1 Introduction

The discovery of a Higgs boson at the CERN LHC [1, 2] has confirmed that the Higgs potential plays a crucial role in the electroweak symmetry breaking (EWSB). While the measured properties of this Higgs boson are so far compatible with the predictions of the Standard Model (SM) within the experimental and theoretical uncertainties, the structure of the Higgs sector and of its potential remain to be determined. Furthermore, in spite of the successes of the SM, it is clear that new, Beyond-the-Standard-Model (BSM) physics is needed to address deficiencies of the SM – such as for instance the lack of an explanation for the baryon asymmetry of the Universe.

In this context, a key quantity to investigate is the trilinear Higgs coupling \(\lambda _{hhh}\). This coupling determines the shape of the Higgs potential away from the electroweak (EW) minimum and in turn controls the nature and strength of the EW phase transition (EWPT). For instance, a strong first-order EWPT, which is a requirement [3] for the scenario of EW baryogenesis [4, 5], is typically associated with a sizeable deviation of \(\lambda _{hhh}\) from its SM prediction, as was discussed first in Refs. [6, 7] (see also Ref. [8] for a more recent example). Even beyond its crucial role in the context of the EWPT, \(\lambda _{hhh}\) provides a unique opportunity to find signs of BSM physics arising from extended Higgs sectors. In particular, the loop contributions in models with additional scalars can cause the trilinear Higgs coupling to deviate by up to several hundred percent from its prediction in the SM if there is a substantial mass splitting between the BSM mass scales. This was found first at the one-loop level for the case of a Two-Higgs-Doublet Model (THDM) in Refs. [9, 10] but is now known to occur for a wide range of BSM models. The genuine physical nature of these effects was confirmed in Refs. [11, 12], where next-to-leading order (NLO), i.e. two-loop, corrections to those large one-loop effects were investigated and found to obey the expected perturbative behaviour. Unlike what is the case for most couplings of the SM-like Higgs boson at 125 GeV, large deviations in \(\lambda _{hhh}\) are possible even in scenarios where all its couplings are very close to the SM values at tree level, such as in aligned scenarios [13]. Meanwhile, it has recently been shown in Ref. [14] that the experimental limits (discussed in further detail below) have become sufficiently strong to probe these potentially large loop effects, and thus the comparison of the predictions for \(\lambda _{hhh}\) with the latest experimental bounds constitutes a powerful new method for constraining the parameter space of BSM theories (probes of BSM parameter space using the di-Higgs production cross-section directly have also been discussed in e.g. Ref. [15]). In this context it is important to keep in mind that \(\lambda _{hhh}\) cannot be directly measured experimentally. The crucial experimental quantity where \(\lambda _{hhh}\) enters at leading order is the process of Higgs pair production. The computation of the trilinear Higgs coupling \(\lambda _{hhh}\) constitutes a necessary intermediate result for the prediction of di-Higgs boson production. In fact, for the case where \(\lambda _{hhh}\) receives large loop corrections, the additional contributions to the di-Higgs production process may be of sub-leading order [14]. More generally, computations of trilinear Higgs couplings are also important for investigating decays of BSM Higgs bosons and BSM decays of the SM-like Higgs boson.

The present experimental information on the trilinear Higgs coupling \(\lambda _{hhh}\) is by far not as precise as what has been achieved for other couplings of the Higgs boson [16, 17]. Indeed, the current best limits on \(\lambda _{hhh}\) were obtained by the ATLAS collaboration using a combination of data from searches for (non-resonant) di-Higgs production and from the experimental results for single-Higgs production processes; they bound the ratio \(\kappa _\lambda ,\) defined as

$$\begin{aligned} \kappa _\lambda \equiv \frac{\lambda _{hhh}}{(\lambda _{hhh}^\text {SM})^{(0)}}, \end{aligned}$$
(1)

to be within the range \(-0.4<\kappa _\lambda <6.3\) at the 95% confidence level (CL) [18, 19]. In Eq. (1), \((\lambda _{hhh}^\text {SM})^{(0)}\) denotes the tree-level prediction for the trilinear coupling in the SM. The CMS collaboration has obtained similar results [17], namely \(-1.24<\kappa _\lambda <6.49.\) The quoted limits were obtained under the assumption that besides a variation of \(\kappa _\lambda \) all other couplings entering the analyses are fixed to their SM values. The current experimental limits leave ample room for BSM deviations, which would so far remain unobserved, but could be accessed in the foreseeable future, given the expected prospects for probing \(\lambda _{hhh}\) at the LHC and future colliders – see Ref. [20] for a review. Specifically, at the high-luminosity upgrade of the LHC (the HL-LHC), the projection for \(\kappa _\lambda \) at 95% CL is \(0.1<\kappa _\lambda <2.3\) [21]. At a future \(e^+e^-\) linear collider and a 100-TeV hadron collider it is expected that \(\kappa _\lambda \) can be determined at the level of \({{{\mathcal {O}}}}(10\%)\) [20, 22,23,24,25,26]. It should be noted that these projections were obtained under the assumption that \(\kappa _\lambda = 1\) is realised in nature and may significantly change if the actual value of \(\kappa _\lambda \) is different. In particular, for an enhanced value of \(\kappa _\lambda \) the prospects for extracting \(\lambda _{hhh}\) from the process \(e^+e^- \rightarrow Zhh\) at a linear collider with about 500 GeV would improve, while as a consequence of destructive interference contributions the prospects at the HL-LHC would deteriorate, see e.g. Refs. [27,28,29].

As stressed above, the most direct probe of the trilinear Higgs coupling are searches for di-Higgs production, because this process involves \(\lambda _{hhh}\) already at the leading order (LO). Single-Higgs production involves contributions of the trilinear Higgs coupling starting at the next-to-leading order (NLO) and EW precision observables at the next-to-next-to-leading order (NNLO) – see for instance Refs. [30, 31]. Of course, a general analysis should not be restricted to the case where BSM contributions enter exclusively via the trilinear Higgs coupling. On the other hand, in scenarios where large loop corrections to \(\lambda _{hhh}\) constitute the leading contributions to di-Higgs production, an effective coupling approach where the dominant corrections are incorporated into \(\kappa _\lambda \) provides a convenient framework to efficiently constrain BSM models with available experimental results, as discussed e.g. in Refs. [14, 32]. We will discuss in this paper in more detail the applicability of experimental constraints set on \(\lambda _{hhh}\).

A number of computations of the trilinear Higgs couplings in BSM theories have been carried out in the literature. At one-loop order, corrections were first computed in the SM and the Minimal Supersymmetric Standard Model (MSSM) in Refs. [33,34,35] (see also Refs. [36, 37] for the case of the MSSM with complex parameters). One-loop calculations of \(\lambda _{hhh}\) have since also been performed in the Next-to-MSSM (NMSSM) in Ref. [38] as well as for various non-supersymmetric extensions of the SM: with singlets [32, 39,40,41], additional doublets [9, 10, 41,42,43,44,45], and triplets [46,47,48]. For some of these models, specific results for \(\lambda _{hhh}\) are available in the public programs H-COUP [49, 50] and BSMPT [51, 52]. At two-loop order, Refs. [11, 53] obtained the two-loop \({\mathcal {O}}(\alpha _t\alpha _s)\) and \({\mathcal {O}}(\alpha _t^2)\) corrections to \(\lambda _{hhh}\) in the SM. In supersymmetric theories, Refs. [54, 55] investigated the \({\mathcal {O}}(\alpha _t\alpha _s)\) corrections to \(\lambda _{hhh}\) in the MSSM and the NMSSM, respectively, and recently Ref. [56] extended the NMSSM calculation to include also \({\mathcal {O}}(\alpha _t^2)\) effects. Regarding non-supersymmetric models, the leading two-loop BSM contributions to \(\lambda _{hhh}\) (arising from BSM scalars and, potentially, top quarks) are known for the Inert Doublet Model (IDM) [11, 12, 53], THDMs [11, 12], O(N)-symmetric real-singlet extensions of the SM [12, 57], and for various models with classical scale invariance [57].

In this work, we present the Python package anyH3, which takes a big step forward in facilitating the prediction of \(\lambda _{hhh}\).Footnote 1anyH3 allows the analytic and/or numerical computation of the trilinear Higgs coupling for general renormalisable theories to full one-loop order. For user convenience, the model definitions needed in anyH3 in order to enable the application of generic results to specific theories can be provided in the form of the widely-employed UFO format [58, 59]. anyH3 also offers a high level of flexibility in the renormalisation schemes used in calculations – with pre-defined commands for standard scheme choices and the additional possibility for the user to define other choices of counterterms. Furthermore, the tool allows the user to modify the treatment of tadpole contributions (for recent discussions see e.g. Refs. [60,61,62,63], Appendix A of Ref. [64], and section 4 of Ref. [65]). anyH3 is part of the wider anyBSM framework, where developments for further (pseudo-)observables are foreseen in the future. An additional anyBSM feature that is already available is the module anyPerturbativeUnitarity, which allows efficient and reliable verifications of perturbative unitarity constraints (at leading order, and in the high-energy limit).

This paper is organised as follows: we start by discussing in Sect. 2 the main elements of our automated computation of \(\lambda _{hhh}\), as well as the interpretation of the obtained results. Next, we present in Sect. 3 the workflow of anyH3 before presenting a brief tutorial of the program in Sect. 4. In Sect. 5, we discuss the cross-checks that were performed for the various models that are installed along with anyH3. Finally, we present in Sect. 6 examples of applications of anyH3 with an emphasis on new results. We summarise our results in Sect. 7. A number of Appendices provide additional details on the program and the considered models: Appendix A contains our conventions for general renormalisable models and for the generic expressions included in anyH3 as well as the conventions or restrictions on UFO model files; Appendix B discusses the various ways of generating compatible UFO models for new BSM theories; Appendix C presents the different models discussed in this paper, including details on renormalisation prescriptions and the treatment of tadpole contributions; Appendix D lists modifications of the UFO standard that are applied by anyBSM internally; Appendix E explains the caching that is available in anyH3, while Appendix F describes the Python interface pyCollier to the Fortran library COLLIER employed for computing loop functions numerically.

2 A generic approach to the trilinear Higgs coupling

In the following we outline the steps for the calculation of the trilinear Higgs coupling where the SM-like Higgs boson h appears at each of the three external legs. A generalisation to trilinear Higgs couplings involving one or more BSM Higgs boson(s) is left for future work.

All calculations performed in the program are based on results for general renormalisable theories. These generic results were obtained by the following steps:

  • For the different types of contributions entering the calculation, as specified in Eq. (2) below, all possible Feynman diagram topologies were identified.

  • For each of these topologies, all possible insertions of generic fields (scalars, fermions, vector bosons, ghosts) were processed.

  • Each of these generic diagrams was calculated using a generic Lagrangian with the help of FeynArts [66,67,68] and FormCalc [69].

The resulting generic expressions were hard-coded into the Python program code to be applied to a specific model upon run-time.

The steps described above include the introduction of several conventions in how the generic Lagrangian and its resulting Feynman rules are written. For instance, all fermion-fermion-scalar operators, \(F_1 F_2 S_3,\) are written in terms of left- and right-handed projectors, \((c_L^{123}P_L +c_R^{123}P_R)F_1{\bar{F}}_2 S_3,\) where the explicit form of \(c_{L,R}^{123}\) (depending on the corresponding operator within the considered model) is yet unspecified. A detailed description of all used conventions is given in Appendix A. The generic results, which are expressed in terms of generic couplings, have been implemented into the Python code. Upon run-time of the program, the couplings of the specified model are mapped onto the generic Lagrangian allowing one to directly obtain results for all contributing Feynman diagrams. A similar approach is followed for example in SARAH [70,71,72,73,74] and TLDR [75].

