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

\smartqed\journalname
11institutetext: Universidad de Alcalá (UAH), E-28805 Alcalá de Henares, Madrid, Spain

Unwinding the rare Ω\rm\Omega sector: Fragmentation of
fully charmed baryons from HL-LHC to FCC

Francesco Giovanni Celiberto [Uncaptioned image]

Abstract

By adopting a hadron-structure-oriented approach, we present and discuss the release of the novel OMG3Q1.0 set of collinear fragmentation functions for fully charmed, rare Ω{\rm\Omega} baryons. Our methodology combines diquark-like proxy model inputs for both charm-quark and gluon channels, calculated at the initial energy scales, with a DGLAP evolution that ensures a consistent treatment of heavy-quark thresholds, following directly from the HF-NRevo scheme. We complement our work with a phenomenological study of NLL/NLO+ resummed Ω3c{\rm\Omega}_{3c} plus jet distributions using (sym)JETHAD at the HL-LHC and the future FCC. Unraveling the production mechanisms of rare, yet-unobserved hadrons, as provided by the OMG3Q1.0 functions, stands as a key asset for deepening our understanding of QCD at future high-energy hadron colliders.   Keywords:
Hadronic structure, Precision QCD,
Heavy flavor, Rare baryons, Omega sector,
Fragmentation, Resummation,
OMG3Q1.0 FF release

1 Hors d’œuvre

Heavy-flavored hadrons stand at the forefront of the search for New Physics. As natural gateways to potential Beyond the Standard Model (BSM) interactions, they serve as probes of unknown forces, thanks to their expected coupling to hypothetical particles beyond the Standard Model spectrum. Their unique sensitivity to rare processes and symmetry violations renders them prime observables in high-precision experiments and theoretical studies.

At the same time, the study of heavy flavors plays a central role in our quest to decode the microscopic structure of matter. It offers an unparalleled window into the dynamics of strong interactions, bridging perturbative and nonperturbative regimes of Quantum Chromodynamics (QCD). In this context, triply heavy baryons, such as the Ω3c{\rm\Omega}_{3c} and Ω3b{\rm\Omega}_{3b}, occupy a unique position in the hadronic spectrum predicted by QCD. Composed entirely of charm or bottom valence quarks, these particles are free from light-quark contributions, and thus provide an exceptional laboratory to probe the dynamics of color confinement in the heavy sector Bjorken (1985); Fleck and Richard (1989); Martynenko (2008); Martynenko and Trunin (2014); Karliner and Rosner (2014); Yoshida et al. (2015). As color-singlet bound states of three heavy quarks, their masses, decay widths, production mechanisms, and potential observability are the subject of intense theoretical scrutiny Ebert et al. (2002); Roberts and Pervin (2008); Chen and Wu (2011); Padmanath et al. (2014); Brown et al. (2014); Meinel (2012).

While numerous studies have primarily addressed baryons containing two heavy quarks, triply heavy systems like the Ω3c{\rm\Omega}_{3c} have also been investigated, often emerging as a natural extension within the same theoretical frameworks Karliner and Rosner (2014); Flynn et al. (2003); Shah et al. (2016). The Ω3c{\rm\Omega}_{3c} baryon, in particular, predicted to have a mass around 4.8GeV4.8\,\text{GeV} Shah and Rai (2017), is expected to be stable against strong decays and to undergo weak decays with a lifetime comparable to that of singly charmed baryons Bagan et al. (1994). Its experimental detection remains elusive, mainly due to the high production threshold and the challenges associated with reconstructing its multibody decay topologies in hadronic environments Chang et al. (2006); Gomshi Nobary and Sepahvand (2006). Fragmentation-based predictions for triply heavy baryons suggest extremely suppressed production rates at current colliders, with fragmentation probabilities in the [105÷107][10^{-5}\div 10^{-7}] range and total cross sections in the nb regime, significantly complicating the direct observation of Ω3c{\rm\Omega}_{3c} states Chen and Wu (2011); Gomshi Nobary and Sepahvand (2006). Although direct observation of triply heavy baryons remains challenging, their presence might be inferred in decay products of exotic multiquark states, as suggested by theoretical predictions Gershtein et al. (2000) and recent experimental signals of pentaquarks with open charm and strangeness Aaij et al. (2015a, 2019, 2021a).

Triply heavy baryons also play a crucial role in the broader program of heavy-hadron spectroscopy and the classification of bound states in QCD Brambilla et al. (2020); Esposito et al. (2017); Lebed et al. (2017); Faustov and Galkin (2022). Their clean internal structure, devoid of light-quark effects, makes them ideal candidates for testing potential models, effective theories, and predictions from lattice QCD Mathur and Padmanath (2019); Padmanath et al. (2014); Francis et al. (2019). Comparisons between triply heavy states and quarkonia, especially with regard to radial excitations and spin splittings, can help refine parameters such as the heavy-quark potential, effective coupling, and confining scale Eichten and Quigg (1994); Godfrey and Isgur (1985). In particular, analogies with fully heavy tetraquarks |QQ¯QQ¯|Q\bar{Q}Q\bar{Q}\rangle and pentaquarks |QQQQ¯Q|QQQ\bar{Q}Q\rangle Ali et al. (2017); Aaij et al. (2020); Karliner and Rosner (2020) suggest that Ω3c{\rm\Omega}_{3c} could be viewed as the baryonic anchor point of a larger spectrum of heavy-flavor objects, enabling unified treatments of their production through parton fragmentation. Such analogies also emerge in effective field theory descriptions, where nonrelativistic approximations lead to compact color configurations analogous to hydrogen-like or helium-like systems in atomic physics Pineda (2012); Celiberto et al. (2024a); Celiberto and Gatto (2025); Celiberto (2025a).

Triply heavy baryons composed of charm and bottom quarks are particularly valuable for disentangling the interplay of fundamental forces at different distance scales in nuclear systems. Due to their Coulomb-like structure, the ground-state radii of these baryons scale inversely with the product of the quark mass and the strong coupling, making them significantly more compact than ordinary baryons. As a result, meson-exchange and color-exchange dynamics differ substantially from those in conventional nuclear matter. Several theoretical predictions for their mass spectra exist, ranging from nonrelativistic variational estimates and potential model analyses Llanes-Estrada et al. (2012); Wei et al. (2017); Yang et al. (2020); Gómez-Rocha et al. (2023); Najjar et al. (2024) to recent quantum computing simulations based on the Cornell potential de Arenaza et al. (2024). In particular, the triply charmed Ω3c{\rm\Omega}_{3c} baryon is expected to lie well within the discovery reach of future hadron colliders such as the High-Luminosity Large Hadron Collider (HL-LHC) Apollinari et al. (2015), with even greater prospects foreseen at the Future Circular Collider (FCC) Benedikt et al. (2025a, b, c), and may also be accessible at present and future lepton facilities like Belle II Abe et al. (2010).

The ATLAS, CMS, and LHCb experiments have already accumulated substantial data sets at 13 and 13.6 TeV, and projections for HL-LHC indicate that searches for triply heavy baryons could be feasible, especially in boosted regimes where modern tracking and vertexing can be exploited Aaboud et al. (2016); Hayrapetyan et al. (2024a); Aaij et al. (2021b). Forward coverage, low transverse-momentum capabilities, and excellent mass resolution are key ingredients that make LHCb particularly well suited for this task. The proposed FCC collider, with its 100 TeV energy and unprecedented luminosity, would enhance discovery prospects even further, offering a unique opportunity to explore the heavy baryon sector beyond current limits.

From a phenomenological standpoint, triply heavy baryons are particularly compelling because of the unique nature of their formation mechanism. More generally, the production of multiply heavy hadrons is expected to proceed predominantly through the fragmentation of high-energy partons into baryonic final states (see, e.g., Refs. Braaten et al. (1995); Braaten and Yuan (1993); Braaten et al. (1993a, 1994, b); Kiselev et al. (1994)). This process is sensitive to both perturbative short-distance dynamics and nonperturbative aspects of hadronization. In the case of heavy baryons, a practical framework for modeling this mechanism is offered by the quark-diquark picture Anselmino et al. (1993); Ebert et al. (1996), where a tightly bound diquark (e.g., |cc|cc\rangle) forms first and subsequently hadronizes into the full baryon via the capture of an additional heavy quark Moosavi Nejad and Sartipi Yarahmadi (2016); Chang et al. (2006).

