section for on the colliding energy for some typical values of and . As shown in this figure, the behaviors of the production cross section as a function of the colliding energy at different values of are quite similar. For , the cross section increases sharply in the range of , reaches its maximum at , and then decreases slowly as the increment of . The existence of the peak at can be attributed to the competition between the phase-space enlargement and the -channel suppression as the increasing of . The maximum value of the cross section can exceed for and decreases to about and for and , respectively. Comparing the upper two plots of Fig.5, we can see that the dependence of the cross section for is very close to that for , but there is a small peak at for . Such resonance induced by loop integrals only occurs above the threshold of , i.e., . As the increment of , this resonance effect becomes more considerable and the peak position moves towards low . As shown in the bottom-left plot of Fig.5, the resonance peak for is located at , and is more distinct compared to that for . As to the dependence of the production cross section for shown in the bottom-right plot of Fig.5, it looks quite different from those for and . There is a sharp peak at for each value of which was also mentioned in previous works Zhu ; Kanemura (2000a); Heinemeyer and Schappacher (2016). This peak is a consequence of competition among the phase-space enlargement, -channel suppression and the resonance induced by loop integrals. We can see that the peak cross section can reach about for , and will decrease to around and when increases to and , respectively. In the region of , the cross section increases quickly as the increment of due to the enlargement of phase space, while in the region , it is close to the result of -channel suppression, especially at high .




IV.2 NLO QCD
In this subsection, we calculate the process at the QCD NLO, and discuss the dependence of the integrated cross section on the colliding energy and the charged Higgs mass as well as the angular distribution of the final-state charged Higgs boson.
The LO, NLO QCD corrected integrated cross sections and the corresponding QCD relative correction for as functions of the colliding energy are depicted in Fig.6, where and . As shown in the upper panel of this figure, the NLO QCD corrected integrated cross section peaks at , it increases sharply when and decreases approximately linearly in the region of as the increment of . From the lower panel of Fig.6, we can see that the QCD relative correction increases rapidly from about to above as the increment of from to and then decreases back to about as increases to . The variation of QCD relative correction with in the region of is also plotted in the inset in the upper panel of Fig.6 for clarity. It clearly shows that the QCD relative correction decreases approximately linearly from about to around as the increment of from to . In Table 1 we list the LO, NLO QCD corrected cross sections and the corresponding QCD relative corrections at some specific colliding energies. At which can be reached by both the International Linear Collider (ILC) Behnke et al. (2013) and the Future Circular Electron-Positron Collider (FCC-ee) Abada et al. (2019), the cross section is about at the QCD NLO. At the ILC with and , the NLO QCD corrected cross sections reach about and , respectively, and the corresponding relative corrections are and .

[GeV] | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|
[fb] | ||||||||||
[fb] | ||||||||||
[%] |
In order to study the dependence of the QCD correction, we plot the NLO QCD corrected cross section, as well as the LO cross section, and the QCD relative correction as functions of in Fig.7 with and . The numerical results for some typical values of are also given in Table 2. From Fig.7, we can see that the LO and NLO QCD corrected cross sections increase from about to around and , respectively, as increases from to , and drop to less than when . Similarly, there is also a notable spike at for QCD relative correction, as shown in the lower panel of Fig.7. The QCD relative correction is less than and thus can be neglected when , while it is expected to increase to about as increases to . When , the QCD relative correction decreases slowly down to about as increases to .

[GeV] | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|
[fb] | ||||||||||
[fb] | ||||||||||
[%] |
The LO, NLO QCD corrected angular distributions of the final-state charged Higgs boson and the corresponding QCD relative corrections for associated production at a collider for and and are depicted in Fig.8, where denotes the scattering angle of with respect to the electron beam direction. Due to the conservation, the distribution of the scattering angle of with respect to the positron beam direction for is the same as the angular distribution of for . From this figure, we can see that the charged Higgs boson is mostly produced in transverse direction for both and . For , the QCD relative correction decreases rapidly from to nearly as the increment of from to , and is steady at around in the region of . It implies that the NLO QCD correction can be neglected in most of the phase space region except when . For , the LO differential cross section is suppressed by the NLO QCD correction in the whole phase space region. The corresponding QCD relative correction decreases from about to as increases from to . This QCD correction should be taken into consideration for precision study of the production at lepton colliders.