Currently anyBSM contains generic results for the scalar three-point function (where in the present implementation the three external scalars are assumed to be identical), the scalar two-point function, the scalar one-point function, the vector-boson two-point function, and the vector-boson–scalar two-point function. An overview of all topologies can be found in Appendix A.2. Making use of these building blocks, the trilinear Higgs coupling \(\lambda _{hhh}\) is calculated by the sub-module anyH3 at the one-loop level,

$$\begin{aligned} \lambda _{hhh} ={}&-\hat{\Gamma }_{hhh}(p_1^2,p_2^2,p_3^2)\nonumber \\ ={}&\lambda ^{(0)}_{hhh} + \delta ^{(1)}_\text {genuine}\lambda _{hhh} + \delta ^{(1)}_\text {tadpoles}\lambda _{hhh}\nonumber \\&+ \delta ^{(1)}_\text {WFR}\lambda _{hhh} + \delta ^{(1)}_\text {CT}\lambda _{hhh}\, , \end{aligned}$$
(2)

where \({\hat{\Gamma }}_{hhh}\) is the renormalised Higgs-boson three-point function. The superscripts indicate the loop order, namely \(\lambda _{hhh}^{(0)}\) denotes the tree-level result for the trilinear Higgs coupling, while \(\delta ^{(1)}_{\text {genuine}}\lambda _{hhh},\) \(\delta ^{(1)}_{\text {tadpoles}}\lambda _{hhh},\) \(\delta ^{(1)}_{\text {WFR}}\lambda _{hhh},\) and \(\delta ^{(1)}_{\text {CT}}\lambda _{hhh}\) are one-loop contributions. Specifically \(\delta ^{(1)}_{\text {genuine}}\lambda _{hhh}\) denotes the genuine vertex corrections, while \(\delta ^{(1)}_{\text {tadpoles}}\lambda _{hhh}\) and \(\delta ^{(1)}_{\text {WFR}}\lambda _{hhh}\) correspond to contributions involving tadpole insertions and external-leg corrections, respectively. The last term of Eq. (2) denotes the counterterm contribution (see below), while the second-last term encodes the contribution from external leg corrections. As indicated in Eq. (2), anyH3 is able to handle arbitrary external momenta. For all the one-loop pieces appearing in Eq. (2), we work in anyH3 only with UV-finite parts, unless otherwise specified. Explicit checks of UV-finiteness can be performed in anyH3, but have been done separately.

We also note that anyH3 allows independent calculations of self-energies and tadpoles. The necessary generic results were derived following the same steps as described above.

2.1 Renormalisation

Renormalisation is a fundamental ingredient of loop calculations. Minimal subtraction schemes like \(\overline{\text {MS}}\) are arguably the easiest schemes to automate since their implementation basically boils down to setting the divergent parts of the appearing loop integrals to zero.

However, for many processes it is known that \(\overline{\text {MS}}\) schemes may result in undesirable features like artificially large loop corrections or gauge dependencies. Also in the specific context of the trilinear Higgs coupling it has been observed that a potentially large part of the one-loop corrections can be absorbed into the Higgs-boson mass, which appears at the tree level, for instance by renormalising it in the on-shell (OS) scheme [12, 34, 35].

For this reason, anyH3 allows the specification of different renormalisation schemes. All bosonic (i.e., of particles with spin zero or spin one) masses appearing in the tree-level expression for the trilinear Higgs coupling can optionally be renormalised in the OS scheme. Moreover, also the vacuum expectation value (\(\text {VEV}\)) entering at lowest order can be renormalised in the OS scheme (see Appendix C.1.2 for more details). In addition, anyH3 allows the user to define custom counterterms, which are then included in the calculation of the trilinear Higgs coupling. In summary, the counterterm contribution to \(\lambda _{hhh}\) reads

$$\begin{aligned} \delta ^{(1)}_{\text {CT}}\lambda _{hhh}= & {} {} \sum _i \frac{\partial }{\partial m_i^2}\lambda ^{(0)}_{hhh} \cdot \delta ^{(1)}_{\text {CT}}m_i^2 + \frac{\partial }{\partial v}\lambda ^{(0)}_{hhh} \cdot \delta ^{(1)}_{\text {CT}}v\nonumber \\{} & {} \quad + \delta ^{(1)}_{\text {custom-CT}}\lambda _{hhh}, \end{aligned}$$
(3)

where h is used to denote the SM-like Higgs boson, \(m_i\) denotes all scalar or vector boson masses appearing in \(\lambda _{hhh}^{(0)},\) and v is used to denote the electroweak \(\text {VEV}\). The notations \(\delta ^{(1)}_{\text {CT}}x\) denote the one-loop counterterms for the parameters x. If one of the masses \(m_i\) is chosen to be renormalised in the OS scheme, its counterterm is determined via

$$\begin{aligned} \delta ^{(1)}_{\text {CT}}m_i^2 = -{{\textrm{Re}}}\Sigma ^{(1)}_{ii}(p^2=m_i^2), \end{aligned}$$
(4)

where \(\Sigma _{ii}\) is the self-energy of the scalar/vector particle i (with mass \(m_i)\), which is defined according to the conventions of e.g. Refs. [75, 76].Footnote 2 For the determination of the electroweak \(\text {VEV}\)  counterterm, we refer to Appendix C.1.2. We note, finally, that if no default or user-defined schemes are provided for a certain parameter, then an \(\overline{\text {MS}}\) renormalisation is employed.

2.2 Tadpole contributions

Besides the renormalisation of the masses and the electroweak \(\text {VEV}\)  appearing at the tree level, also tadpole contributions need to be taken into account. Since the UFO standard does not provide a unified notation to store information about the minimisation of the Higgs potential,Footnote 3 we use as default setting the Fleischer–Jegerlehner tadpole scheme of Ref. [77] (using \({\overline{\text {MS}}} \) tadpole counterterms). As a consequence, all tadpole diagrams have to be calculated explicitly – see Appendix C.1.1 for a detailed discussion. This does not only include explicit tadpole contributions to \(\lambda _{hhh},\) denoted as \(\delta ^{(1)}_{\text {tadpoles}} \lambda _{hhh},\) but also to \(\delta ^{(1)}_{\text {CT}} \lambda _{hhh}\) and \(\delta ^{(1)}_{\text {WFR}} \lambda _{hhh}.\) It should be stressed that the treatment of the tadpole contributions can be adapted by the user. In particular, it is possible to avoid the explicit appearance of tadpole diagrams by an appropriate choice of \(\delta ^{(1)}_{\text {custom-CT}}\lambda _{hhh}\) using e.g. \(\text {OS}\) tadpole counterterms.Footnote 4

2.3 External leg corrections

anyH3 also includes external leg corrections to ensure the proper normalisation of the external scalars. These corrections are given by

$$\begin{aligned} \delta ^{(1)}\lambda _{hhh}^\text {WFR}{} & {} {=}{} \sum _i\left( \frac{1}{2}\Sigma _{hh}^\prime (p_i^2)\lambda ^{(0)}_{hhh} {+} \sum _{j,h_j\ne h}\frac{\Sigma _{h h_j}(p_i^2)}{p_i^2{-}m_{h_j}^2}\lambda ^{(0)}_{h_jhh} \right) \nonumber \\{} & {} \equiv {} \sum _i\left( \frac{1}{2}\delta ^{(1)}Z_h(p_i^2)\lambda ^{(0)}_{hhh}+ \sum _{j,h_j\ne h}\delta ^{(1)}Z_{hh_j}(p_i^2)\lambda ^{(0)}_{h_jhh} \right) ,\nonumber \\ \end{aligned}$$
(5)
Fig. 1
figure 1

Schematical workflow for the calculation of \(\lambda _{hhh} \)

where the prime indicates a derivative with respect to the external momentum squared. The second line of this equation serves to define the notations \(\delta ^{(1)}Z(p^2),\) which we will employ later in this paper. For the on-shell case, \(p_i^2 = m_h^2,\) the first term on the right-hand side yields the usual LSZ factor as it occurs for the case without mixing between different Higgs bosons, while the second term accounts for the contributions from possible scalar mixing effects on the external legs. As explained above, the self-energies appearing in Eq. (5) are meant to contain only the UV-finite contributions. It should be noted that anyH3 by default evaluates the external leg corrections at the same momenta as the vertex corrections. In order to implement different choices of the field renormalisations together with their appropriate wave function normalisation contributions, one can alternatively choose to turn off the automatic calculation of external-leg corrections and re-introduce the corresponding contributions in \(\delta ^{(1)}_{\text {custom-CT}}\lambda _{hhh}.\) This is in particular needed for the case where the result for the trilinear Higgs coupling obtained with anyH3 is meant to be incorporated into the prediction for the cross section for di-Higgs production. In the di-Higgs production process the trilinear Higgs coupling enters with two on-shell external legs, while the third leg is an off-shell internal line of the amplitude for di-Higgs production, see Sect. 6.5 and Appendix C.2 below for more details.

2.4 Interpretation of the result for \(\lambda _{hhh}\)

The trilinear Higgs coupling \(\lambda _{hhh}\) itself is not a physical observable. Its experimental determination from a physical process (typically di-Higgs production, but via higher-order contributions also single-Higgs production provides some sensitivity) relies on assumptions on the other couplings and particles involved in the processes. The current limits on \(\lambda _{hhh}\) from ATLAS [18] and CMS [17] were obtained under the assumption that all other Higgs couplings that are relevant for the respective processes have the SM values and that no other BSM particles contribute to these processes. Moreover, \(\lambda _{hhh}\) has been treated as a constant which does not depend on the inflowing momenta. When comparing the predictions obtained with anyH3 in the considered model to the experimental limits, the user of anyH3 should ensure that these assumptions are fulfilled sufficiently well.

Alternatively, the output of anyH3 can be used as input for other codes calculating di-Higgs production cross sections. In this case, the user can choose to include the momentum dependence and switch-off the external-leg corrections for the internal Higgs propagator (an explicit example for this case is discussed in Fig. 13 and also in Appendix C.2).

3 User- and program-flow

The main objective of this work is to provide one-loop corrections to the trilinear Higgs coupling in wide classes of BSM models. A typical work-flow of how this is organised is shown in Fig. 1. We distinguish two major sections for demonstrative purposes: (A) User input and (B) the actual program flow. User input is required in section (A). In addition, it is also possible to control each of the steps discussed in (B) by using the \(\texttt {anyBSM} \) library. The latter will be discussed in more detail below.

At the user level (A) several inputs are required:

  • (A.1) The model-specific information such as particle content and Feynman rules relies on the UFO standard. In order to obtain a UFO description of the model of interest one can use SARAH, FeynRules, or UFO files obtained with any other tool. For the latter two cases, we provide a converter, which is discussed in Appendix A.1, while in the case of SARAH the conventions for all relevant Lorentz structures match the anyBSM conventions. A detailed description of the UFO format can be found in Refs. [58, 59].

  • (A.2) In order to renormalise the relevant input parameters consistently, the tool needs to know which of the particles defined in the UFO model correspond to the SM particles. This information is specified in an auxiliary file called schemes.yml.

  • (A.3) Once the SM parameters and particles (along with their masses) are identified, one needs to specify in which renormalisation scheme they are given. This is also done in the file schemes.yml. An example specification of this file is shown in Sect. 4.4.

  • (A.4) Numerical values for all input parameters. Additionally, the UFO model may provide analytic relations between input parameters (so called “external” parameters) and e.g. Lagrangian parameters or mixing angles (so called “internal” parameters). The program automatically resolves these dependencies and writes all internal parameters in terms of external parameters. The numerical values are required for the numerical evaluation of the analytically obtained results for \(\lambda _{hhh}\) and are by default read from the UFO model. Moreover, one can change the default parameter values individually or all at once by specifying e.g. a SLHA [78, 79] input file (see Sect. 4.3 for examples).

It should be stressed that the relations between internal and external parameters given via the UFO model in (A.4) are essential for the renormalisation procedure. In particular Eq. (3) will be evaluated once all parameter dependencies have been applied. Thus, if any of the Higgs masses is not defined as an input parameter in the UFO  model, it cannot be renormalised in the \(\text {OS}\) scheme automatically. Instead, the corresponding counterterm contribution would need to be provided manually via the custom counterterm \(\delta ^{(1)}_{\text {custom-CT}}\lambda _{hhh}.\) It is, therefore, recommended to align the chosen parametrisation for input parameters in the UFO model along with the chosen renormalisation schemes.

Fig. 2
figure 2

Class and module structure of anyBSM. The ellipsis denotes additional observables which can be implemented in the future

The following steps are performed automatically using the information gathered before:

  • (B.1) The UFO model is loaded and several checks are performed:

    • Whether all relevant couplings are present (especially quartic scalar couplings which are sometimes excluded in UFO outputs).

    • Whether all relevant couplings are defined through the same Lorentz structures that are also used by anyBSM. Otherwise, one can use the model converter discussed in Appendix A.1.