This diquark-based factorization simplifies the computation of fragmentation functions (FFs), which encode the probability density for a parton to hadronize into a specific final-state baryon carrying a given fraction of its momentum. Since heavy-quark masses lie above the perturbative QCD threshold, the initial-scale inputs of FFs for heavy-flavored baryons incorporate both perturbative and nonperturbative components. These can be modeled through effective parametrizations, or derived from wavefunction overlaps inspired by potential models and nonrelativistic effective frameworks Caswell and Lepage (1986); Bodwin et al. (1995); Braaten and Yuan (1993); Cho and Leibovich (1996a, b); Bodwin et al. (2005).

Theoretical studies of fragmentation into triply heavy baryons have been carried out at leading order (LO) and next-to-leading order (NLO) in αs\alpha_{s} Adamov and Goldstein (1997); Yang (2002); Gomshi Nobary and Sepahvand (2005); Moosavi Nejad and Delpasand (2017); Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019), with particular emphasis on charm- and gluon-initiated channels. These analyses underscore the role of diquark correlations, binding energy corrections, and the underlying color structure in the baryon formation process. To obtain realistic collider predictions, it is essential to incorporate higher-order corrections and parton evolution effects Cacciari and Greco (1994); Buza et al. (1998); Cacciari and Catani (2001); Mitov and Moch (2006).

Experimentally, while no triply heavy baryon has been confirmed, their indirect detection might occur through decay chains of exotic hadrons An et al. (2019); Ortiz-Pacheco and Bijker (2023); Liu et al. (2024). Multiquark states such as tetraquarks and pentaquarks with hidden or open heavy flavor (see Refs. Chen et al. (2016); Esposito et al. (2017); Lebed et al. (2017) for a review) are routinely produced in LHC collisions, and their decay topologies may involve final states three valence heavy quarks. In fact, Ω3c{\rm\Omega}_{3c} can emerge from hadronic transitions of heavy exotic hadrons or as a daughter particle in sequential decays Chen and Wu (2011); Wang and Xu (2018); Yang et al. (2020).

The motivation to construct realistic FFs for triply heavy baryons is thus multifold: improving our theoretical control on hadronization in the heavy sector, enabling predictions for rare baryon production at the LHC and future colliders, and contributing to the broader understanding of exotic matter in QCD, particularly through the role of Ω3c{\rm\Omega}_{3c} states as possible decay products of multiquark exotics such as pentaquarks. To this end, we introduce in this work the OMG3Q1.0 sets Celiberto (2025b), the first public release of collinear FFs for the Ω3c{\rm\Omega}_{3c} baryon. This set extends and complements recent efforts to characterize rare and exotic hadron production via fragmentation, such as those for doubly and fully heavy tetraquarks Celiberto and Gatto (2024a, b); Celiberto (2025c); Celiberto and Gatto (2025); Celiberto (2025a), and fully heavy pentaquarks Celiberto (2025d).

The OMG3Q1.0 functions are built by combining diquark-like proxy model NLO inputs for both the charm-quark Moosavi Nejad (2017) and gluon Delpasand and Moosavi Nejad (2019) channels at the lowest factorization scale. Energy evolution to higher scales is performed using (Dokshitzer-Gribov-Lipatov-Altarelli-Parisi) DGLAP equations in a variable-flavor-number scheme (VFNS) Mele and Nason (1991); Cacciari and Greco (1994); Buza et al. (1998), implemented via the HF-NRevo scheme Celiberto (2024a, 2025e, 2025f). The evolved FFs are provided in LHAPDF6 format Buckley et al. (2015), allowing for direct inclusion in high-energy simulation tools and cross section computations. They can be interfaced with resummation frameworks and parton-level Monte Carlo codes to generate realistic predictions for Ω3c{\rm\Omega}_{3c} production.

As a phenomenological application, we investigate the production of Ω3c{\rm\Omega}_{3c} plus jet systems at center-of-mass energies relevant for the (HL-)LHC and FCC. The semi-inclusive process [ppΩ3c+𝒳+jet{\rm p}{\rm p}\to{\rm\Omega}_{3c}+{\cal X}+{\rm jet}] (see Fig. 4) is studied in a NLL/NLO+{\rm NLL/NLO^{+}} hybrid framework that consistently merges NLO collinear factorization with the high-energy resummation of next-to-leading logarithmic (NLL+) energy contributions and beyond. Observables such as rapidity separations and transverse momentum spectra are computed via the JETHAD code and the symJETHAD symbolic plugin Celiberto (2021a, 2022a, 2023a, 2024b, 2024c).

This work contributes to the broader goal of developing a systematic and evolution-consistent description of the rare Ω\rm\Omega sector via VFNS fragmentation. It offers a benchmark for future studies of triply heavy states and lays the groundwork for comparisons with lattice QCD, potential models, and heavy-flavor QCD phenomenology.

The remainder of this paper is organized as follows. In Section 2, we briefly discuss general features of heavy-flavor fragmentation and then focus on the production of rare Ω\rm\Omega baryons. Section 3 details the architecture of our OMG3Q1.0 sets. The hybrid factorization framework is outlined in Section 4, and predictions for the HL-LHC and FCC are presented and discussed in Section 5. Conclusions and outlook are given in Section 6.

2 Heavy-flavor fragmentation and the Ω\rm\Omega sector

In the first part of this section, we provide a concise overview of key aspects of heavy-flavor fragmentation, covering heavy-light hadrons, quarkonia, and exotic bound states (Section 2.1). We then discuss the main features of the diquark-like proxy model applied to triply heavy baryons (Section 2.2). Finally, we present the general structure of Ω3c{\rm\Omega}_{3c} initial-scale FFs (Section 2.3).

2.1 Highlights on heavy-flavor fragmentation

The fragmentation of heavy-flavored hadrons is inherently more intricate than that of light hadrons. This complexity arises because heavy-quark masses in their lowest Fock states fall within the perturbative QCD domain. As a result, while light-hadron FFs encode only nonperturbative dynamics at the initial energy scale, heavy-hadron FFs require a combination of perturbative and nonperturbative elements.

For singly heavy hadrons such as DD mesons, BB mesons, and ΛQ{\rm\Lambda}_{Q} baryons, the fragmentation process unfolds in two distinct stages Cacciari et al. (1997); Cacciari and Greco (1994); Jaffe and Randall (1994); Kniehl et al. (2005); Helenius and Paukkunen (2018, 2023). The first stage involves a parton ii, produced in a hard-scattering event with large transverse momentum, fragmenting into a heavy quark QQ. Since αs(mQ)<1\alpha_{s}(m_{Q})<1, this step can be computed within perturbative QCD.

This contribution, known as the short-distance coefficient (SDC) for the [iQ][i\to Q] fragmentation process, occurs on a much shorter timescale than hadronization. The first NLO determination of SDCs for singly heavy hadrons was presented in Ref. Mele and Nason (1990, 1991), with subsequent next-to-NLO refinements provided in Refs. Rijken and van Neerven (1996); Mitov and Moch (2006); Blumlein and Ravindran (2006); Melnikov and Mitov (2004); Mitov (2005); Biello and Bonino (2024).

At later timescales, the heavy quark QQ transitions into a bound hadronic state. This second stage of fragmentation is entirely nonperturbative and is commonly modeled through phenomenological approaches Kartvelishvili et al. (1978); Bowler (1981); Peterson et al. (1983); Andersson et al. (1983); Collins and Spiller (1985); Colangelo and Nason (1992) or effective field theories Georgi (1990); Eichten and Hill (1990); Grinstein (1992); Neubert (1994); Jaffe and Randall (1994).

In mathematical terms, for a parton ii fragmenting into a singly heavy hadron Q{\cal H}_{Q} at the initial scale μF,0\mu_{F,0} of the order of the mass of the heavy-quark, mQm_{Q}, one has Cacciari et al. (1997); Cacciari and Greco (1997)

DiQ(z,μF,0)=z1dξξDiQ(ξ,μF,0)D[np]Q(zξ).D_{i}^{{\cal H}_{Q}}(z,\mu_{F,0})=\int_{z}^{1}\frac{{\rm d}\xi}{\xi}D_{i}^{Q}(\xi,\mu_{F,0})\,D_{\rm[np]}^{{\cal H}_{Q}}\left(\frac{z}{\xi}\right)\;. (1)

Here, DiQD_{i}^{Q} is the perturbative part of the initial scale FF, namely the SDC for an outgoing (massless) parton to fragment, via a perturbative QCD cascade, into the constituent (massive) heavy quark. Conversely, D[np]QD_{\rm[np]}^{{\cal H}_{Q}} depicts the nonperturbative part of the fragmentation, which is taken to be universal, independent of the generating parton ii. It also does not depend on μF,0\mu_{F,0}.