V Summary
Searching for exotic Higgs boson and studying its properties are important tasks at future lepton colliders. In this work, we study in detail the associated production at future electron-positron colliders within the framework of the Type-I THDM. We calculate the process at the LO, and investigate the dependence of the production cross section on the THDM parameters ( and ) and the colliding energy. The numerical results show that the cross section is very sensitive to the charge Higgs mass in the vicinity of at a collider, and decreases consistently with the increment of in the low region. The existence of a peak in the colliding energy distribution of the cross section is explained by the resonance effect induced by loop integrals. This resonance occurs only above the threshold of , and the peak position moves towards low colliding energy as the increment of . We also calculate the two-loop NLO QCD corrections to , and provide some numerical results for the NLO QCD corrected integrated cross section and the angular distribution of the final-state charged Higgs boson. For and , the QCD relative correction varies smoothly in the range of as the increment of from to , except in the vicinity of . The QCD relative correction is sensitive to the charged Higgs mass and strongly depends on the final-state phase space. For and , the QCD relative correction to the production at a collider increases from about to as the scattering angle of increases from to . Compared to hadron colliders, the measurement precision of Higgs associated production at future high-energy electron-positron colliders is much higher. The expected experimental error of Higgs production in association with a weak gauge boson at high-energy electron-positron colliders is less than through the recoil mass of the associated vector boson. For example, the measurement precision of production at FCC-ee and CEPC can reach about and , respectively Peskin ; Bicer et al. (2014); Ruan (2016). Due to the high -tagging efficiency at colliders, the measurement precision of production at a high-energy collider can be also less than . Thus, the two-loop QCD correction should be taken into consideration in precision study of the associated production at future lepton colliders.
VI Acknowledgements
This work is supported in part by the National Natural Science Foundation of China (Grants No. 11775211 and No. 11535002) and the CAS Center for Excellence in Particle Physics (CCEPP).
References
- Aad et al. (2012) G. Aad et al. (ATLAS), Phys. Lett. B716, 1 (2012).
- Chatrchyan et al. (2012) S. Chatrchyan et al. (CMS), Phys. Lett. B716, 30 (2012).
- Lee (1973) T. D. Lee, Phys. Rev. D8, 1226 (1973).
- Gunion et al. (2000) J. F. Gunion, H. E. Haber, G. L. Kane, and S. Dawson, Front. Phys. 80, 1 (2000).
- Gunion and Haber (2003) J. F. Gunion and H. E. Haber, Phys. Rev. D67, 075019 (2003).
- Branco et al. (2012) G. C. Branco, P. M. Ferreira, L. Lavoura, M. N. Rebelo, M. Sher, and J. P. Silva, Phys. Rept. 516, 1 (2012).
- Turok and Zadrozny (1991) N. Turok and J. Zadrozny, Nucl. Phys. B358, 471 (1991).
- Davies et al. (1994) A. T. Davies, C. D. froggatt, G. Jenkins, and R. G. Moorhouse, Phys. Lett. B336, 464 (1994).
- Fromme et al. (2006) L. Fromme, S. J. Huber, and M. Seniuch, JHEP 11, 038 (2006).
- Deshpande and Ma (1978) N. G. Deshpande and E. Ma, Phys. Rev. D18, 2574 (1978).
- Dolle and Su (2009) E. M. Dolle and S. Su, Phys. Rev. D80, 055012 (2009).
- Sirunyan et al. (2017) A. M. Sirunyan et al. (CMS), Phys. Rev. Lett. 119, 141802 (2017).
- Sirunyan et al. (2020) A. M. Sirunyan et al. (CMS), JHEP 01, 096 (2020).
- Aaboud et al. (2018a) M. Aaboud et al. (ATLAS), JHEP 09, 139 (2018a).
- Aaboud et al. (2018b) M. Aaboud et al. (ATLAS), JHEP 11, 085 (2018b).
- Santos and Barroso (1997) R. Santos and A. Barroso, Phys. Rev. D56, 5366 (1997).
- Degrande (2015) C. Degrande, Comput. Phys. Commun. 197, 239 (2015).
- Krause et al. (2016) M. Krause, R. Lorenz, M. Muhlleitner, R. Santos, and H. Ziesche, JHEP 09, 143 (2016).
- Denner et al. (2016) A. Denner, L. Jenniches, J.-N. Lang, and C. Sturm, JHEP 09, 115 (2016).
- Altenkamp et al. (2017) L. Altenkamp, S. Dittmaier, and H. Rzehak, JHEP 09, 134 (2017).
- Arhrib and Moultaka (1999) A. Arhrib and G. Moultaka, Nucl. Phys. B558, 3 (1999).
- Heinemeyer and Schappacher (2016) S. Heinemeyer and C. Schappacher, Eur. Phys. J. C76, 535 (2016).
- (23) S. H. Zhu, hep-ph/9901221.
- Arhrib et al. (2000) A. Arhrib, M. Capdequi Peyranere, W. Hollik, and G. Moultaka, Nucl. Phys. B581, 34 (2000), [Erratum: Nucl. Phys.2004,400(2004)].