  • (B.2) Definition of the SM-like particles and parameters based on the inputs made in (A.2) in the file schemes.yml. The program is also capable of finding the SM particles and parameters automatically based on their PDG codes and numerical (mass) values. This functionality is also used to cross-check the user-input in order to avoid erroneous configurations.

  • (B.3) and (B.4) All possible field-insertions into the generic diagrams are determined. The corresponding couplings and masses are inserted into the generic results. The calculation of any n-point function involved in the counterterm contributions follows the same procedure. Finally, the result for every n-point function is stored on-disk for caching/later use (see Appendix E).

  • (B.5) Collection of the individual results and construction of the expression for the renormalised \(\lambda _{hhh}\).

  • (B.6) Numerical or analytical evaluation. For diagrams with non-zero external momentum, the loop functions are evaluated using pyCollier (see Appendix F), which is a Python interface for COLLIER [80]. The analytical evaluation can be simplified/modified using SymPy [81].

We want to stress that this particular strategy for obtaining a prediction for \(\lambda _{hhh}\) in a given model has a number of ingredients in common with the calculation of many other observables. For this reason, the code anyH3 for calculating \(\lambda _{hhh}\) is embedded into a larger program called anyBSM. The program anyBSM provides many utilities capable of performing the steps described above to set up the calculation of a particular observable. Utilities of this kind are for instance the interface to UFO or the insertion of UFO particles and Feynman rules into generic Feynman diagrams. These ingredients are used in submodules – of which anyH3 is the first one that has been implemented – to define actual quantities to compute. In fact, anyH3 only takes care of (B.5) of the program points mentioned above while all other steps are taken care of in decoupled classes/modules of the program anyBSM.

This class and module structure is depicted in Fig. 2. The class anyModel encodes all information about the used model (Feynman rules, particles, interactions, etc.). This information is then inherited by the anyProcess class, which is used to calculate generic quantities like two- or three-point functions. This class makes use of various internal and external modules to derive the necessary diagrams and loop functions. The generic results of the anyProcess class are then used by the anyH3 class to specifically calculate the trilinear Higgs coupling. Future extensions of anyBSM featuring the calculation of new observables can be built on the basis of the anyProcess class (indicated by the ellipsis). In the end, the anyBSM class collects the classes for the different observables into a single object.

4 Program tutorial

In this section, we describe the basic features of anyBSM and anyH3. As explained above anyH3 is part of anyBSM, which provides a flexible Python framework for precision calculations (not only of \(\lambda _{hhh}\)). This section is not meant to be a detailed manual but instead is intended to give an overview of the overall functionality. A detailed description of all available methods and options is available in the online manual, which can be found at

https://anybsm.gitlab.io/anybsm.

4.1 Installation

The anyBSM source code is hosted at

https://gitlab.com/anybsm/anybsm.

Running the code requires at least Python version 3.5. The code is most easily used by installing the corresponding Python package by running

figure a

which will automatically download and install anyBSM as well as all necessary dependencies. One necessary requirement not handled automatically by pip is the presence of a Fortran compiler (required for the compilation of COLLIER for the pyCollier dependency) such as gfortran or CLANG which can be installed from the systems package repository. Upon the first run of anyBSM, the model repository will be downloaded and saved into a user-specified location.Footnote 5 The model repository is available online at

https://gitlab.com/anybsm/anybsm_models

and can be updated using the version control system git. The model repository will be expanded over time, and community contributions, in particular git merge requests, for tested models are welcome.

4.2 Basic syntax

anyBSM can either be integrated into Python scripts as a Python package or run directly from the command line. In addition, a Mathematica interface exists as well.

4.2.1 Python package mode

After starting a new Python session and importing anyBSM via

figure b

a model – here for instance the SM – can be initialised via

figure c

Alternatively, a path to a UFO model directory can also be given. As an overview, the dictionary anyBSM.built_in _models contains a list of all pre-installed UFO models and their installation directories. During the initialisation step, anyBSM will try to automatically identify the SM-like Higgs boson, for which \(\lambda _{hhh}\) is calculated, as well as all other SM particles.

After the model initialisation, \(\lambda _{hhh}\) can be calculated by running

figure d

which returns

figure e

Here, “total” denotes the total value for \(\lambda _{hhh}\) in GeV; “treelevel”, the tree-level value; “genuine”, the genuine one-loop contribution; “wfr”, the contribution from external-leg corrections (diagonal and off-diagonal); “tads”, the contribution from tadpole diagrams; “massren”, the contribution from mass renormalisation; “vevren”, the contribution from the renormalisation of the electroweak vev; and, “customren”, the contribution from a custom counterterm. It should be noted that by default \(\lambda _{hhh}\) is evaluated for vanishing external momenta. The option for using non-zero external momenta is described in Sect. 4.3.

4.2.2 Command line mode

As an alternative to using anyBSM as a Python package, it can also be called directly from the command line. A simple example is

figure f

which returns

figure g

An overview of the available options for the command line can be displayed by running

figure h

To view more options and details about a specific model one can also add the -h flag to the model name. For example

figure i

lists the particle content of the SM and the options for setting numerical values of all parameters (such as the top-quark mass) and their default values from the UFO model.

It should be noted that the command line tool provides access only to the basic functionalities of the anyBSM library, unlike the Python package mode and the Mathematica mode discussed below.

4.2.3 Mathematica mode

The Mathematica interface can be conveniently installed as follows:

figure j

which checks for all requirements and adds the anyBSM interface to Mathematica’s $Path variable. Afterwards, the interface can be used as follows:

figure k

The result lambda is a Mathematica Association object similar to the Python dictionary object obtained in Sect. 4.2.1 using the Python library. However, by default the analytical rather than numerical results are returned, after conversion to valid Mathematica expressions. A list of all available functions (such as for e.g. the calculation of self-energies and tadpoles) within Mathematica is stored in the variable $AnyFunctions. More information is given in the online documentation. Furthermore, comprehensive Mathematica notebooks that demonstrate the use of anyBSM ’s Mathematica mode are provided in the examples repository.

The Mathematica mode has access to the full functionalities of the Python backend (i.e. the anyBSM library). For the sake of clarity, we will restrict ourselves to a description of the Python package in the next sections.

4.3 Setting parameters

anyBSM uses the default parameters defined in the respective UFO model. To change e.g. the value of the top-quark mass in the example SM calculation discussed above, we can run

figure l

where in this example an effective (running) top-quark mass of 165 GeV is used. Alternatively, a LHA file can be used as input,

figure m

The program also defines a few additional UFO parameters in case they are not found in the UFO model. For instance, if the model does not define an external parameter named Qren (used for the renormalisation scale \(Q_{\text {ren.}})\), the code introduces it internally with a default value of \(\texttt {Qren}=172.5\,\text {{GeV}}.\) The full list of additionally introduced parameters is discussed in Appendix D as well as in the online documentation.

The external momenta entering the computation of \(\lambda _{hhh}\) can be specified by passing the momenta attribute to the lambdahhh function, e.g.

figure n

where Mh is the Higgs mass parameter defined in the SM UFO model file and is automatically replaced by its numerical value. By default, momenta = [0, 0, 0] is chosen.

4.4 Renormalisation

Information about the renormalisation is saved in the file schemes.yml in the model directory. A simple example file for the SM is

figure o

Here, the first SM_names block defines the names of various SM fields and parameters. The renormalization_schemes block can be used to define different renormalisation schemes. In the present example, the scheme OS is defined such that the mass of the field h as well as the \(\text {VEV}\)  counterterm are renormalised in the OS scheme. This scheme is set as default scheme via the default_scheme directive. In addition, the scheme MS is defined so that the mass of the field h as well as the \(\text {VEV}\)  counterterm are renormalised in the \(\overline{\text {MS}}\) scheme. It should be stressed that this does not mean that all inputs are converted from \(\text {OS}\) to \(\overline{\text {MS}}\) parameters but rather that the physical interpretation of these parameters is changed from \(\text {OS}\) to \(\overline{\text {MS}}\). However, for a consistent conversion of the parameters, all ingredients (i.e. two-point functions) are provided by the program. A proper conversion between the schemes will be demonstrated in Sect. 6.1.

If a non-default scheme should be used, this can e.g. be specified during the model initialisation:

figure p

As an alternative to using schemes predefined in schemes.yml, renormalisation schemes can also be generated interactively during the run time by using a new name that is not yet used in the schemes.yml file for the scheme directive during the model initialisation or by calling e.g. afterwards. The new scheme will then be saved into the schemes. yml file. It is also possible to change the renormalisation scheme, e.g. between two calls of SM.lambdahhh(), using the appropriate method:

figure r

If no schemes.yml file is present in the UFO model directory, it will be generated automatically upon the first creation of a renormalisation scheme which automatically searches for all SM-like parameters (particles) based on their numerical (mass) values (and PDG identifiers), cf. Sect. 3.

4.5 Evaluation modes and output formats

anyBSM supports three different evaluation modes:

  • abbreviations: all results are given in analytical form using the UFO coupling abbreviations (GC_1, GC_2, etc.);

  • analytical: all results are given in analytical form using the full analytical form for all couplings;

  • numerical: the numerical values for all parameters are inserted, and a numerical result is returned.

The evaluation mode can be set e.g. via

figure s

if using anyBSM as a Python package. The default evaluation mode is numerical.

A detailed breakdown of the results (including results for individual diagrams) in the form of a PDF document can be produced by using draw = True as an additional argument for the lambdahhh function or via the -t option when using the command line interface

figure t

The individual results listed along with the diagrams are represented in a way which depends on the chosen evaluation mode (e.g. numerical or analytical/using abbreviations). The resulting PDF file is saved to the current working directory as well as the model directory. In order to make use of this feature, LaTeX needs to be installed.

In addition to the Mathematica mode, analytical expressions can be exported from within a Python session to Mathematica with the help of SymPy

figure u

Note that anyBSM includes a caching system which automatically saves the analytic results into json files (into the cache directory located in the model directory). This leads to a significant speed-up of consecutive runs, see Appendix E.

4.6 Getting help

All Python classes and methods defined in anyBSM and anyH3 have meaningful doc-strings which can be issued by e.g.

figure v

or directly using existing class instances (such as help(SM.lambdahhh) in the examples above). In addition, the online documentation makes use of these doc-strings and provides a search functionality.

The usage of the command line tool anyBSM is returned by the command anyBSM -h. For a given model, one can obtain further help by issuing the command anyBSM<model name> -h from the command line. The Mathematica interface of anyBSM also provides documentation for all its functions by issuing ?<function name> such as e.g. ?lambdahhh. Furthermore, it provides a list of available functions stored in the variable $AnyFunctions.

In addition, the anyBSM examples repository provides basic and concrete examples for all three interfaces and for the generation of new model files.

5 Built-in models and cross-checks

The models currently distributed alongside anyH3 are

  • the Standard Model (SM);

  • the real-singlet extension of the SM (SSM);

  • the Two-Higgs-Doublet Model (THDM) – all four Yukawa types;

  • the Inert-Doublet Model (IDM);

  • the Next-to-Two-Higgs-Doublet Model (NTHDM) – i.e. the real-singlet extension of the THDM;

  • triplet extensions of the SM with either a real triplet with hypercharge \(Y=0\) or a complex triplet with \(Y=1.\) The two theories are denoted respectively TSM\(_{Y=0}\) and TSM\(_{Y=1}\);

  • the Georgi–Machacek model (a general version, as well as an aligned version);

  • a \(U(1)_{B-L}\) extension of the SM (BmLSM);

  • the Minimal Supersymmetric Standard Model (MSSM).

These models, and associated conventions, are described in more detail in Appendix C. We emphasise again that additional models can also be included by the user in a convenient and fast way, as described in Appendix B. In this section, we present details about a variety of analytical and numerical cross-checks we performed to validate anyH3.

5.1 Cross-checks using analytical computations

To cross-check the routines implemented in anyH3, we compared the analytical results to calculations performed using FeynArts and FormCalc. We found full agreement for the following pieces of the calculation performed in anyH3:

  • Higgs and Goldstone boson self-energies (and momentum derivatives thereof) including self-energies with two distinct external scalars as well as charged scalar self-energies;

  • Higgs tadpoles;

  • vector boson self-energies including mixing self-energies (e.g. \(\gamma -Z\) mixing);

  • genuine one-loop corrections to scalar three-point functions;

  • one-particle-reducible contributions to scalar three-point functions.