To construct a fully evolved set of FFs within the VFNS, it is crucial to account for energy evolution effects. Starting from initial-scale inputs, assumed to be free from scaling violations, the functions are evolved numerically using DGLAP timelike evolution equations at the required level of perturbative accuracy.

The two-step initial-scale fragmentation framework, originally devised for singly heavy hadrons, can be extended to quarkonia. In this case, the simultaneous presence of a heavy quark QQ and its antiquark Q¯\bar{Q} within the lowest Fock state |QQ¯|Q\bar{Q}\rangle adds an additional layer of complexity to the fragmentation description. The modern theoretical treatment of quarkonium production is based on nonrelativistic QCD (NRQCD) Caswell and Lepage (1986); Thacker and Lepage (1991); Bodwin et al. (1995); Cho and Leibovich (1996a, b); Leibovich (1997); Bodwin et al. (2005) (for a pedagogical introduction, see Refs. Grinstein (2000); Krämer (2001); Brambilla et al. (2004); Lansberg (2005, 2020)).

NRQCD treats heavy-quark and antiquark fields as nonrelativistic degrees of freedom, enabling a systematic factorization between SDCs, which govern the perturbative production of the [QQ¯][Q\bar{Q}] pair, and long-distance matrix elements (LDMEs), which encapsulate the nonperturbative hadronization dynamics. A physical quarkonium state is then expressed as a linear superposition of all possible Fock states, organized through a systematic expansion in both the strong coupling αs\alpha_{s} and the relative velocity v𝒬v_{\cal Q} of the heavy quark-antiquark pair.

A key feature of NRQCD is its ability to describe quarkonium production mechanisms across a wide range of transverse momenta. At low transverse momentum, qTq_{T}, the dominant process involves the short-distance formation of the [QQ¯][Q\bar{Q}] pair in the hard scattering, followed by its nonperturbative evolution into a physical bound state. At higher qTq_{T}, however, an alternative mechanism—where a single parton fragments into the quarkonium state plus additional radiation—begins to compete with, and eventually surpasses, the short-distance process.

From a theoretical perspective, short-distance quarkonium production can be interpreted as a fixed-flavor number scheme (FFNS) Alekhin et al. (2010) two-parton fragmentation, capturing higher-twist power corrections (see Refs. Fleming et al. (2012); Kang et al. (2014); Echevarria (2019); Boer et al. (2023); Celiberto (2024a, 2025e, 2025f) for more details). By contrast, the single-parton mechanism at large qTq_{T} is a collinear VFNS fragmentation, its energy evolution being governed by the DGLAP equations.

The first LO calculations of the initial-scale inputs for both gluon and heavy-quark FFs of SS-wave color-singlet vector quarkonia were performed in the early 1990s Braaten and Yuan (1993); Braaten et al. (1993a). However, corresponding NLO refinements became available only recently Zheng et al. (2019a, 2022).

Building upon these inputs, a pioneering set of VFNS, DGLAP-evolving FFs for vector quarkonia, ZCW19+, was introduced in Refs. Celiberto and Fucilla (2022a); Celiberto (2023a). Shortly thereafter, the ZCFW22 extension was developed to incorporate BcB_{c} mesons Celiberto (2022b, 2024d). Here, high-energy resummed production rates of charmed BB mesons, derived using the ZCFW22 FF framework, reinforced the experimental observation by the LHCb Collaboration Aaij et al. (2015b, 2017); Celiberto (2024d) that the relative production rate of Bc(1S0)B_{c}(^{1}S_{0}) mesons compared to singly bottomed BB mesons remains below 0.1% Celiberto (2024d). This result simultaneously benchmarked for both the high-energy resummation framework and the NRQCD fragmentation applied to BcB_{c} mesons.

Remarkably, recent studies indicate that NRQCD factorization can be suitably extended to explore the production mechanism of di-J/ψJ/\psi excitations Aaij et al. (2020); Aad et al. (2023); Hayrapetyan et al. (2024b), interpreting these resonances as fully charmed, exotic tetraquarks Zhang and Ma (2020); Zhu (2021). From a NRQCD viewpoint, the production of a fully heavy T4QT_{4Q} tetraquark originates from the generation of a heavy-quark pair and its corresponding antiquark pair at short distances, with a characteristic energy scale set by the inverse of the heavy-quark mass. As with singly heavy hadrons and quarkonia, asymptotic freedom allows for the usual heavy-flavor fragmentation mechanism in two steps: an initial perturbative stage followed by a nonperturbative hadronization phase.

The first NRQCD calculation of the initial-scale FF for the [gT4Q][g\to T_{4Q}] SS-wave channel in a color-singlet configuration appeared in Ref. Feng et al. (2022a). Building on this foundation, the recent TQ4Q1.0 VFNS sets combined this NRQCD-based gluon channel with an initial-scale input for the [QT4Q][Q\to T_{4Q}] function, derived by adapting a Suzuki-model approach Suzuki (1977, 1986); Amiri and Ji (1987); Moosavi Nejad and Amiri (2022).

Expanding this methodology, the first VFNS FFs for heavy-light tetraquarks, named TQHL1.0 determinations, were derived in Refs. Celiberto and Papa (2024); Celiberto (2024b). The latest release, the TQ4Q1.1 and TQHL1.1 families Celiberto and Gatto (2025); Celiberto (2025a), further refined the description by incorporating NRQCD-based modeling of also the [QT4Q][Q\to T_{4Q}] initial-scale FF Bai et al. (2024), and improving the treatment of XQqQ¯q¯X_{Qq\bar{Q}\bar{q}} fragmentation.

Finally, Ref. Celiberto (2025d) accompanies and presents the first release of collinear FFs for fully charmed |cccc¯c|ccc\bar{c}c\rangle pentaquarks. The resulting set PQ5Q1.0 embodies a consistent heavy-quark threshold DGLAP evolution from two possible initial-scale inputs for the heavy-quark channel: a compact direct multicharm state Farashaeian and Moosavi Nejad (2024a) or a dicharm-charm-dicharm configuration Farashaeian and Moosavi Nejad (2024b).

2.2 The diquark-like proxy model

The quark-diquark model provides a simplified, yet powerful framework for describing the internal structure of baryons, wherein two quarks are treated as a tightly bound subsystem, referred to as a diquark, that interacts with the third quark Gell-Mann (1964). This approach has been widely used in hadron spectroscopy and in the analysis of production and decay mechanisms of baryons, particularly those containing heavy quarks (see Refs. Maiani et al. (2005); Jaffe and Wilczek (2003); Guo et al. (2013); De Sanctis et al. (2016) for related discussions). Within this picture, diquarks are modeled either as scalar objects with spin-0 or as axial-vector states with spin-1. Their internal structure, inherently nonperturbative, is typically encoded through phenomenological form factors.

Scalar diquarks involve a single form factor and are generally associated with simpler spin configurations, while axial-vector diquarks require multiple form factors to describe their richer internal dynamics. Both types have been utilized in the modeling of spin-dependent processes in baryon formation, and also in the parameterization of polarized quark and gluon distributions inside the nucleon, especially within spectator models Bacchetta et al. (2008, 2010, 2020, 2024a); Chakrabarti et al. (2023); Banu et al. (2024).

Early implementations of the quark-diquark framework to describe the fragmentation of light and heavy baryons can be found in Refs. Nzar and Hoodbhoy (1995); Ma et al. (2002); Yang (2002) and Falk et al. (1994); Adamov and Goldstein (1997); Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019), respectively. The same formalism has also been extended to the analysis of exotic hadrons, such as pentaquarks Maiani et al. (2015), and to the mass spectroscopy of doubly and fully heavy tetraquarks using relativistic quasipotential approaches based on diquark-antidiquark interactions Faustov et al. (2020, 2021, 2022).

In the present work, we adopt a modeling strategy that has become standard in the literature Adamov and Goldstein (1997); Martynenko and Saleev (1996); Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019), in which the diquark is considered to be a scalar constituent. This choice is primarily motivated by computational convenience. It simplifies the spin algebra, reduces the number of independent form factors, and results in a more manageable analytical structure. It has been widely used in the modeling of unpolarized J=1/2J=1/2 baryons in different approaches. For example, in Adamov and Goldstein (1997), the scalar diquark model is used to calculate collinear FFs within a perturbative QCD framework, emphasizing its tractability. Similarly, the works in Refs. Martynenko and Saleev (1996); Gomshi Nobary and Sepahvand (2007) show that adopting a scalar diquark allows one to construct models that are consistent with available data, while keeping the formalism analytically viable.