- Kanemura (2000a) S. Kanemura, Eur. Phys. J. C17, 473 (2000a).
- Drees and Roy (1991) M. Drees and D. P. Roy, Phys. Lett. B269, 155 (1991).
- Roy (1999) D. P. Roy, Phys. Lett. B459, 607 (1999).
- Capdequi Peyranere et al. (1991) M. Capdequi Peyranere, H. E. Haber, and P. Irulegui, Phys. Rev. D44, 191 (1991).
- Kanemura (2000b) S. Kanemura, Phys. Rev. D61, 095001 (2000b).
- Hernandez-Sanchez et al. (2004) J. Hernandez-Sanchez, M. A. Perez, G. Tavares-Velasco, and J. J. Toscano, Phys. Rev. D69, 095008 (2004).
- Arhrib et al. (2007) A. Arhrib, R. Benbrik, and M. Chabab, J. Phys. G34, 907 (2007).
- Farris et al. (2004) T. Farris, H. E. Logan, and S.-f. Su, Phys. Rev. D69, 035005 (2004).
- Kanemura et al. (2011) S. Kanemura, K. Yagyu, and K. Yanase, Phys. Rev. D83, 075018 (2011).
- Logan and Su (2002) H. E. Logan and S.-f. Su, Phys. Rev. D66, 035001 (2002).
- Logan and Su (2003) H. E. Logan and S.-f. Su, Phys. Rev. D67, 017703 (2003).
- Eriksson et al. (2010) D. Eriksson, J. Rathsman, and O. Stal, Comput. Phys. Commun. 181, 833 (2010).
- Grinstein et al. (2016) B. Grinstein, C. W. Murphy, and P. Uttayarat, JHEP 06, 070 (2016).
- Nie and Sher (1999) S. Nie and M. Sher, Phys. Lett. B449, 89 (1999).
- Akeroyd et al. (2000) A. G. Akeroyd, A. Arhrib, and E.-M. Naimi, Phys. Lett. B490, 119 (2000).
- Sirunyan et al. (2019) A. M. Sirunyan et al. (CMS), Eur. Phys. J. C79, 564 (2019).
- Hahn (2001) T. Hahn, Comput. Phys. Commun. 140, 418 (2001).
- Mertig et al. (1991) R. Mertig, M. Bohm, and A. Denner, Comput. Phys. Commun. 64, 345 (1991).
- Shtabovenko et al. (2016) V. Shtabovenko, R. Mertig, and F. Orellana, Comput. Phys. Commun. 207, 432 (2016).
- Maierhöfer et al. (2018) P. Maierhöfer, J. Usovitsch, and P. Uwer, Computer Physics Communications 230, 99 (2018).
- Laporta (2000) S. Laporta, Int. J. Mod. Phys. A15, 5087 (2000).
- Liu et al. (2018) X. Liu, Y.-Q. Ma, and C.-Y. Wang, Phys. Lett. B779, 353 (2018).
- Davydychev and Tausk (1993) A. I. Davydychev and J. B. Tausk, Nucl. Phys. B397, 123 (1993).
- Broadhurst (1999) D. J. Broadhurst, Eur. Phys. J. C8, 311 (1999).
- Kniehl et al. (2017) B. A. Kniehl, A. F. Pikelner, and O. L. Veretin, JHEP 08, 024 (2017).
- Ahnert et al. (2011) K. Ahnert, M. Mulansky, T. E. Simos, G. Psihoyios, C. Tsitouras, and Z. Anastassi, AIP Conference Proceedings 1389, 1586 (2011).
- Prausa (2017) M. Prausa, Comput. Phys. Commun. 219, 361 (2017).
- Bernreuther et al. (2005) W. Bernreuther, R. Bonciani, T. Gehrmann, R. Heinesch, T. Leineweber, P. Mastrolia, and E. Remiddi, Nucl. Phys. B706, 245 (2005).
- Tanabashi et al. (2018) M. Tanabashi et al. (Particle Data Group), Phys. Rev. D98, 030001 (2018).
- Behnke et al. (2013) T. Behnke, J. E. Brau, B. Foster, J. Fuster, M. Harrison, J. M. Paterson, M. Peskin, M. Stanitzki, N. Walker, and H. Yamamoto, “The International Linear Collider Technical Design Report - Volume 1: Executive Summary,” (2013).
- Abada et al. (2019) A. Abada et al. (FCC), Eur. Phys. J. ST 228, 261 (2019).
- (56) M. E. Peskin, arXiv:1207.2516.
- Bicer et al. (2014) M. Bicer et al. (TLEP Design Study Working Group), JHEP 01, 164 (2014).
- Ruan (2016) M. Ruan, Nucl. Part. Phys. Proc. 273-275, 857 (2016).