These checks have been performed in the SM, meaning that all contributions to the renormalised trilinear Higgs coupling arising in the SM have been cross-checked. Contributions that do not exist in the SM (e.g. scalar-mixing self-energies) have been cross-checked in the THDM.

Fig. 3
figure 3

\(\kappa _\lambda \) in the SSM as a function of the singlet mass (left) and one off-shell external momentum (right). The results match those of Ref. [40] Fig. 6 and 7 (upper-right), respectively

As an additional cross-check, we have verified the cancellation of ultraviolet divergences in the SM, the THDM, the SSM, the TSM\(_{Y=0}\), and the TSM\(_{Y=1}\). Moreover, we have found full agreement for the overall one-loop result for \(\lambda _{hhh}\) with independent calculations in the SM and the THDM, performed with FeynArts and FormCalc, as well as with the results for the TSM\(_{Y=1}\) from Ref. [48] (see Appendix C for a more detailed description of the models).

5.2 Numerical cross-checks

In addition to analytical cross-checks, we have performed a series of numerical cross-checks by reproducing results from the literature.

5.2.1 SSM

As a first check, we reproduced the SSM results for \(\kappa _\lambda \) derived in Ref. [40] (following the choice made in this reference and for the sake of comparison, we set the mixing between the CP-even states to zero). This reproduction is shown in Fig. 3 which is to be compared with Fig. 6 and 7 (upper-right) of Ref. [40]. In this figure, the momentum of two of the three external Higgs boson legs is always set on-shell \(\sqrt{p_1^2}=\sqrt{p_2^2}=m_{h_1}=125\,\text {{GeV}}.\) In the left plot of Fig. 3, the momentum of the third external Higgs leg is fixed at \(\sqrt{p_3^2}\equiv \sqrt{p^2}=251\,\text {{GeV}}\) and \(\kappa _\lambda \) is shown as a function of the singlet mass. In the right plot, the singlet mass is fixed to 200 GeV and the external momentum of the third Higgs boson leg is varied.

The behaviour of both plots reproduces the behaviour found in Ref. [40]. However, the exact numerical values in the left panel are shifted due to different treatments of the external-leg corrections. The different treatments of external momenta also lead to a slightly different peak structure in the right plot. However, at \(p^2=m_{h_1}^2\) the different treatments coincide. To show this equality we use the shift to \(\kappa _\lambda \) caused by the BSM sector, \(\delta _{hhh}^{(1)} = \kappa _\lambda - \nicefrac {\lambda _{hhh} ^{(1),\,\text {SM}}}{\lambda _{hhh} ^{(0),\,\text {SM}}},\) which was introduced in Eq. (25) of Ref. [40]. The external leg contribution to \(\delta _{hhh}^{(1)}\) in the two different treatments reads

(6)

where \(\text {B}_0(p^2)\equiv B_0(p^2,m_{h_2}^2, m_{h_2}^2)\) is the one-loop Passarino–Veltman two-point function [82, 83]. Thus, the two approaches yield the same external leg correction factors in the limit of \(p^2\rightarrow m_{h_1}^2.\) We made use of this relation to cross-check the full analytical result obtained with anyH3 with the result derived in Ref. [40] and found full agreement at \(p^2=m_{h_1}^2.\) For demonstrative purposes, we provide this cross-check using the Mathematica interface of anyBSM (cf. Sect. 4) as an example usage in the anyBSM examples repository.

5.2.2 TSM\(_{Y=1}\)

Fig. 4
figure 4

Reproduction of Fig. 11 of Ref. [46] showing results for \(\kappa _\lambda \) for the \(Y=1\) triplet model. Left: \(\kappa _\lambda \) contours are shown in the \((\lambda _4,\Delta m)\) parameter plane with \(\Delta m = m_{D^\pm } - m_{D^{\pm \pm }}\) and \(m_{D^{\pm \pm }} = 300\,\, \textrm{GeV}.\) Right: \(\kappa _\lambda \) contours are shown in the \((\lambda _4,\Delta m)\) parameter plane with \(\Delta m = m_{D^\pm } - m_{D^0}\) and \(m_{D^0} = 300\,\, \textrm{GeV}\)

As a further verification of anyH3, we reproduced results in the literature for the TSM\(_{Y=1}\) model [46]. Fig. 11 of Ref. [46] shows deviations of \(\lambda _{hhh}\) from the SM prediction in the plane of the coupling \(\lambda _4\) and the mass difference between the lightest and second-lightest BSM states (see Appendix C.7 for further details about the model). Our reproduction of this figure is shown in Fig. 4. In the left panel, the lightest BSM states are the doubly-charged Higgs bosons; in the right panel, the lightest BSM states are the two neutral BSM Higgs bosons. Overall, we observe a very good agreement between our results and the results presented in Ref. [46]. The remaining small differences can be traced back to different SM input parameters used in Ref. [46].

5.2.3 MSSM

Fig. 5
figure 5

Reproduction of Fig. 2 of Ref. [34] showing results for \({\mathcal {O}}(m_t^4)\) one-loop corrections \(\Delta \lambda _{hhh}\) to the trilinear Higgs coupling (normalised to the tree-level coupling) in the MSSM. Left: \(\Delta \lambda _{hhh}\) as a function of \(M_A\) for \(\tan \beta = 5\) (blue curve), \(\tan \beta = 10\) (orange curve), and \(\tan \beta = 30\) (green curve). Right: \(\Delta \lambda _{hhh}\) as a function of \(\tan \beta \) for \(M_A = 1\,\, \textrm{TeV}\)

As a cross-check of the MSSM implementation, we reproduced the results of Ref. [34]. In this work, the leading \({\mathcal {O}}(m_t^4)\) corrections to the trilinear Higgs coupling originating from scalar top quarks were calculated in the limit of vanishing electroweak gauge couplings. Setting the SUSY (and SM) parameters as in Ref. [34] (i.e., setting \(M_{{\tilde{Q}}} = M_{{\tilde{U}}} = 15\,\, \textrm{TeV},\) \(\mu = |A_t| = 1.5\,\, \textrm{TeV})\), we find very good agreement with their results (see our Fig. 5 in comparison to Fig. 2 of Ref. [34]).

5.2.4 Recovering the SM result in the decoupling limit

As an additional non-trivial cross-check of anyH3 (and also of the model files distributed alongside it), we have verified that the BSM contributions decouple if the masses of the BSM scalars are increased in a uniform way (see below for details), so that the SM result for \(\lambda _{hhh}\) is recovered in this limit.

Fig. 6
figure 6

For the shown models all masses have been chosen to be degenerate with the value \(M_{\textrm{BSM}}.\) Soft symmetry-breaking parameters are set to \(\sqrt{M_{\textrm{BSM}}-(250\,\text {{GeV}})^2}.\) Individually the relevant model parameters are fixed as follows. SSM: \(\alpha =0,\, \kappa _S=\kappa _{SH}=-800\,\text {{GeV}}\) and \(v_S=300\,\text {{GeV}}.\) IDM: \(\sqrt{M_{\textrm{BSM}}^2-\mu _2^2}=250\,\text {{GeV}}\) and \(m_{A,H,H^+}=M_{\textrm{BSM}}.\) THDM-II: \(\sqrt{M_{\textrm{BSM}}^2-M^2}=250\,\text {{GeV}},\)\(\tan \beta =2,\,\sin (\beta -\alpha )=1\) and \(m_{A,{h_2},{H^+}}=M_{\textrm{BSM}}.\) NTHDM-II: as in the THDM-II with \(v_S=M_{\textrm{BSM}},\) \(\alpha _1+\alpha _3=\beta -\pi /2,\) \(\alpha _2=\pi /2.\) TSM\(_{Y=0}\): \(\lambda _{T\Phi }=2.5\) and \(M_H^+=M_{\textrm{BSM}}.\) TSM\(_{Y=1}\): \(\lambda _4=2.5\) and \(m_{D^+}=m_{D^{++}}=M_{\textrm{BSM}}.\) GeorgiMachacek: \(\sqrt{M_{\textrm{BSM}}^2-M_\eta ^2}=250\,\text {{GeV}},\) \(M_5=M_3=M_{\textrm{BSM}}\) and \(\sin (H)=0.\) See Appendix C for details about the various models

This verification is shown in Fig. 6, where \(\kappa _\lambda \) is displayed as a function of the BSM mass scale \(M_{\text {BSM}}.\) All BSM masses in each model have been chosen to be degenerate with each other with the mass value \(M_{\text {BSM}}.\) The results for \(\kappa _\lambda \) are shown in the SSM (green), the IDM (light blue), the THDM-II (dark blue), the NTHDM-II (stars), the TSM\(_{Y=0}\) (red), the TSM\(_{Y=1}\) (orange), and the Georgi–Machacek model (brown). In all models we chose appropriate input parameters such that the lowest-order couplings of the SM-like Higgs boson to the other SM states are exactly as in the SM. This setting is referred to as aligned scenario in the following. Further details about the chosen parameters of the models are specified in the caption. As a reference, the SM result for \(\kappa _\lambda \) is indicated as a black line. It is clearly visible that for increasing \(M_{\text {BSM}}\) \(\kappa _\lambda \) quickly approaches the SM result for the chosen parameter settings. For \(M_{\text {BSM}}\gtrsim 1\,\, \textrm{TeV},\) the deviations from the one-loop SM result are below \(\sim 0.05.\) We note that, as will be seen in Sect. 6.4, values below the SM prediction are also possible in BSM models. Additionally, the details of the decoupling patterns in the different models strongly depend on the chosen parameters. Therefore, from the shown example no general conclusions can be drawn about how quickly decoupling occurs with increasing BSM mass scale in the various models.

6 Example applications

After having discussed cross-checks for validation, we present here a series of example applications. We first discuss estimates of the remaining theoretical uncertainties and then provide examples of results that go beyond existing studies in the literature.

6.1 Estimation of uncertainties in the computation of \(\lambda _{hhh}\)

When comparing predictions for an observable (or a pseudo-observable) with experimental measurements or limits, an important consideration to ensure the reliability of the comparison is to estimate the theoretical uncertainty associated with the obtained prediction. We devote this section to discussing different contributions to the theoretical uncertainty associated with computations of the trilinear Higgs coupling: on the one hand, uncertainties due to missing higher-order terms, and on the other hand, parametric uncertainties due to the limited precision with which quantities entering the calculation of \(\lambda _{hhh}\) are known experimentally.

6.1.1 Uncertainty from missing higher-order corrections in the SM

Focusing at first on the calculation of \(\lambda _{hhh}\) in the SM, we begin by investigating the possible size of missing higher-order contributions. While in the SM two-loop corrections of \({\mathcal {O}}(\alpha _t\alpha _s)\) and \({\mathcal {O}}(\alpha _t^2)\) are known [11, 12, 53], we refer here to all higher-order corrections that go beyond the full one-loop result that is obtained with anyH3 (see below for a discussion of the impact of the known two-loop corrections). Since calculations that are carried out at a given order in different renormalisation schemes differ by contributions that go beyond the calculated order, the comparison of different results of this kind for the same parameter point can be used as an estimate for the size of missing higher-order corrections – provided that the perturbative behaviour in the different schemes is of similar quality. We employ this method and compare the result obtained with anyH3 in the OS scheme with the one in the \(\overline{\text {MS}}\) scheme.