In contrast, the use of axial-vector diquarks introduces greater theoretical complexity, including the need for additional form factors and a more elaborate treatment of spin correlations, which limits their use in practical calculations. Nevertheless, axial-vector diquarks are essential for describing baryons with spin-3/23/2 and for incorporating polarization effects in both spin-1/21/2 and spin-3/23/2 states. In particular, any realistic modeling of the polarized fragmentation or production of the Ω3c\rm\Omega_{3c}^{*} resonance, with J=3/2J=3/2, requires a vector diquark configuration.

The scalar diquark approximation employed in this study should therefore be interpreted as an effective model for initiating a perturbative QCD-based description of fragmentation into unpolarized triply heavy baryons such as Ω3c{\rm\Omega}_{3c} states. More sophisticated treatments incorporating spin-1 diquarks could be pursued in future work to provide a more complete account of polarization and spin structure.

Interestingly, our adoption of the scalar diquark picture in the description of unpolarized Ω3c{\rm\Omega}_{3c} and Ω3b{\rm\Omega}_{3b} production is fully consistent with the color and spin symmetry requirements of the baryonic wavefunction, provided that the process is described within the two-step fragmentation mechanism. In this framework, the heavy quark first fragments into a color-antitriplet heavy diquark, which subsequently hadronizes into the triply heavy baryon through nonperturbative QCD effects.

Although scalar diquarks alone do not manifest the full symmetrization structure expected for a J=1/2J=1/2 baryon composed of three identical fermions, this structure can be effectively recovered through hadronization effects encoded in the purely nonperturbative component of the FF. This interpretation, while not explicitly stated in Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2020), is in line with the modeling strategies adopted therein, where scalar diquarks are used in the indirect channel to compute FFs into Ω3c{\rm\Omega}_{3c} and Ω3b{\rm\Omega}_{3b}, yielding consistent quantum number assignments and phenomenologically viable results.

2.3 General structure of Ω3c{\rm\Omega}_{3c} FFs

Refer to caption
Refer to caption
Figure 1: Representative leading diagrams for the diquark-like proxy model of the initial-scale collinear fragmentation of a constituent heavy quark (left) or a gluon (right) into a color-singlet SS-wave Ω3Q{\rm\Omega}_{3Q} baryon. Double lines stand for 𝒟{\cal D} or 𝒟¯\bar{{\cal D}} diquark states, while orange blobs represent the nonperturbative hadronization component of corresponding FFs. Black bullet vertices denote scalar diquark form-factor couplings.

The FF initial-scale inputs used in this study are extracted from existing calculations in the literature, which focus on the production of triply heavy baryons in a color-singlet configuration. Specifically, we rely on perturbative results for the quark-induced channel computed at LO in Ref. Moosavi Nejad and Delpasand (2017) and at NLO in Ref. Moosavi Nejad (2017), while for the gluon-induced channel we use the NLO result from Ref. Delpasand and Moosavi Nejad (2019). In all cases, the baryon is modeled as a color-singlet state formed from three charms, and the diquark is treated as a spin-0 scalar constituent.

Two main production scenarios are typically considered for phenomenological explorations. The first is the direct fragmentation mechanism, in which the heavy quark fragments directly into the three-quark baryon without invoking any internal substructure. This approach involves three-body diagrams and includes the direct interaction among the three constituent quarks. The second scenario, which we adopt here, is the diquark fragmentation model. In this framework, two of the three quarks are assumed to form a tightly bound scalar diquark, which then combines with the remaining quark to form the baryon. As mentioned before, this approximation significantly reduces the complexity of the calculation and has been widely used to describe the unpolarized production of J=1/2J=1/2 heavy baryons.

The so-called indirect mechanism, as referred to in Refs. Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019), corresponds precisely to the two-stage fragmentation framework discussed in Section 2.1, adapted to the diquark picture. From a theoretical standpoint, this structure offers a robust and systematically extensible framework for modeling the fragmentation of heavy-flavored hadrons. First, it separates the perturbative production of an intermediate colored object (the diquark) from its nonperturbative hadronization into the physical baryon. Second, both the quark and gluon FFs have been calculated at NLO within this setup, making it a consistent and reliable foundation for a VFNS analysis.

The Suzuki model Suzuki (1986) is among the ingredients of the framework adopted in Refs. Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019). In that context, it plays a role analogous to that of NRQCD in quarkonium production. Originally developed to describe the fragmentation of heavy quarks into baryons or mesons, the Suzuki picture provides both a structural and a computational foundation for the two-stage fragmentation process. It defines a perturbative FF transition amplitude, where the constituent heavy quark fragments into a constituent diquark and a spectator, followed by the hadronization of the diquark into a baryonic bound state. The model incorporates kinematic simplifications, such as the collinearity of final-state momenta, and introduces non-perturbative elements through the wave function at the origin and an effective diquark form factor.

The Suzuki model is used in both direct and diquark fragmentation scenarios studied in the literature, although its role is far more central in the latter. In the diquark-based case, it provides the full structure of the two-body fragmentation mechanism and enables the derivation of compact analytic expressions for the SDCs. This is particularly useful for modeling Ω3c{\rm\Omega}_{3c} baryons, where handling full QCD three-body fragmentation becomes analytically and symbolically prohibitive.

Adopting Suzuki to compute the perturbative part of the FF is motivated by both analytical practicality and physical consistency. While rooted in perturbative QCD, the approach forgoes operator-based expansions in favor of a more “diagrammatic” strategy. This yields a simplified yet effective tool for evaluating SDCs, especially in the case of complex multi-quark final states such as triply heavy baryons.

First, the Suzuki picture offers an analytically accessible framework that enables the closure of phase-space integrals and yields manageable expressions for amplitudes. Spin and color algebra are simplified, and the use of the wave function at the origin effectively absorbs the bound-state dynamics without invoking full nonperturbative QCD machinery. This is especially valuable in contexts where rigorous QCD calculations requiring renormalization, operator regularization, or twist expansions are unfeasible or unavailable.

Second, the Suzuki model retains an explicit connection to the internal structure of the hadron. Rather than treating the baryon as a pointlike final state, it describes the fragmentation process using constituent-level vertices, including spin-dependent terms, propagators for internal degrees of freedom, and color couplings. This construction ensures that the hard subprocess remains sensitive to the underlying baryonic wave function, preserving the link between perturbative production and hadron-level observables.

Third, the model is particularly well suited for baryons like Ω3c{\rm\Omega}_{3c}, where direct short-distance transitions in full perturbative QCD involve the algebraic disentangling of intertwined Dirac, Lorentz, and color structures. By rephrasing the process as a two-body fragmentation, where the diquark is treated as an effective, quasi-elementary object, the Suzuki model circumvents these challenges while preserving the overall dynamical content.

The analogy with NRQCD is particularly instructive. Just as NRQCD allows for a systematic separation of short-distance and long-distance contributions in quarkonium production, the Suzuki model supports a clean factorization between perturbative dynamics, described by the SDCs, and nonperturbative hadronization. Both frameworks rely on the use of effective composite degrees of freedom (diquark or heavy-quark nonrelativistic fields), which serve to simplify the treatment of fragmentation processes while maintaining fidelity to the underlying physics.

A crucial component in the construction of our initial-scale FFs is the modeling of the nonperturbative transition from the constituent diquark 𝒟{\cal D} to the physical baryon Ω3c{\rm\Omega}_{3c}. Following Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019), in this work we adopt the phenomenological Peterson-Schlatter-Schmitt-Zerwas (PSSZ) parametrization Peterson et al. (1983), which has been widely used in heavy-flavor fragmentation studies and is particularly suited for scenarios involving hadrons with multiple heavy quarks. We have

D𝒟Ω3c(z)D[np][PSSZ](z)=𝒩𝒫z(1z)2[(1z)2+b𝒫z]2,D_{\cal D}^{{\rm\Omega}_{3c}}(z)\equiv D_{\rm[np]}^{\rm[PSSZ]}(z)={\cal N}_{\cal P}\frac{z(1-z)^{2}}{\left[(1-z)^{2}+b_{\cal P}z\right]^{2}}\;, (2)

with the normalization factor

𝒩𝒫={b𝒫26b𝒫+4(4b𝒫)4b𝒫b𝒫2×[arctanb𝒫4b𝒫b𝒫2+arctan2b𝒫4b𝒫b𝒫2]+12lnb𝒫+14b𝒫}.\begin{split}{\cal N}_{\cal P}&\,=\,\left\{\frac{b_{\cal P}^{2}-6b_{\cal P}+4}{(4-b_{\cal P})\sqrt{4b_{\cal P}-b_{\cal P}^{2}}}\right.\\ &\times\,\left[\arctan\frac{b_{\cal P}}{\sqrt{4b_{\cal P}-b_{\cal P}^{2}}}+\arctan\frac{2-b_{\cal P}}{\sqrt{4b_{\cal P}-b_{\cal P}^{2}}}\right]\\ &+\,\left.\frac{1}{2}\ln b_{\cal P}+\frac{1}{4-b_{\cal P}}\right\}\;.\end{split} (3)

being obtained by summing over all hadrons containing the charm quark Cacciari and Greco (1997) (see also Section IV of Ref. Peterson et al. (1983)). From the original derivation of the model, we note that the b𝒫b_{\cal P} cannot be calculated exactly, since it is intrinsically connected to nonperturbative effects.