In the SM, we use for the quantities that enter the tree-level expression of \(\lambda _{hhh}\)  — i.e., the Higgs-boson mass, the W- and Z-boson masses, and the electromagnetic coupling \(\alpha _{\text {em}}\) (the latter three quantities are in turn used to compute the Higgs \(\text {VEV}\), see the discussion in Appendix C.1.2) – the following OS input values

$$\begin{aligned}{} & {} M_h=125.1\,\, \textrm{GeV},\quad M_W=80.379\,\, \textrm{GeV}\nonumber \\{} & {} M_Z=91.187\,\, \textrm{GeV}\nonumber \\{} & {} \alpha _{\text {em}}^{-1}(0)=137.035999679, \end{aligned}$$
(7)

where the notation \(M_i\) indicates the OS mass of particle i (we deviate here from the lower-case notation \(m_h\) employed for the Higgs mass in the rest of the paper in order to avoid ambiguities between OS and \(\overline{\text {MS}}\) masses). This yields for the tree-level prediction of the trilinear coupling a value of \(\lambda _{hhh}^{(0)}=187.3\,\, \textrm{GeV}.\) For the full one-loop predictions of \(\lambda _{hhh} \) in the on-shell scheme (where the tadpoles are renormalised in the OS scheme) we obtain for the two cases of vanishing external momenta and for the choice \(p_1^2=(200\,\, \textrm{GeV})^2\) and \(p_2^2=p_3^2=M_h^2\)

$$\begin{aligned}{} & {} \lambda _{hhh}^{(1),\,{\text { OS}}}(0,0,0)=176.2\,\, \textrm{GeV},\nonumber \\{} & {} \lambda _{hhh}^{(1),\,{\text { OS}}}((200\,\, \textrm{GeV})^2,M_h^2,M_h^2)=180.8\,\, \textrm{GeV}. \end{aligned}$$
(8)

In order to compare these values with those in the \(\overline{\text {MS}}\) scheme, we must first convert the OS input parameters \(M_h,\) \(M_W,\) \(M_Z,\) \(\alpha _{\text {em}}\) to the \(\overline{\text {MS}}\) scheme – this conversion (and all other scheme conversions in this paper) will be performed at one-loop order. Working at \(Q=172.5\,\, \textrm{GeV},\) and employing again an OS renormalisation of the tadpoles (see the discussion in Appendix C.1.1 for further details), we find after the one-loop conversion

$$\begin{aligned}{} & {} m_h^{\overline{\text {MS}}} = 121.4\,\, \textrm{GeV},\nonumber \\{} & {} m_W^{\overline{\text {MS}}} = 80.1\,\, \textrm{GeV},\nonumber \\{} & {} m_Z^{\overline{\text {MS}}} = 91.6\,\, \textrm{GeV},\nonumber \\{} & {} (\alpha _{\text {em}}^{\overline{\text {MS}}})^{-1}= 128.34. \end{aligned}$$
(9)

Using now these values as inputs for the calculation of \(\lambda _{hhh}\) in the \(\overline{\text {MS}}\) scheme, we obtain at \(Q=172.5\,\, \textrm{GeV}\) (again at full one-loop order, and with the same two choices of external momenta as above)

(10)

The difference of about \(0.3{-}0.4\,\, \textrm{GeV}\) between the results obtained in the OS and the \(\overline{\text {MS}}\) schemes constitutes a first estimate of a part of the unknown higher-order corrections; in relative size, the obtained shifts correspond to a difference of less than 0.2%. We note that if we had chosen to convert also the value of the squared Higgs mass used for the external momenta, the result for \(\lambda _{hhh}^{(1),{\overline{\text {MS}}}}((200\,\, \textrm{GeV})^2,(m_h^{\overline{\text {MS}}})^2,\)\((m_h^{\overline{\text {MS}}})^2)\) would have decreased by \(0.2\,\, \textrm{GeV},\) giving rise to only a slight change of our uncertainty estimate.

Concerning the interpretation of the uncertainty estimates obtained so far, it should be emphasised that the scheme comparison done above in fact does not capture corrections to \(\lambda _{hhh}\) involving the strong coupling \(\alpha _s,\) because the performed scheme conversions of the quantities \(M_h,\) \(M_W,\) \(M_Z,\) \(\alpha _{\text {em}}\) do not involve this coupling at one-loop order. Such effects can on the other hand be estimated by converting the input value used for the top-quark mass (in contrast to the quantities entering the prediction of the trilinear Higgs coupling at tree-level, converting the top-quark mass entering at the one-loop level from the OS to the \(\overline{\text {MS}}\) scheme directly gives rise to a two-loop effect in the prediction for \(\lambda _{hhh}\)). Starting from the OS value of \(M_t=172.5\,\, \textrm{GeV},\) a conversion including the leading \({{{\mathcal {O}}}}(\alpha _s)\) and \({{{\mathcal {O}}}}(\alpha _t)\) contributions to the top-quark self-energy (involving the strong gauge and top Yukawa couplings) yields an \(\overline{\text {MS}}\) value of \(m_t^{\overline{\text {MS}}} (Q=172.5\,\, \textrm{GeV})=166.3\,\, \textrm{GeV}.\) Using this \(\overline{\text {MS}}\) value in the computation of \(\lambda _{hhh}\), we obtain

(11)

which corresponds to a total deviation of about \(2\,\, \textrm{GeV}\) compared to the OS results above where \(M_t\) was used as input value. For the sake of comparison, we note that employing the OS computation of \(\lambda _{hhh}\) using \(M_h,\) \(M_W,\) \(M_Z,\) and \(\alpha _{\text {em}}^{-1}(0)\) from Eq. (7) but with the \(\overline{\text {MS}}\) value of \(m_t,\) we obtain

$$\begin{aligned}{} & {} \lambda _{hhh}^{(1),\text { OS}}(0,0,0)\big |^{m_t^{\overline{\text {MS}}}}=178.6\,\, \textrm{GeV},\nonumber \\{} & {} \lambda _{hhh}^{(1),\text { OS}}((200\text { GeV})^2,M_h^2,M_h^2)\big |^{m_t^{\overline{\text {MS}}}}=182.8\,\, \textrm{GeV}. \end{aligned}$$
(12)

As explained above, the dominant two-loop \({{{\mathcal {O}}}}(\alpha _t\alpha _s)\) and \({{{\mathcal {O}}}}(\alpha _t^2)\) corrections to \(\lambda _{hhh}\) are in fact known [11, 12, 53], and amount to about \(+3\,\, \textrm{GeV}\) (specifically, the \({{{\mathcal {O}}}}(\alpha _t\alpha _s)\) corrections amount to \(+4.1\) GeV, while those of \({{{\mathcal {O}}}}(\alpha _t^2)\) amount to \(-1.1\) GeV). Thus, our above estimates of higher-order corrections that are not included in the computation of anyH3 turn out to be close to the actual size of the known higher-order corrections. While for the case of the SM those two-loop corrections could be incorporated into the anyH3 prediction, in many other models for which anyH3 can be employed the corresponding corrections are not fully known. For reasons of uniformity we also restrict the SM prediction for \(\lambda _{hhh}\) in anyH3 to the full one-loop level. An extension of the code providing the incorporation of higher-order contributions is left for future work.

6.1.2 Parametric uncertainties

Another source of theoretical uncertainty in the prediction of \(\lambda _{hhh}\) arises from the experimental errors of the input parameters. In order to investigate the impact of these parametric uncertainties, we take into account the \(1 \, \sigma \) ranges of the experimental input parameters as given in Ref. [84],

$$\begin{aligned}{} & {} \Delta M_h^{\text {exp.}}=\pm 0.17\,\, \textrm{GeV},\nonumber \\{} & {} \Delta M_W^{\text {exp.}}=\pm 0.012\,\, \textrm{GeV},\nonumber \\{} & {} \Delta M_Z^{\text {exp.}}=\pm 0.0021\,\, \textrm{GeV},\nonumber \\{} & {} \Delta (\alpha ^{\text {exp.}}_{\text {em}}(0))^{-1}=\pm 2.1\times 10^{-8}, \end{aligned}$$
(13)

while for the top-quark mass we use

$$\begin{aligned} \Delta M_t=\pm 1 \,\, \textrm{GeV}. \end{aligned}$$
(14)
Table 1 Theoretical uncertainties in the calculation of \(\lambda _{hhh}\) arising from the experimental errors of the input parameters

It should be noted that the variation of \(\pm 1 \,\, \textrm{GeV}\) for \(M_t\) is indicated for illustration. The parametric uncertainty of the top-quark mass receives a contribution both from the experimental error of the measured mass parameter of \(\pm 0.30\,\, \textrm{GeV}\) at the \(1\,\sigma \) level [84] and from the systematic uncertainty arising from relating the measured quantity to a theoretically well-defined top-quark mass. The parametric uncertainties that are induced by the masses of the other quarks and the leptons are negligible. We furthermore note that we do not consider a parametric uncertainty from the strong gauge coupling because it does not enter the expression of \(\lambda _{hhh}\) at the one-loop level.

Varying each of the indicated experimental errors independently, we find the theoretical uncertainties induced in \(\lambda _{hhh}\) shown in Table 1. As expected, the largest effect on \(\lambda _{hhh}\), with an induced uncertainty of \(\pm 0.5\,\, \textrm{GeV},\) originates from the experimental error of the mass of the detected Higgs boson. Indeed the Higgs-boson mass enters the prediction for \(\lambda _{hhh}\) already at the tree level, and while it is already known to a high level of accuracy its experimental error is still larger – by more than an order of magnitude – than the experimental errors of \(M_W\) and \(M_Z\) (and much larger than the parametric uncertainty associated with \(\alpha _{\text {em}})\). The theoretical uncertainties that are induced by the gauge-boson masses have only effects at the level of some tens of MeV or less. On the other hand, the experimental uncertainty of the top-quark mass has a stronger impact on \(\lambda _{hhh}\) even though it only enters at the one-loop level. It should be noted that if in the future BSM parameters are measured, the parametric uncertainties in the prediction for \(\lambda _{hhh}\) induced by their experimental errors should also be taken into account.

6.1.3 Uncertainty from missing higher-order corrections in a BSM model: the example of the IDM

When considering the computation of \(\lambda _{hhh}\) in BSM theories, BSM parameters can enter the tree-level expressions. This is not the case in aligned scenarios, like the IDM, where the tree-level prediction for the trilinear coupling is the same as in the SM and only the mass of the detected Higgs boson and the associated vacuum expectation value enter the lowest-order prediction. It should be noted that also in this case a comparison between the OS and \(\overline{\text {MS}}\) results for \(\lambda _{hhh}\) computed at one-loop order with anyH3 with a one-loop conversion of \(M_h\) and v would not be sensitive to the type of contributions that can give rise to the largest effects at the two-loop level. For the specific case of the IDM one can infer from simple arguments of dimensional analysis that the leading two-loop corrections to \(\lambda _{hhh}\) are of \({\mathcal {O}}(g_{h\Phi \Phi }^5/M_\Phi ^4)\) and \({\mathcal {O}}(\lambda _2g_{h\Phi \Phi }^3/M_\Phi ^2)\) (which were computed in Ref. [11]), where \(\Phi \) denotes either of the BSM scalars of the IDM, \(g_{h\Phi \Phi }\) is a coupling between the Higgs boson at 125 GeV and two BSM scalars, and \(\lambda _2\) is the Lagrangian self-coupling of the inert doublet (c.f. Appendix C.4 for more details). These types of contributions are not generated by a one-loop conversion of \(M_h\) or v.

Instead, the size of these contributions can be estimated via a one-loop conversion of the BSM scalar masses, which affect the size of the dominant one-loop corrections to \(\lambda _{hhh}\) of \({\mathcal {O}}(g_{h\Phi \Phi }^3/M_\Phi ^2).\) In the following, we investigate for four different example benchmark points – labelled BP1, BP2, BP3, and BP4 (defined in Table 2) – the potential size of these leading two-loop effects. We choose BP1 and BP2 with small splittings between the BSM scalar masses and the BSM mass parameter \(\mu _2,\) so that the couplings \(g_{h\Phi \Phi }\) (which are proportional to the difference \(M_\Phi ^2-\mu _2^2)\) remain small, while we choose larger splittings for BP3 and BP4. Additionally, we set \(\lambda _2\) to zero in BP1 and BP3, in order to investigate only terms of the form \({\mathcal {O}}(g_{h\Phi \Phi }^5/M_\Phi ^4),\) while for BP2 and BP4 we set \(\lambda _2=2\) to also include effects of \({\mathcal {O}}(\lambda _2g_{h\Phi \Phi }^3/M_\Phi ^2).\) We present the results obtained with the code anyH3 for the one-loop conversions of the scalar masses and for \(\lambda _{hhh}\) in Table 2 (note that for the computation of \((\lambda _{hhh}^{(1)})^{\text {OS}}\) and for the scheme conversion of the BSM scalar masses, the tadpole contributions are renormalised on-shell).