Indeed, for a singly charmed hadron, say a DD meson with a leading Fock state |cq¯|c\bar{q}\rangle, the value of b𝒫b_{\cal P} is related to the charm-quark mass by b𝒫Λ/mcb_{\cal P}\approx{\rm\Lambda}/m_{c}, where Λ{\rm\Lambda} stands for a hadronic scale of the order of the mass of the constituent light antiquark, mq¯m_{\bar{q}}. In such a case, one finds that z=1b𝒫\langle z\rangle=1-\sqrt{b_{\cal P}}, which is in line with the prediction of heavy-quark FFs scaling linearly with the heavy-quark mass Suzuki (1977); Bjorken (1978); Kinoshita (1986); Cacciari et al. (1997). On the contrary, for a fully charmed baryon Ω3c{\rm\Omega}_{3c}, the b𝒫b_{\cal P} is connected to the charm and anticharm masses, which are above the perturbative threshold. Following Refs. Adamov and Goldstein (1997); Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019), we set b𝒫=[mc/(mc+mc¯)]2=1/4b_{\cal P}=[m_{c}/(m_{c}+m_{\bar{c}})]^{2}=1/4.

This choice is supported by previous studies, where the Peterson form is explicitly used to model the nonperturbative stage of the diquark hadronization in fragmentation to triply heavy baryons Moosavi Nejad (2017); Delpasand and Moosavi Nejad (2019). Work in Ref. Moosavi Nejad (2017) clearly emphasize that the Peterson function effectively captures the essential nonperturbative dynamics of the fragmentation process and is compatible with the overall two-step fragmentation scheme defined through the Suzuki model. Similarly, Ref. Delpasand and Moosavi Nejad (2019) supports the use of the Peterson parametrization on the basis of its widespread adoption and its ability to encode realistic hadronization patterns for heavy baryons. Moreover, the Peterson form complements the output of the perturbative calculation by providing a universal and physically interpretable nonperturbative component that fits naturally into collinear factorization frameworks.

The Peterson model, by contrast, was originally formulated to describe the fragmentation of heavy quarks in which the initiating heavy quark carries most of the hadron momentum. This structure is naturally suited to hadrons containing heavy constituents, since the resulting hadron tends to follow the direction and kinematics of the fragmenting parton. In the case of triply heavy baryons, where all valence quarks are heavy, the same reasoning applies: the baryon is formed predominantly from the massive fragments of the initial parton, and thus inherits a large fraction of its momentum. In this respect, the Peterson function provides a realistic and physically consistent parametrization of the nonperturbative transition, capturing the expected peak at large z and the suppression at small z, in line with the kinematics of systems such as triply heavy baryons.

This behavior is particularly well suited for modeling the hadronization of a heavy diquark into a triply heavy baryon, where the recombination process is dominated by massive constituents and the baryon inherits a substantial fraction of the momentum of the initiating quark or gluon. The simple analytic structure of the model, governed by a single shape parameter b𝒫b_{\cal P}, allows for the adjustment to match the internal mass hierarchy and spatial configuration of the final-state baryon, while ensuring smooth integration within collinear convolution schemes.

In this context, the Peterson form meets several important criteria: it is grounded in heavy-flavor dynamics, it offers analytical tractability, it aligns with the factorized structure emerging from the Suzuki picture, and it preserves consistency with the DGLAP evolution in the VFNS. For these reasons, we consider it a natural and robust choice for describing the nonperturbative input to the Ω3c{\rm\Omega}_{3c} initial-scale FF in our two-step formalism.

The compact expression for the FF of a parton ii generating a Ω3Q{\rm\Omega}_{3Q} baryon at the initial scale μF,0\mu_{F,0} reads

DiΩ3c(z,μF,0)=z1dξξDi𝒟(ξ,μF,0)D𝒟Ω3c(zξ).D_{i}^{{\rm\Omega}_{3c}}(z,\mu_{F,0})=\int_{z}^{1}\frac{{\rm d}\xi}{\xi}D_{i}^{\cal D}(\xi,\mu_{F,0})\,D_{\cal D}^{{\rm\Omega}_{3c}}\left(\frac{z}{\xi}\right)\;. (4)

Equation (4) essentially matches the general structure of the heavy-hadron initial-scale FF as in Eq. (1), with two suitable adjustments. We notice that

DiQ(ξ,μF,0)\displaystyle D_{i}^{Q}(\xi,\mu_{F,0})\quad Di𝒟(ξ,μF,0),\displaystyle\longrightarrow\quad D_{i}^{\cal D}(\xi,\mu_{F,0})\;, (5a)
D[np]Q(zξ)\displaystyle D_{\rm[np]}^{{\cal H}_{Q}}\left(\frac{z}{\xi}\right)\quad D𝒟Ω3c(zξ).\displaystyle\longrightarrow\quad D_{\cal D}^{{\rm\Omega}_{3c}}\left(\frac{z}{\xi}\right)\;. (5b)

Here we have SDCs describing the perturbative splitting of a massless parton ii into a diquark state 𝒟{\cal D} (Eq. (5a)), followed by the long-distance, nonperturbative hadronization of the diquark into the observed triply heavy baryon Ω3c{\rm\Omega}_{3c} (Eq. (5b)). In other words, the diquark 𝒟{\cal D} replaces the QQ entering the lowest Fock state of the singly heavy hadron Q{\cal H}_{Q}. As mentioned, the nonperturbative D𝒟Ω3cD_{\cal D}^{{\rm\Omega}_{3c}} will be modeled according to PSSZ as in Eq. (2), while gluon and charm SDCs will be presented and discussed in the next section.

3 The OMG3Q1.0 architecture

In this section, we outline our scheme for developing the hadron-structure-oriented OMG3Q1.0 VFNS FF determinations for fully charmed Ω{\rm\Omega} baryons, starting from diquark-like proxy model inputs for both the charm-quark and gluon channels at the initial energy scales, and accounting from a threshold-consistent DGLAP as follows from the HF-NRevo methodology Celiberto (2024a, 2025e, 2025f).

All calculations necessary to build our OMG3Q1.0 set were carried out using symJETHAD, the newly developed Mathematica Inc. plugin of JETHAD Celiberto (2021a, 2022a, 2023a, 2024b, 2024c), designed for the symbolic manipulation of analytical formulas pertinent to the hadronic structure and precision QCD.

Sections 3.1 and 3.2 detail the charm and gluon inputs, respectively. The DGLAP/HF-NRevo timelike evolution of the OMG3Q1.0 functions is then analyzed and discussed in Section 3.3.

3.1 Initial-scale charm fragmentation

The initial-scale FF for the [cΩ3c][c\to{\rm\Omega}_{3c}] transition in the diquark-like scenario is diagrammatically shown in the left panel of Fig. 1. In this representation, the double lines denote dicharm states, 𝒟|cc{\cal D}\equiv|cc\rangle, or their corresponding antidiquark partners, 𝒟¯|c¯c¯\bar{{\cal D}}\equiv|\bar{c}\bar{c}\rangle. The orange blob illustrates the nonperturbative hadronization component of the FF, labeled as D[np]QD𝒟Ω3cD_{\rm[np]}^{{\cal H}_{Q}}\equiv D_{\cal D}^{{\rm\Omega}_{3c}}. The black bullet vertex represents the scalar diquark form factor coupling.111While not explicitly shown in the formulæ of this Section, our treatment mirrors that of Ref. Moosavi Nejad (2017), where the form factor is assumed to exhibit a simple pole in transverse-momentum space at 5 GeV.

We note that the diagram on the left panel of Fig. 1 corresponds to the [c(c,𝒟)+𝒟¯][c\to(c,{\cal D})+\bar{\cal D}] splitting. By swapping all charm quarks with anticharms, one obtains the [c¯(c¯,𝒟¯)+𝒟][\bar{c}\to(\bar{c},\bar{\cal D})+{\cal D}] process. A detailed analysis of potential asymmetries in the production of triply charmed Ω\rm\Omega baryons and antibaryons is left for future dedicated studies. In this work, we assume full symmetry in the formation mechanisms of Ω3c{\rm\Omega}_{3c} and Ω¯3c\bar{{\rm\Omega}}_{3c} states. Accordingly, our predictions are based on observables sensitive to the inclusive, charge-averaged emission of baryons and antibaryons. Assuming this symmetry holds, the cc and c¯\bar{c} fragmentation channels are treated on equal footing (for comparison with the light-hadron case, see Ref. Bertone et al. (2018)).

Using symJETHAD Celiberto (2021a, 2022a, 2023a, 2024b, 2024c) and FeynCalc Mertig et al. (1991); Shtabovenko et al. (2016, 2020), we symbolically calculated and obtained the explicit form of the LO and NLO initial-scale charm SDC, finding agreement with Refs. Moosavi Nejad and Delpasand (2017) and Moosavi Nejad (2017), respectively. As for the LO case, we write

Dc𝒟(z,μF,0)αs2dc[LO](z;μF,0)+αs3dc[NLO](z;μF,0).\begin{split}D_{c}^{\cal D}(z,\mu_{F,0})\,&\equiv\alpha_{s}^{2}{\rm d}_{c}^{\rm[LO]}(z;\mu_{F,0})\\[2.84544pt] \,&+\,\alpha_{s}^{3}\,{\rm d}_{c}^{\rm[NLO]}(z;\mu_{F,0})\;.\end{split} (6)

with αsαs(5mc)\alpha_{s}\equiv\alpha_{s}(5m_{c}) and

dc[LO](z;μF,0)=𝒩c(z)𝒮c[LO](z;qT/c)𝒯c[LO](z;qT/c),{\rm d}_{c}^{\rm[LO]}(z;\mu_{F,0})={\cal N}_{c}(z)\frac{{\cal S}_{c}^{\rm[LO]}(z;{\cal R}_{q_{T}/c})}{{\cal T}_{c}^{\rm[LO]}(z;{\cal R}_{q_{T}/c})}\;, (7)

where

qT/c2qT 2/mc2,{\cal R}_{q_{T}/c}^{2}\equiv\langle{\vec{q}}_{T}^{\;2}\rangle/m_{c}^{2}\;, (8)
𝒩c(z)=π26f2CF2z3(1z)3,{\cal N}_{c}(z)=\frac{\pi^{2}}{\sqrt{6}}\,f_{\cal B}^{2}C_{F}^{2}\,z^{3}(1-z)^{3}\;, (9)
𝒮c[LO](z;qT/c)= 4(41z280z+96+256/z2)+qT/c2(8z3+5z216z+16)+qT/c4 4z2,\begin{split}{\cal S}_{c}^{\rm[LO]}&(z;{\cal R}_{q_{T}/c})\,=\,4(41z^{2}-80z+96+256/z^{2})\\ \,&+\,{\cal R}_{q_{T}/c}^{2}(8z^{3}+5z^{2}-16z+16)\\ \,&+\,{\cal R}_{q_{T}/c}^{4}\,4z^{2}\;,\end{split} (10)

and

𝒯c[LO](z;qT/c)=(qT/c2+(43z)2)2×qT/c2z2+z216z+16)2.\begin{split}{\cal T}_{c}^{\rm[LO]}&(z;{\cal R}_{q_{T}/c})\,=\,({\cal R}_{q_{T}/c}^{2}+(4-3z)^{2})^{2}\\ \,&\times\,{\cal R}_{q_{T}/c}^{2}\,z^{2}+z^{2}-16z+16)^{2}\;.\end{split} (11)

The NLO correction to the charm SDC was originally derived in Ref. Moosavi Nejad and Sartipi Yarahmadi (2016) for singly heavy mesons, and then employed for triply-heavy baryons (see Ref. Moosavi Nejad (2017)). It reads

dc[NLO](z;μF,0)=𝒩c(z)𝒮c[NLO](z;qT/c)𝒯c[NLO](z;qT/c),{\rm d}_{c}^{\rm[NLO]}(z;\mu_{F,0})={\cal N}_{c}(z)\frac{{\cal S}_{c}^{\rm[NLO]}(z;{\cal R}_{q_{T}/c})}{{\cal T}_{c}^{\rm[NLO]}(z;{\cal R}_{q_{T}/c})}\;, (12)

where

𝒮c[NLO](z;qT/c)= 96πz[qT/c14(1720z)z14+qT/c12(240z41664z3+ 3813z24130z+1711)z12+qT/c10(4368z639072z5+140635z4 269144z3+300490z2 189616z+52243)z10+qT/c8(20976z8208040z7+ 870151z62070634z5+3255115z4 3533740z3+2600201z2 1202330z+268205)z8+qT/c6 4(1z)2(11256z875035z7 151779z6+2535886z5 8913460z4+15864355z3 15844619z2+8395110z1823490)z6+qT/c4 4(1z)4(12324z822830z7 2045973z6+14641206z542277675z4+ 65338590z357387186z2+ 26330832z4933872)z4+qT/c2 48(1z)6(563z8+2272z7 252314z6+1752016z55068605z4+ 7770780z36800544z2+ 3505032z849528)z2+ 144(1z)8(41z8+438z736063z6+ 258240z5833328z4+1614600z3 1793016z2+676512z198288)],\begin{split}{\cal S}_{c}^{\rm[NLO]}&(z;{\cal R}_{q_{T}/c})\,=\,96\pi z\left[\right.{\cal R}_{q_{T}/c}^{14}\,(17-20z)z^{14}\\ \,&+\,{\cal R}_{q_{T}/c}^{12}\,(240z^{4}-1664z^{3}\\ \,&+\,3813z^{2}-4130z+1711)z^{12}\\ \,&+\,{\cal R}_{q_{T}/c}^{10}\,(4368z^{6}-39072z^{5}+140635z^{4}\\ \,&-\,269144z^{3}+300490z^{2}\\ \,&-\,189616z+52243)z^{10}\\ \,&+\,{\cal R}_{q_{T}/c}^{8}\,(20976z^{8}-208040z^{7}\\ \,&+\,870151z^{6}-2070634z^{5}+3255115z^{4}\\ \,&-\,3533740z^{3}+2600201z^{2}\\ \,&-\,1202330z+268205)z^{8}\\ \,&+\,{\cal R}_{q_{T}/c}^{6}\,4(1-z)^{2}(11256z^{8}-75035z^{7}\\ \,&-\,151779z^{6}+2535886z^{5}\\ \,&-\,8913460z^{4}+15864355z^{3}\\ \,&-\,15844619z^{2}+8395110z-1823490)z^{6}\\ \,&+\,{\cal R}_{q_{T}/c}^{4}\,4(1-z)^{4}(12324z^{8}-22830z^{7}\\ \,&-\,2045973z^{6}+14641206z^{5}-42277675z^{4}\\ \,&+\,65338590z^{3}-57387186z^{2}\\ \,&+\,26330832z-4933872)z^{4}\\ \,&+\,{\cal R}_{q_{T}/c}^{2}\,48(1-z)^{6}(563z^{8}+2272z^{7}\\ \,&-\,252314z^{6}+1752016z^{5}-5068605z^{4}\\ \,&+\,7770780z^{3}-6800544z^{2}\\ \,&+\,3505032z-849528)z^{2}\\ \,&+\,144(1-z)^{8}(41z^{8}+438z^{7}-36063z^{6}\\ \,&+\,258240z^{5}-833328z^{4}+1614600z^{3}\\ \,&-\,1793016z^{2}+676512z-198288)\left.\right]\;,\end{split} (13)

and

𝒯c[NLO](z;qT/c)=[qT/c2z2+4(32z)2]2×[qT/c2z2+(6z)2]2×[qT/c2z2+(1z)2]×[qT/c2z2+36(1z)2]2×[qT/c2z2+z235z+36].\begin{split}{\cal T}_{c}^{\rm[NLO]}(z;{\cal R}_{q_{T}/c})\,&=\,[{\cal R}_{q_{T}/c}^{2}\,z^{2}+4(3-2z)^{2}]^{2}\\ \,&\times\,[{\cal R}_{q_{T}/c}^{2}\,z^{2}+(6-z)^{2}]^{2}\\ \,&\times\,[{\cal R}_{q_{T}/c}^{2}\,z^{2}+(1-z)^{2}]\\ \,&\times\,[{\cal R}_{q_{T}/c}^{2}\,z^{2}+36(1-z)^{2}]^{2}\\ \,&\times\,[{\cal R}_{q_{T}/c}^{2}\,z^{2}+z^{2}-35z+36]\;.\end{split} (14)

Based on LO kinematics, we set the initial scale of charm fragmentation to μF,05mc\mu_{F,0}\equiv 5m_{c} (see Fig. 1, left panel).