Table 2 One-loop predictions for \(\lambda _{hhh}\) in the IDM for different example scenarios in the OS and the \(\overline{\text {MS}}\) scheme, as well as the relative difference \(\Delta .\) For the conversion of masses from the OS to the \(\overline{\text {MS}}\) scheme, as well as the \(\overline{\text {MS}}\) calculation of \(\lambda _{hhh}\), the renormalisation scale is chosen to be \(Q=300\,\, \textrm{GeV}.\) The values of the \(\overline{\text {MS}}\) masses are also given. For all four benchmark scenarios we set the OS masses to \(M_H=400\,\, \textrm{GeV},\) \(M_A=410\,\, \textrm{GeV},\) \(M_{H^\pm }=415\,\, \textrm{GeV},\) while the other free IDM parameters are given in the “Inputs” columns

As could be expected, we find that the OS and \(\overline{\text {MS}}\) results are in very good agreement – differing only by 1.4% and 2.5% for the two choices of \(\lambda _2\) – for the scenarios with small mass splittings (BP1 and BP2). For BP3 and BP4 featuring larger splittings, the discrepancy between the two results increases to about 5–6%. This confirms the known fact that the inclusion of two-loop corrections to \(\lambda _{hhh}\) is increasingly important for parameter regions with larger splittings between the different BSM masses. Finally, we observe that the relative size of the \({\mathcal {O}}(\lambda _2g_{h\Phi \Phi }^3/M_\Phi ^2)\) pieces compared to the \({\mathcal {O}}(g_{h\Phi \Phi }^5/M_\Phi ^4)\) ones decreases for larger mass splittings, which simply follows from the lower power dependence on \(g_{h\Phi \Phi }\propto (M_\Phi ^2-\mu _2^2).\)

6.2 Comparison of renormalisation scheme choices for the TSM\(_{Y=0}\)

Fig. 7
figure 7

\(\kappa _\lambda \) in the \(Y=0\) triplet extension of the SM as a function of \(\lambda _{HT},\) comparing the OS (red curves) and \(\overline{\text {MS}}\) (blue curves) renormalisation schemes for the Higgs-boson mass and the EW VEV. For the \(\overline{\text {MS}}\) case, we consider different choices of the renormalisation scale, shown by the solid, dashed and dotted curves. The charged Higgs mass is set to \(M_{H^+}=1\text { TeV},\) while \(\lambda _T=1.5.\) Left: Results using the FJ scheme for the tadpoles. Right: Results using OS-renormalised tadpoles

Fig. 8
figure 8

\(\kappa _\lambda \) in the \(Y=0\) triplet extension of the SM. Left: \(\kappa _\lambda \) as a function of \(\lambda _{HT},\) comparing results employing the OS- and \(\overline{\text {MS}}\)-renormalised charged Higgs mass. The OS scheme is used for the Higgs mass, the EW VEV, and the tadpoles. Right: Results for \(\kappa _\lambda \) (calculated in the OS scheme) shown in the \((\lambda _{HT},M_{H^+})\) parameter plane

Choosing a suitable renormalisation scheme is a crucial step for the calculation of \(\kappa _\lambda \). We illustrate this in Figs. 7 and 8 for the \(Y=0\) triplet extension of the SM (see Appendix C.6 for details about the model and its implementation).

In Fig. 7, \(\kappa _\lambda \) is shown for different renormalisation schemes as a function of the quartic interaction between the SM Higgs doublet and the BSM triplet \((\lambda _{HT})\) with fixed \(M_{H^+} = 1\,\, \textrm{TeV}\) and \(\lambda _T = 1.5.\) In the left plot, the tadpoles are treated in the FJ prescription, and therefore enter both the calculation of \(\lambda _{hhh}\) and the parameter conversion explicitly, while in the right plot an OS renormalisation is used for the tadpoles. If the mass of the SM-like Higgs boson and the EW \(\text {VEV}\) are renormalised in the on-shell scheme (red curve, identical in both plots), the dependence of \(\kappa _\lambda \) on \(\lambda _{HT}\) is very small. This is expected since the BSM masses are chosen at the TeV scale implying that all BSM corrections should be small as a consequence of decoupling. If the mass of the SM-like Higgs boson and the EW \(\text {VEV}\) are renormalised in the \(\overline{\text {MS}}\) scheme (blue curves), the result for \(\kappa _\lambda \) depends strongly on the choice of the renormalisation scale, as well as on the chosen treatment of the tadpoles. For all curves, \(m_h^{\text {OS}} = 125.1 \,\, \textrm{GeV}\) and \(v^{\text {OS}} \simeq 250.7 \,\, \textrm{GeV}\) are used as input which are then converted in the first step to the \(\overline{\text {MS}}\) scheme. Then, these \(\overline{\text {MS}}\) quantities are used to calculate \(\kappa _\lambda \). For the considered scenario, the conversion can lead to very large shifts between the OS and \(\overline{\text {MS}}\) quantities if the renormalisation scale is not chosen appropriately. If for example \(\mu _R = m_t\) is chosen (solid blue curve), we encounter artificially large corrections to \(\kappa _\lambda \) for large positive \(\lambda _{HT}\) when employing FJ tadpoles. We note that this is due exclusively to the impact of the tadpoles on the \(\overline{\text {MS}}\) parameters obtained from OS inputs, because the tadpole contributions in the calculation of \(\lambda _{hhh}\) itself are the same independently of the employed scheme, as shown in Appendix C.1.1. Moreover, for this scale choice the \(\overline{\text {MS}}\) mass of the SM-like Higgs boson quickly becomes tachyonic for negative \(\lambda _{HT}\) in the case with FJ tadpoles, and for positive \(\lambda _{HT}\) in the case with OS tadpoles. Similar issues appear for the choice of \(\mu _R = M_{H^+}\) (blue dotted curve) for which the \(\overline{\text {MS}}\) mass of the SM-like Higgs boson becomes tachyonic for \(\lambda _{HT} \gtrsim 1\) \((\lambda _{HT}\lesssim -1/2)\) when using FJ (OS) tadpoles. For \(\mu _R = (m_t + M_{H^+})/2\) (blue dashed curve), however, the corrections are quite well-behaved, and a result close to the OS curve is obtained – although in the case of OS tadpoles, the Higgs mass once again becomes tachyonic for \(\lambda _{HT}\gtrsim 3.\) Overall, the choice of OS tadpoles leads to more moderate effects in \(\kappa _\lambda ,\) and appears to be (if implemented) a preferable option for calculations in which scheme conversions of parameters are performed.

In the left plot of Fig. 8, we next present results for \(\kappa _\lambda \) using OS inputs for the Higgs mass and EW VEV (and OS-renormalised tadpoles), but with the charged Higgs boson mass in the OS (red curve) and the \(\overline{\text {MS}}\) scheme (orange curve). As in Fig. 7, we compare here results for the same parameter point, defined by \(\lambda _T=1.5\) and an OS charged Higgs mass of \(M_{H^+}^{\text {OS}}=500\,\, \textrm{GeV}\) (which for the \(\overline{\text {MS}}\) curve is converted to the \(\overline{\text {MS}}\) scheme). Because \(M_{H^+}\) only enters the prediction for \(\lambda _{hhh}\) at the one-loop order, the difference between the red and orange curves is formally of two-loop order and, as was discussed for the IDM in Sect. 6.1.3, can serve as an estimate of the size of the unknown higher-order contributions to \(\lambda _{hhh}\). We can observe, as expected, that the results in the two schemes remain very close for small values of \(|\lambda _{HT}\lesssim 2|,\) and only differ for larger couplings. Finally, additional results for \(\kappa _\lambda \), calculated in the OS scheme, are shown the \((\lambda _{HT},M_{H^+})\) parameter plane (for fixed \(\lambda _T = 1.5)\) in the right panel of Fig. 8.Footnote 6 It is clearly visible that large corrections to \(\kappa _\lambda \) are obtained for low \(M_{H^+}\) and large \(|\lambda _{HT}|.\) The solid black contour lines indicate the parameter region that is excluded by the current LHC bounds on \(\kappa _\lambda \) [18],Footnote 7 while the dashed black contour lines show the region that will be probed with the projected sensitivity based on the full HL-LHC dataset [21].

Fig. 9
figure 9

In all shown models the mass of the lightest BSM state which is charged under the \(SU(2)_L\) gauge group is set to \(M_L=400\,\text {{GeV}}.\) For the different models the following parameter choices have been made: IDM: \(M_{H}=\mu _2=M_L.\) THDM-II: \(M=M_{H}=M_L.\) TSM\(_{Y=1}\): \(m_{D^{++}}=M_L.\) GeorgiMachacek: \(M_{h_2}=M_\eta =M_L.\) All other parameters are chosen as in Fig. 6. In particular, the other BSM masses are degenerate with a mass value of \(M_{\text {BSM}}\)

6.3 Comparison of BSM effects arising from mass splittings

In Sect. 5.2.4 we showed for various models that the BSM contributions to \(\lambda _{hhh}\) vanish once all BSM states are decoupled simultaneously in an appropriate way. This behaviour is in accordance with the decoupling theorem [85] which states that, in the case of heavy new physics, all BSM contributions can be incorporated into coefficients \({\mathcal {C}}^{(n)}_i\) of higher-dimensional SM operators,

$$\begin{aligned} {\mathcal {L}}^{d>4} = \sum _{n=5}^\infty \frac{{\mathcal {C}}_i^{(n)}}{M_{\text {BSM}}^{n-4}}{{{\mathcal {O}}}}^{(n)}_i, \end{aligned}$$
(15)

such that all BSM effects vanish for \(M_{\text {BSM}}\rightarrow ~\infty .\) A crucial requirement for this decoupling behaviour is that the \({\mathcal {C}}_i^{(n)}\) are small and do not increase with \(M_{\text {BSM}}.\) Mass splittings between the BSM particles can modify this behaviour. This can happen for instance if some of the masses of the heavy BSM states \(\phi _{\text {BSM}}\) are mostly generated via the comparably small SM \(\text {VEV}\),

$$\begin{aligned}{} & {} \frac{1}{2}\left( \lambda _{XH}\Phi _\text {SM}^\dagger \Phi _\text {SM}+ M_L^2\right) \phi _\text {BSM}^2 \nonumber \\{} & {} \xrightarrow {\text {SSB}} \frac{1}{2}\left( \frac{1}{2}\lambda _{XH}v^2 + M_L^2\right) \phi _\text {BSM}^2= \frac{M_{\text {BSM}}^2}{2} \phi _\text {BSM}^2, \end{aligned}$$
(16)

where \(\phi ^2_{\text {BSM}}\) schematically stands for the quadratic term of some BSM scalar, and \(\Phi _{\text {SM}}\) is the SM-like doublet. Thus, \(\phi _{\text {BSM}}\) receives mass contributions both from the quartic coupling \(\lambda _{XH}\) and the mass parameter \(M_L.\) For the case of \(M_L\sim v\) a large scalar mass \(M_{\text {BSM}}~\gg ~v\) can be realised via \(\lambda _{XH}~\gg ~1.\) The quartic interaction of BSM scalars to SM-like Higgs bosons can lead to large contributions in this case,

$$\begin{aligned} c_{hh \phi _{\text {BSM}} \phi _{\text {BSM}}}\propto \lambda _{XH} = 2\frac{M_{\text {BSM}}^2-M_L^2}{v^2} . \end{aligned}$$
(17)

Another way of understanding the origin of such large contributions is related to the symmetry argument of the decoupling theorem i.e. that the decoupling of a heavy particle must not break any symmetries of the resulting effective theory (EFT). To demonstrate this we consider the states \(X_i\) of some irreducible \(SU(2)_L\) multiplet \(\varvec{X}\) with masses \(M_{X_{i\ne L}}\equiv M_{\text {BSM}}\) and \(M_{X_{L}}\equiv M_L\ll M_{\text {BSM}}.\) Taking the limit \(M_{\text {BSM}}\rightarrow \infty \) leads to an EFT which is not \(SU(2)_L\)-invariant anymore as \(X_L\) cannot be incorporated into a smaller \(SU(2)_L\) multiplet. As a consequence, portal couplings as in Eq. (16) between the SM- and the BSM-Higgs bosons become large for a large mass splitting. In fact, for all models implemented in anyH3 with additional states charged under \(SU(2)_L\) we found the couplings \(c_{hh X_i X_j}\) to behave as in Eq. (17), provided that appropriate parameterisations for the input masses are assumed. Note that the above discussion applies not only to the \(SU(2)_L\) gauge symmetry but more generally to global and gauge symmetries of the BSM theories.