Our charm FF differs from the one originally introduced in Ref. Moosavi Nejad (2017) in two main aspects. First, in that study, the constant 𝒩c(z){\cal N}_{c}(z) in Eq. (9) was not explicitly computed but was fixed by means of a normalization condition. Second, the choice of the qT 2\langle{\vec{q}}_{T}^{\;2}\rangle parameter, directly connected with the use of the Suzuki picture, requires further refinement.

The Suzuki model effectively accounts for spin correlations and serves as a proxy for transverse-momentum-dependent (TMD) FFs. In the collinear limit, instead of integrating over the squared transverse momentum of the outgoing charm quark, one may replace it with its average value, qT 2\langle{\vec{q}}_{T}^{\;2}\rangle. This substitution renders qT 2\langle{\vec{q}}_{T}^{\;2}\rangle a free parameter that must be fixed phenomenologically. Increasing qT 2\langle{\vec{q}}_{T}^{\;2}\rangle shifts the FF peak to lower values of zz while reducing the overall normalization Gomshi Nobary (1994).

The initial-scale quark FF presented in Ref. Moosavi Nejad (2017) was obtained by adopting qT 2=1GeV2\langle{\vec{q}}_{T}^{\;2}\rangle=1\,\text{GeV}^{2}, regarded as an upper-limit estimate for the average squared transverse momentum. In this work, we introduce a more refined selection of qT 2\langle{\vec{q}}_{T}^{\;2}\rangle, based on a balanced and phenomenologically motivated approach consistent with the exploratory goals of our analysis. This improvement is based on a preliminary refinement introduced in our previous investigation of collinear fragmentation into fully charmed tetraquarks Celiberto et al. (2024a). There, we drew on phenomenological insights from fragmentation studies of various hadrons produced in proton collisions. In particular, heavy-quark FFs for both light hadrons Celiberto et al. (2016a, 2017); Bolognino et al. (2018a); Celiberto (2021a) and heavy-flavored ones Celiberto et al. (2021a, b); Celiberto and Fucilla (2022a); Celiberto (2022b) were observed to be typically probed at average longitudinal momentum fractions z>0.4\langle z\rangle>0.4.

Based on a numerical analysis, we determined that setting qT 2T4c=70GeV2\langle{\vec{q}}_{T}^{\;2}\rangle_{T_{4c}}=70\,\text{GeV}^{2} for charm-to-T4cT_{4c} FFs results in an average zz greater than 0.4 and preserves a comparable order of magnitude between the quark and gluon channels. Extending this strategy to fully charmed |cccc¯c|ccc\bar{c}c\rangle pentaquarks, we adopted qT 2P5c=90GeV2\langle{\vec{q}}_{T}^{\;2}\rangle_{P_{5c}}=90\,\text{GeV}^{2} in Ref. Celiberto (2025d). Here, we fix qT 2Ω3c\langle{\vec{q}}_{T}^{\;2}\rangle_{{\rm\Omega}_{3c}} by correlating it with the peak position of the initial charm FFs. Following the same methodology used for qT 2T4c\langle{\vec{q}}_{T}^{\;2}\rangle_{T_{4c}} and qT 2P5c\langle{\vec{q}}_{T}^{\;2}\rangle_{P_{5c}}, we performed a numerical scan and selected qT 2Ω3c=60GeV2\langle{\vec{q}}_{T}^{\;2}\rangle_{{\rm\Omega}_{3c}}=60\,\text{GeV}^{2}.

A deeper justification supports our choice, going beyond heuristic arguments. Foundational studies on heavy-flavor fragmentation Suzuki (1977); Bjorken (1978); Kinoshita (1986); Peterson et al. (1983) have shown that heavy-quark FFs typically peak at large values of zz, with binding effects scaling in proportion to the heavy-quark mass. To illustrate this behavior, consider the fragmentation process leading to a DD meson, whose lowest Fock state is |cq¯|c{\bar{q}}\rangle, with total momentum kk and mass MDM_{D}.

In this context, the constituent heavy quark and the light antiquark are assumed to move with approximately the same velocity, vvcvqv\equiv v_{c}\simeq v_{q}. Their momenta can thus be expressed as kczk=mcvk_{c}\equiv zk=m_{c}v for the charm quark and kq=Λqvk_{q}={\rm\Lambda}_{q}v for the light antiquark, where Λq{\rm\Lambda}_{q} is a hadronic mass scale of order ΛQCD{\rm\Lambda_{\rm QCD}}. Since the mass of the DD meson satisfies MDmcM_{D}\approx m_{c}, one finds mcvk=kc+kq=zmcv+Λqvm_{c}v\approx k=k_{c}+k_{q}=zm_{c}v+{\rm\Lambda}_{q}v. This implies the approximate relation zc1Λq/mc\langle z\rangle_{c}\approx 1-{\rm\Lambda}_{q}/m_{c}, where the subscript ‘cc’ indicates the [cD][c\to D] FF.

As pointed out in Refs. Celiberto et al. (2024a); Celiberto (2025d), this behavior does not generally apply to fully heavy-flavored states such as quarkonia, triply charmed baryons, tetracharms, or pentacharms. In the valence configuration of these systems, no soft scale is present, as their lowest Fock components contain only heavy quarks. For fully heavy rare baryons or exotics like T4cT_{4c} or P5cP_{5c} states, the complex dynamics among the three, four, or five heavy constituents hinders a simple kinematic determination of the peak position of the FF.

As a preliminary example, we present the zz dependence of the [cΩ3c][c\to{{\rm\Omega}_{3c}}] initial-scale FFs used in the OMG3Q1.0 set. To estimate uncertainties at the starting scale, we apply a simplified, diagonal DGLAP evolution at NLO, similar to that in Ref. Celiberto et al. (2024a), but using only the PqqP_{qq} kernel for the charm FF. The left panel of Fig. 2 shows the zz-weighted charm FF, with μF,0=5mc\mu_{F,0}=5m_{c} and scale variations in the range 4mc<μF<6mc4m_{c}<\mu_{F}<6m_{c}. The zz-weighted FF peaks in the 0.45<z<0.50.45<z<0.5 range and drops to zero at the endpoints, as expected.

The small negative dip visible near z0.8z\gtrsim 0.8 may appear anomalous at first glance. However, dedicated numerical tests show that this feature originates from the choice of a relatively large value of qT 2\langle{\vec{q}}_{T}^{\;2}\rangle, higher than the one adopted in Ref. Moosavi Nejad (2017). This has no practical impact, as our tests confirm that FFs contribute dominantly to hadroproduction observables in the region 0.4z0.60.4\lesssim z\lesssim 0.6 (see Section 3.3). For z>0.8z>0.8, we have explicitly verified that no instability arises in high-energy cross-section predictions.

3.2 Initial-scale gluon fragmentation

The initial-scale FF for the [gΩ3c][g\to{\rm\Omega}_{3c}] transition in the diquark-like scenario is diagrammatically shown in the right panel of Fig. 1. Although both the charm- and gluon-initiated channels are considered leading within their respective topologies, they differ in their perturbative order. As shown in Fig. 1, the gluon-induced process involves one additional QCD vertex compared to the charm-induced one, thus contributing at a higher order in the strong coupling expansion. Specifically, the [cΩ3c][c\to{\rm\Omega}_{3c}] channel is known both at LO, corresponding to the left diagram, and at NLO, as discussed in Section 3.1. In contrast, the [gΩ3c][g\to{\rm\Omega}_{3c}] channel is currently known only at 𝒪(αs3)\mathcal{O}(\alpha_{s}^{3}), which corresponds to the diagram shown in the right panel.

This difference comes from the fact that a gluon, which does not carry a charm quantum number, must generate at least three charm quarks to produce Ω3c{\rm\Omega}_{3c}. Therefore, the first perturbative order at which such a process can occur is 𝒪(αs3)\mathcal{O}(\alpha_{s}^{3}), since the gluon must emit a [cc¯][c\bar{c}] pair to supply the free charm quark, generate a |cc|cc\rangle system to form the diquark (either directly or through an equivalent intermediate process) and include at least one additional QCD vertex to complete the baryon formation diagram. For consistency with this perturbative hierarchy, we denote the initial-scale gluon FF as the NLO function, while reserving the LO label for the lowest-order charm FF.