Fig. 10
figure 10

Predictions for \(\kappa _\lambda \) (red, blue, and black lines) and for \(\kappa _t\) (green line) in the NTHDM and THDM, as a function of the NTHDM mixing angle of the \(\mathcal{C}\mathcal{P}\)-even scalar sector \(\alpha _2.\) Results for \(\kappa _\lambda \) are shown at tree level (dashed curves) and at one loop (solid lines). The masses of the BSM scalars are taken to be degenerate at 300 GeV, while the BSM mass scales – \({\tilde{\mu }},\) defined by \({\tilde{\mu }}^2=m_{12}^2/(\cos \beta \sin \beta ),\) for the NTHDM and M for the THDM – are chosen to be 100 GeV, and \(\tan \beta =2.\) For the singlet vev in the NTHDM we consider two scenarios: \(v_S=300\) GeV in blue and \(v_S=3\) TeV in red

It should be noted that there can be regions in parameter space where the splitting of the mass parameters \(M_{\text {BSM}}^2-M_L^2\) is relatively large compared to the electroweak scale while the model can still be described perturbatively. From the THDM it is known that such large couplings can be constrained by the current experimental bounds on \(\kappa _\lambda \) while being in agreement with all other experimental and theoretical constraints [14]. With the help of anyH3 one can easily go beyond the THDM and study the effect of couplings determined by Eq. (17) onto \(\kappa _\lambda \) in other SM extensions. To demonstrate this, we use the example of the different \(SU(2)_L\) extensions from Sect. 5.2.4 and fix one of the BSM scalar masses to \(M_L=400~\text {{GeV}}\) rather than having all masses degenerate at \(M_{\text {BSM}}.\) Figure 9 shows the resulting \(\kappa _\lambda \) prediction as a function of \(M_{\text {BSM}}\) for the IDM (light blue), the THDM (blue), the TSM\(_{Y=1}\) (orange) and the Georgi–Machacek model (brown). We want to emphasise again that all shown parameter points are chosen to be in the alignment limit, i.e. they have a tree-level prediction of \(\kappa _\lambda ^{(0)}=1.\) In agreement with Eq. (17), we observe in all models that \(\kappa _\lambda \approx \kappa _\lambda ^{\text {SM}}\) for \(M_{\text {BSM}}\approx M_L=400{GeV}.\) For increasing values of \(M_{\text {BSM}},\) corrections proportional to couplings of the form of Eq. (17) lead to a large increase of \(\kappa _\lambda \) so that it can become close to or even larger than the current experimental constraint (red horizontal line). The projection for the sensitivity on \(\kappa _\lambda \) at the HL-LHC (grey horizontal line) shows that it will be possible to probe mass splittings down to 150–200 GeV in the displayed examples. We explicitly verified, using the anyPerturbativeUnitarity module of anyBSM, that all models fulfil the tree-level perturbative unitarity constraint in the high-energy limit for all shown values of \(\kappa _\lambda \).

It is also important to stress that this discussion is not restricted to the \(SU(2)_L\) gauge symmetry of the SM but also applies to any other symmetry within or beyond the SM.Footnote 8

6.4 Phenomenological results for \(\lambda _{hhh}\) in the NTHDM

In this section we discuss an example of phenomenological results for \(\lambda _{hhh}\) in the NTHDM (other investigations of \(\lambda _{hhh}\) in the NTHDM have also been performed with BSMPT in Ref. [87]). The scalar sector of this model contains three \(\mathcal{C}\mathcal{P}\)-even scalars, which can mix, and in turn three mixing angles are required to diagonalise the \(3\times 3\) \(\mathcal{C}\mathcal{P}\)-even scalar mass matrix (see model definitions and details in Appendix C.5). It is of interest in this context to investigate the potential impact of mixing on the prediction for the trilinear Higgs coupling.

As a brief illustration of phenomenological studies made possible with anyBSM, we present in Fig. 10 results for \(\kappa _\lambda \) as a function of the second \(\mathcal{C}\mathcal{P}\)-even mixing angle \(\alpha _2.\) We choose a scenario of the NTHDM where \(h_2\) is identified with the detected Higgs boson at 125 GeV, so that the alignment limit is reached for \(\alpha _1+\alpha _3\rightarrow \beta -\pi /2\) and \(\alpha _2\rightarrow \pi /2.\) We consider in Fig. 10 scenarios of the NTHDM and THDM were the BSM scalars \((h_2,\) \(h_3,\) A\(H^\pm \) for the NTHDM; \(h_2,\) A\(H^\pm \) for the THDM) are mass-degenerate with a mass value of 300 GeV. We fix the BSM mass scales of both models \(({\tilde{\mu }}\) for the NTHDM and M for the THDM) to 100 GeV, resulting in a sizeable mass splitting giving rise to a significant contribution to \(\kappa _\lambda \). Additionally, we set \(\tan \beta =2\) and \(\alpha _1+\alpha _3=\beta -\pi /2,\) while for the singlet \(\text {VEV}\), \(v_S,\) we adopt two values: \(v_S=300\) GeV (blue curves) and \(v_S=3\) TeV (red curves). Regarding our choice of renormalisation scheme, we employ here an OS renormalisation of all scalar masses and of the Higgs \(\text {VEV}\), while the other parameters are renormalised in the \(\overline{\text {MS}}\) scheme. Note that this scenario is devised as a simple setting in which to demonstrate calculations that are made possible by anyBSM. The range of parameters shown in Fig. 10 is not allowed in its entirety, however, we do not explicitly indicate the exclusion limits since we are not aiming at a thorough phenomenological analysis here.

Fig. 11
figure 11

Upper panel: Momentum dependence of \(\kappa _\lambda \) in the THDM of type-I, with \(M=m_{h_2}=400\,\, \textrm{GeV},\) \(m_A=m_{H^\pm }=700\,\, \textrm{GeV},\) and \(\tan \beta =2.\) Lower panel: Same as upper panel but the imaginary part of \(\lambda _{hhh}\) is shown

The solid curves in Fig. 10 show the full one-loop results for \(\kappa _\lambda \), while the dashed lines correspond to the tree-level results. For \(\alpha _2\rightarrow \pi /2,\) we observe – as expected – that we recover the alignment limit, and for both possible values of \(v_S\) the tree-level and one-loop predictions for \(\kappa _\lambda \) converge to the results in the THDM, indicated by the black horizontal lines. In this limit, the additional singlet decouples entirely, and the dependence of \(\kappa _\lambda \) on \(v_S\) vanishes. A sizeable BSM contribution remains in this limit, yielding a value of \(\kappa _\lambda \sim 1.25,\) which arises from the corrections involving the THDM-like scalars \((h_2,\) A,  and \(H^\pm )\). On the other hand, away from the alignment limit, and for \(\alpha _2\lesssim \pi /4,\) the relative importance of the loop corrections to \(\lambda _{hhh}\) decreases significantly. It should be pointed out here, for completeness, that deviations of \(\kappa _t\) from the SM are already constrained by experimental data to be below \({\mathcal {O}}(20\%)\) (see for instance Ref. [16]). This implies that values of \(\alpha _2\lesssim \pi /4\) are already excluded in this scenario.

Furthermore, we can observe the interesting feature that the prediction for \(\kappa _\lambda \) becomes negative as \(\alpha _2\) decreases – i.e. as one departs from the alignment limit. At this point, it is however important to remark that the sign of \(\lambda _{hhh}\) is not a physical observable. A quantity that is of physical relevance is the relative sign between the trilinear Higgs coupling and other couplings of the Higgs boson, e.g. its coupling to top quarks. For this reason, we also present in Fig. 10 results (green line) for the coupling modifier of the top Yukawa interaction, which we denote \(\kappa _t,\) at tree level. We find that for the entire range of \(\alpha _2,\) \(\kappa _t^{(0)}\) remains positive, so that a change in the relative sign between the trilinear Higgs coupling and the top Yukawa does occur – this can in principle lead to a significant increase in the Higgs pair production cross-section, as the destructive interference between the box and triangle diagrams occurring in the SM is avoided in this case. We leave a more thorough investigation of scenarios with negative trilinear Higgs couplings in the NTHDM for future work.

Fig. 12
figure 12

Upper panel: Momentum dependence of \(\kappa _\lambda \) in the THDM of type-I, with \(M=m_{h_2}=600\,\, \textrm{GeV},\) \(m_A=m_{H^\pm }=1000\,\, \textrm{GeV},\) and \(\tan \beta =2.\) Lower panel: Same as upper panel but the imaginary part of \(\lambda _{hhh}\) is shown

6.5 Momentum-dependent effects in \(\lambda _{hhh}\)

By default, anyH3 evaluates the trilinear Higgs coupling setting the momentum of all external legs to zero. The code, however, also allows the evaluation of \(\lambda _{hhh}\) for finite momenta (via the argument momenta of the lambdahhh function).

Fig. 13
figure 13

Upper panel: Momentum dependence of \(\kappa _\lambda \) evaluated in the TSM\(_{Y=1}\) with \(m_{D^{\pm \pm }}=400\,\, \textrm{GeV},\) \(m_{D^\pm } = 500\,\, \textrm{GeV},\) and \(\lambda _4 = 4.\) Lower panel: Same as upper panel but the imaginary part of \(\lambda _{hhh}\) is shown

We demonstrate this for the THDM of type I (see Appendix C.3) and for the \(Y=1\) triplet extension of the SM (TSM\(_{Y=1}\), see Appendix C.7) in Figs. 11, 12 and 13. In all three scenarios, \(\kappa _\lambda =1\) at the tree level. For the THDM-I (Figs. 11 and 12), we present results for two scenarios, with \(M=m_{h_2}=400\) GeV, \(m_A=m_{H^\pm }=700\) GeV for the first and \(M=m_{h_2}=600\) GeV, \(m_A=m_{H^\pm }=1000\) GeV for the second, and with \(\tan \beta =2\) for both. Next, for the TSM\(_{Y=1}\) (Fig. 13), we set \(m_{D^{\pm \pm }}=400\,\, \textrm{GeV},\) \(m_{D^\pm } = 500\,\, \textrm{GeV},\) and \(\lambda _4 = 4.\) All three scenarios are chosen such that significant BSM effects occur in the trilinear Higgs coupling, and additionally the second THDM-I scenario is devised specifically to obtain a value of \(\kappa _\lambda \) larger than the current upper experimental bound of 6.3.

In the upper panels of Figs. 11, 12 and 13, we show \(\kappa _\lambda \) as a function of a varying external momentum scale \(\sqrt{p^2}.\) The orange dashed line denotes the \(\kappa _\lambda \) value if all squared external momenta \(p_i^{2}\) are set to zero. For the blue curve, two of the external momenta are set equal to the mass of the SM-like Higgs boson, while the momentum of the third leg is kept general and set to \(p^2.\) For this off-shell leg, we also do not include an external wave-function renormalisation, see also Sect. 2.3. This means that the result shown in blue corresponds precisely to the quantity that enters the evaluation of the triangle diagram contributing to di-Higgs production. Comparing the solid blue and the orange-dashed curves, we observe that, for low to intermediate ranges of \(\sqrt{p^2},\) the momentum effects are small in comparison to the overall size of the BSM effects, which shift \(\kappa _\lambda \) to about 2.95,  7.9,  and 3.3 respectively for the three figures. It is only for larger values of \(\sqrt{p^2}\) – namely \(\sqrt{p^2}\gtrsim 600-700\) GeV for the THDM-I scenarios and \(\sqrt{p^2} \gtrsim 1.2\,\, \textrm{TeV}\) for the TSM\(_{Y=1}\) – that the momentum effects become sizeable and cause a significant decrease in \(\kappa _\lambda \).

Finite external momenta can also induce imaginary parts for \(\lambda _{hhh}\) (for the calculation of \(\kappa _\lambda \), we take the real part of \(\lambda _{hhh}\)). These are shown in the lower panels of Figs. 11, 12 and 13. Several particle thresholds are visible (corresponding to what can be seen in the upper panels of the corresponding figures): e.g., the di-Higgs threshold around \(\sqrt{p^2}\sim 250\,\, \textrm{GeV},\) the di-top threshold around \(\sqrt{p^2}\sim 350\,\, \textrm{GeV},\) and for instance for the TSM\(_{Y=1}\) (Fig. 13), also the \(D^{\pm \pm }D^{\mp \mp }\) threshold \((\sqrt{p^2}\sim 800\,\, \textrm{GeV})\) and the \(D^{\pm }D^{\mp }\) threshold \((\sqrt{p^2}\sim 1000\,\, \textrm{GeV})\). Note that for the THDM-I (Figs. 11 and 12), there are no \(h_2h_2\) thresholds, because the \(\lambda _{h_1h_2h_2}\) coupling vanishes (due to the equality \(M=m_{h_2})\) and hence diagrams with internal \(\mathcal{C}\mathcal{P}\)-even scalars \(h_2\) do not contribute to \(\lambda _{hhh}.\) In Fig. 11, the \(AA/H^+H^-\) threshold is visible, as expected, at \(\sqrt{p^2}=1.4\,\, \textrm{TeV}.\)

Fig. 14
figure 14

Left: Mass of the SM-like Higgs boson computed with SARAH/SPheno [74, 89] at the one- and two-loop levels in the CMSSM as a function of the SUSY scale parameter \(m_0.\) Right: Same as the left panel, but results for the trilinear Higgs coupling are shown

The results shown in Figs. 11, 12 and 13 can directly be applied to di-Higgs boson production by treating \(\lambda _{hhh}\) as a momentum-dependent quantity entering the cross-section calculation. In this context, it is important to note that the di-Higgs invariant mass distribution typically peaks around \(\sqrt{p^2}\sim 400\,\, \textrm{GeV}\) (see e.g.  Ref. [88]) and then quickly falls off (by several orders of magnitude) as \(\sqrt{p^2}\) increases. This implies that the sizeable momentum dependence found for larger values of \(\sqrt{p^2}\) only has a small impact on the total di-Higgs boson production cross section. For the representative value of \(\sqrt{p^2}=400\,\, \textrm{GeV},\) i.e. around the peak of the di-Higgs differential cross-section, the momentum-dependence contributes positive shifts of 3.4% for the first THDM-I scenario, 1.0% for the second THDM-I case, and 3.6% for the TSM\(_{Y=1}\) scenario. Furthermore, we observe that, in all three considered scenarios, the imaginary part of \(\lambda _{hhh}\) remains minute until \(\sqrt{p^2}\gtrsim 350-400\,\, \textrm{GeV}\) (i.e. until the di-top threshold), and only reach sizeable values for \(\sqrt{p^2}\) well above \(400\,\, \textrm{GeV},\) and thus far from the peak of the di-Higgs invariant mass distribution. Consequently, we find that evaluating \(\lambda _{hhh}\) at zero external momenta is a good approximation of the full result for the scenarios considered here. We leave a more detailed investigation of the impact of non-zero momenta in di-Higgs production for future work. Moreover, we observe that for the points with the most sizeable BSM deviations in \(\kappa _\lambda \)  – such as the second THDM-I scenario – the relative magnitude of the momentum-dependent effects are the smallest, compared to the overall value of \(\kappa _\lambda \). Similarly, the magnitude of the imaginary part of \(\lambda _{hhh}\), which originates dominantly from the SM-like loop contributions, remains approximately the same in Figs. 11 and 12 (apart from the additional threshold at \(1.4\,\, \textrm{TeV}\) in Fig. 11). Thus, its relative size compared to its real part diminishes for scenarios with larger \(\kappa _\lambda .\)

Fig. 15
figure 15

One-loop corrections to trilinear scalar self-couplings in the SSM as a function of the tree-level singlet mass. Left: The SM-like Higgs self-coupling. Right: The scalar singlet coupling for a fixed renormalisation scale, \(\mu _R=m_t\) (dash-dotted), and a dynamically chosen scale, \(\mu _R=(m_t+m_s)/2\) (solid). Tree-level couplings are shown in orange (dashed). Lower panels: Relative difference between one-loop and tree-level predictions in percent

6.6 Use of anyH3 together with a spectrum generator: an example in the MSSM

anyBSM can easily be interfaced with spectrum generator tools. This is demonstrated in Fig. 14 for the constrained Minimal Supersymmetric Standard Model (CMSSM). For this example, we use SARAH/SPheno to generate the mass spectrum for different values of the SUSY scale parameter \(m_0\) (fixing the other BSM parameters via \(m_0 = m_{1/2} = - A_0,\) \(\tan \beta = 10,\) \(\text {sgn}(\mu )=1)\). The resulting predictions for the mass of the SM-like Higgs boson are shown in the left panel of Fig. 14: the solid curve shows the two-loop results, while the dotted curves corresponds to the one-loop result. The experimental value of \(\sim 125\,\, \textrm{GeV}\) is reached for \(m_0 \sim 4.4\,\, \textrm{TeV}\) in this example.

The loop-corrected mass spectrum computed with SARAH/SPheno is then passed to anyH3 using the SLHA interface (see Sect. 4.3). The resulting prediction for the trilinear Higgs coupling is shown in the right panel of Fig. 14. The orange curve shows the tree-level prediction, which is given in terms of the tree-level Higgs mass \((M_{h,\text {tree}}^2\simeq M_Z^2 \cos ^2{2\beta })\) divided by the electroweak VEV. The finite part of the genuine one-loop corrections to the trilinear Higgs coupling, represented by the green curve, is roughly of the same size as the tree-level prediction. Adding both contributions (together with the additional counterterm, external-leg, and tadpole contributions), the full one-loop result is obtained (blue curve). For \(m_0\sim 4.4\,\, \textrm{TeV},\) for which \(M_h^{(2)} \sim 125\,\, \textrm{GeV},\) we find \(\lambda _{hhh} \simeq 180\,\, \textrm{GeV}.\) This result is very close to the one-loop SM value of \(\sim 176\,\, \textrm{GeV}.\) For comparison, we also show in red the result of the effective lowest-order contribution \(3 M_h^2/v,\) where \(M_h\) incorporates the corrections to the mass of the SM-like Higgs boson up to the one-loop (red dotted curve) or the two-loop level (red solid curve). These results, which are quite close to the full one-loop result (within \(\sim 10\,\, \textrm{GeV})\), indicate that the bulk of the corrections to \(\lambda _{hhh}\) in the CMSSM enters via the loop-corrected prediction to the mass of the SM-like Higgs boson.

6.7 Non standard couplings

While Sect. 2 discusses specifically the calculation of the trilinear coupling of the SM-like Higgs boson, the program anyH3 is in general able to calculate any trilinear self-coupling \(\lambda _{h_i h_i h_i}\) where the three external Higgs bosons are the same. To demonstrate this feature we again consider the real singlet extension of the SM, the SSM, and compute both the SM-like Higgs coupling \(\lambda _{hhh} \) and the singlet coupling \(\lambda _{sss}.\) For simplicity, we set the singlet–doublet mixing angle to zero, \(\alpha =0,\) which leads to the tree-level expressions

$$\begin{aligned} \lambda _{hhh} ^{(0)}= & {} \frac{3 m_h^2}{v}, \quad \text {and} \end{aligned}$$
(18a)
$$\begin{aligned} \lambda _{sss}^{(0)}= & {} \frac{3 m_s^2}{v_S} - \kappa _S + \frac{3}{2}\frac{v^2}{v_S^2}\kappa _{SH}, \end{aligned}$$
(18b)

where we fix the mass of the SM-like Higgs boson to \(m_h=125\,\text {{GeV}}\) and allow the singlet mass \(m_s\) to be either larger or smaller than \(m_h\) (see Appendix C.2 for more details about the model). In this case the renormalisation of \(\lambda _{hhh} \) at the one-loop order is identical to the SM. For the trilinear singlet coupling, we choose to renormalise \(\kappa _S\) and \(\kappa _{SH}\) in the \(\overline{\text {MS}}\) and \(m_s\) in the OS scheme. The renormalisation of the singlet \(\text {VEV}\) \(v_S\) involves contributions from the singlet tadpole \(t_s.\) anyH3 provides the necessary ingredients to compute the one-loop prediction for \(\lambda _{sss}\) in the considered scenario.

Figure 15 shows the prediction for \(\lambda _{hhh}\) (left) and \(\lambda _{sss}\) (right) for \(\kappa _S=0,\) \(\kappa _{SH}=-400\,\text {{GeV}}\) and \(v_S=200\,\text {{GeV}}\) as a function of the singlet mass \(m_s\) in the interval \(80\,\text {GeV}\le m_s \le 200\,\text {GeV}.\) This corresponds to the scenario \(\lambda _{\Phi S}=1\) that has been studied in Ref. [40].Footnote 9 The corrections to \(\lambda _{hhh} \) reach up to 20% for small values of \(m_s\) in this scenario. For increasing \(m_s\) the prediction for \(\lambda _{hhh} \) approaches the one of the one-loop SM result. We find corrections to \(\lambda _{sss}\) between \(-50\) to \(+25\%\) in the considered singlet mass range. For non-zero soft-\({\mathbb {Z}}_2\)-breaking parameters the result for \(\lambda _{sss}\) depends on the chosen renormalisation scale (while \(\lambda _{hhh}\) does not depend on the renormalisation scale since the \(\overline{\text {MS}}\) parameters \(\kappa _S\) and \(\kappa _{SH}\) do not enter the prediction for \(\lambda _{hhh}\) at the tree-level). In order to demonstrate the dependence on the renormalisation scale we calculate \(\lambda _{sss}\) for the two options of using a fixed scale \(\mu _R=m_t\) and a dynamical scale \(\mu _R=(m_t+m_s)/2.\) We find that the difference in the \(\lambda _{sss}\) prediction for these two scale choices is at least about a factor 3 smaller than the overall size of the corrections for most of the considered \(m_s\) range. We have explicitly verified (besides the UV-finiteness of the result obtained with anyH3) that \(\lambda _{sss}\) is independent of the renormalisation scale for \(\kappa _{SH}=\kappa _S=0.\)

7 Conclusions

Obtaining information about the trilinear Higgs coupling \(\lambda _{hhh}\) is crucial for determining the shape of the Higgs potential and for gaining a better understanding of the nature of the electroweak phase transition. BSM contributions to \(\lambda _{hhh}\) can be large even for cases where all the couplings of the detected Higgs boson to gauge bosons and fermions are very close to the SM predictions. Thus, the comparison of the theoretical predictions for \(\lambda _{hhh}\) in different models – taking into account contributions at the quantum level – with the available experimental constraints on \(\lambda _{hhh}\) plays an important role for discriminating between the SM and extensions or alternatives of it.

It is therefore the main purpose of the public Python code anyH3, which we have presented in this paper, to provide precise predictions for \(\lambda _{hhh}\) in a wide variety of models. anyH3, which is part of the broader anyBSM framework, calculates the trilinear Higgs coupling in the SM and renormalisable BSM extensions of it at the one-loop level. For the model input, the code supports the widely used UFO format. This allows the user to easily extend the library of models shipped alongside the code. Already 14 models are provided in this library.

The code implements generic one-loop corrections which are mapped to the respective UFO model. For renormalisation, semi-automatic routines allow the user to easily implement different renormalisation schemes. Besides calculating the trilinear Higgs coupling, anyBSM also supports the calculation of other quantities like scalar and vector boson self-energies.

We have validated the results of anyH3 by explicit analytical cross-checks, various consistency checks (such as cancellation of UV divergences and decoupling of BSM contributions), and numerical cross-checks against known results in the literature. Besides comparing with known results, we have also presented new results for various models. In this context we have investigated different aspects like renormalisation scheme dependence, non-decoupling effects, momentum dependence, and negative trilinear Higgs couplings.

anyH3 can be used in the form of a Python module, called from the command line, or accessed via a Mathematica interface. All output quantities can either be evaluated analytically or numerically. To evaluate the required loop functions, anyH3 employs a link to the COLLIER library, which is available as the independent Python module pyCollier. Besides detailed examples (including scripts to reproduce all plots in this paper), we also provide an extensive online documentation.

Table 3 UFO Lorentz structures used in anyBSM for vertices involving scalars \(S_i,\) fermions \(F_i,\) vectors \(V_i^\mu ,\) and ghosts \(U_i.\) The four-momentum \(p_i^{\mu _j}=\texttt {'P(j,i)'}\) in the second and third columns is carried by the field with the label i in the first column. In this table, \(\gamma ^\mu \) denotes Gamma matrices, \(g^{\mu \nu }\) the metric tensor, and \(P_{L,R}\) the left-/right-handed projectors defined as \(P_{L,R}=(1\mp \gamma _5)/2\)

The code base of anyBSM is not restricted to the calculation of \(\lambda _{hhh}\), but can be easily extended to support the calculation of other observables like trilinear Higgs couplings with different external scalars (i.e. of the form \(\lambda _{ijk})\) or electroweak precision observables. We leave this for future work.