With symJETHAD Celiberto (2021a, 2022a, 2023a, 2024b, 2024c) and through the interface to FeynCalc Mertig et al. (1991); Shtabovenko et al. (2016, 2020), we symbolically computed the explicit form of the NLO initial-scale gluon SDC, originally provided in Ref. Delpasand and Moosavi Nejad (2019). One has

Dg𝒟(z,μF,0)αs3dg[NLO](z;μF,0).D_{g}^{\cal D}(z,\mu_{F,0})\,\equiv\,\alpha_{s}^{3}\,{\rm d}_{g}^{\rm[NLO]}(z;\mu_{F,0})\;. (15)

with αsαs(6mc)\alpha_{s}\equiv\alpha_{s}(6m_{c}) and

dg[NLO](z;μF,0)=𝒩g(z)𝒮g[NLO](z;qT/c)𝒯g[NLO](z;qT/c),{\rm d}_{g}^{\rm[NLO]}(z;\mu_{F,0})={\cal N}_{g}(z)\frac{{\cal S}_{g}^{\rm[NLO]}(z;{\cal R}_{q_{T}/c})}{{\cal T}_{g}^{\rm[NLO]}(z;{\cal R}_{q_{T}/c})}\;, (16)

where

𝒩g(z)=2π33f2CF2z3(1z)2,{\cal N}_{g}(z)=\frac{2\pi^{3}}{3}\,f_{\cal B}^{2}C_{F}^{2}\,z^{3}(1-z)^{2}\;, (17)
𝒮g[NLO](z;qT/c)= 8(16z232z+15)+qT/c2 2z2(4z220z+17)+qT/c4z4\begin{split}{\cal S}_{g}^{\rm[NLO]}(z;{\cal R}_{q_{T}/c})\,&=\,8(16z^{2}-32z+15)\\ \,&+\,{\cal R}_{q_{T}/c}^{2}\,2z^{2}(4z^{2}-20z+17)\\ \,&+\,{\cal R}_{q_{T}/c}^{4}\,z^{4}\end{split} (18)

and

𝒯g[NLO](z;qT/c)=(4+z2qT/c2)5.\begin{split}{\cal T}_{g}^{\rm[NLO]}&(z;{\cal R}_{q_{T}/c})\,=\,(4+z^{2}{\cal R}_{q_{T}/c}^{2})^{5}\;.\end{split} (19)

As in Section 3.1 (see Eq. (8)), here we set qT/c2qT 2/mc2{\cal R}_{q_{T}/c}^{2}\equiv\langle{\vec{q}}_{T}^{\;2}\rangle/m_{c}^{2}, with qT 2qT 2Ω3c=60GeV2\langle{\vec{q}}_{T}^{\;2}\rangle\equiv\langle{\vec{q}}_{T}^{\;2}\rangle_{{\rm\Omega}_{3c}}=60\,\text{GeV}^{2}. In contrast to the charm case, the LO kinematics of gluon fragmentation set the initial scale at μF,06mc\mu_{F,0}\equiv 6m_{c} (see Fig. 1, right diagram).

Similarly to the charm-induced case, we apply an analogous strategy to examine the [gΩ3c][g\to{{\rm\Omega}_{3c}}] fragmentation channel. As a preliminary illustration, we show the zz dependence of the gluon initial-scale FF adopted in the OMG3Q1.0 set. To evaluate uncertainties at the starting scale, we perform a simplified, diagonal NLO DGLAP evolution using only the PggP_{gg} kernel, following the same methodology employed in Ref. Celiberto et al. (2024a) for gluon-induced tetraquark fragmentation.

The right panel of Fig. 2 displays the zz-weighted gluon FF, evaluated at μF,0=6mc\mu_{F,0}=6m_{c}, and evolved within the range 5mc<μF<7mc5m_{c}<\mu_{F}<7m_{c}. As expected, the function vanishes at z0z\to 0 and z1z\to 1, and exhibits a pronounced peak around z0.25z\simeq 0.25. Compared to charm FF, the gluon curve is wider and slightly shifted to lower zz, reflecting a more uniform sharing of energy among final-state partons in gluon-initiated splittings.

We note that charm FF dominates by several orders of magnitude over its gluon counterpart. This suppression is not merely an artifact of the scalar diquark model, rather it reflects a structural feature of fragmentation into triply heavy baryons. While final states with an even number of heavy quarks, such as quarkonia or fully heavy tetraquarks, can naturally emerge from gluon splittings, the exclusive production of three heavy quarks in the valence state—with no accompanying light partons—poses a significant challenge for gluon fragmentation.

The scalar diquark framework, which presumes an 𝒪(αs)\mathcal{O}(\alpha_{s}) suppression compared to the quark-induced channel, emphasizes this aspect without rigidly imposing it. Instead, it offers a phenomenologically motivated picture wherein the hierarchical suppression of gluon-induced baryon production is not imposed, but rather emerges as a natural outcome of the model.

3.3 The OMG3Q1.0 functions from HF-NRevo

Refer to caption
Refer to caption
Figure 2: Initial-scale charm (left) and gluon (right) channels to Ω3c{\rm\Omega}_{3c} for our OMG3Q1.0 FFs. The shaded bands represent the effect of DGLAP evolution within the range 4mc<μF<6mc4m_{c}<\mu_{F}<6m_{c} for the charm quark and 5mc<μF<7mc5m_{c}<\mu_{F}<7m_{c} for the gluon.

To complete the construction of our OMG3Q1.0 collinear FFs for triply charmed Ω{\rm\Omega} baryons, we perform a proper DGLAP evolution of the initial-scale inputs presented before. A key distinction from light-hadron fragmentation is that in this case both the (anti-)charm and gluon channels are subject to evolution thresholds. This feature stems directly from the perturbative nature of partonic splittings into the lowest Fock state: gluon-initiated and charm-initiated processes correspond to the right and left panels of Fig. 1, respectively, and are governed by their associated SDCs.

As outlined earlier, kinematic constraints dictate that the minimum invariant mass required for quark fragmentation is 5mc5m_{c}, which we adopt as the threshold scale for the charm channel. In contrast, gluon-induced production requires a higher threshold 6mc6m_{c}, which we define as the starting scale for the (anti)charm FF. To handle the evolution while properly incorporating these thresholds, we adopt a dedicated strategy based on the recently developed HF-NRevo scheme Celiberto (2024a, 2025e, 2025f).

This framework is tailored to describe DGLAP evolution of heavy-hadron FFs starting from nonrelativistic inputs and relies on three core components: interpretation, evolution, and uncertainty quantification. The first component allows us to interpret the low-transverse-momentum production mechanism as a two-parton fragmentation process, as outlined in Section 2.1. This forms the basis for a consistent matching between FFNS and VFNS calculations. The third component provides a systematic approach to estimate the impact of missing higher-order uncertainties (MHOUs) by varying the evolution thresholds.

Although originally developed to bridge precision QCD predictions with a hadron-structure-oriented vision for quarkonium fragmentation, the HF-NRevo methodology has recently been extended to address exotic matter production, with promising results. Applications to scalar (0++0^{++}) and tensor (2++2^{++}) T4cT_{4c} states Celiberto et al. (2024a); Celiberto and Gatto (2025), as well as to rare baryons such as Ω3c{\rm\Omega}_{3c} (as explored in this work), have established HF-NRevo as a flexible framework for evolving FFs that involve both constituent heavy-quark and gluon initial-scale inputs. This dual-channel structure introduces the need for a dedicated treatment of evolution thresholds, tailored to each partonic species.

In this exploratory study on Ω3c{\rm\Omega}_{3c} fragmentation, we postpone the implementation of matching procedures and uncertainty quantification. Instead, we concentrate on developing a robust strategy for handling DGLAP evolution in the presence of quark and gluon thresholds.

According to HF-NRevo, the DGLAP evolution of heavy-hadron FFs is implemented in two successive stages. In the context of rare baryon production, we begin by evolving the channel with the lowest threshold, namely the charm-initiated one. The initial condition is provided by the (anti)charm FF at μF,0=5mc\mu_{F,0}=5m_{c}, as obtained in Section 3.1. This FF is then evolved up to 6mc6m_{c}, the gluon threshold, using only the PqqP_{qq} kernel. Since this evolution is both expanded in powers of αs\alpha_{s} and decoupled from other partonic contributions, it can be performed analytically using the symJETHAD plugin Celiberto (2021a, 2022a, 2023a, 2024b, 2024c).

The second stage begins by combining the evolved (anti)charm FF at 6mc6m_{c} with the gluon FF input introduced in Section 3.2. Starting from this common scale, we apply a full numerical DGLAP evolution to produce the NLO OMG3Q1.0 set, which we release in LHAPDF format. We define the starting point of this step as the evolution-ready scale, Q0Q_{0}, corresponding to the highest threshold among active partonic species, 6mc6m_{c} in this case. The Q0Q_{0} scale is the energy at which the numerical evolution is initialized. For this purpose, we use APFEL++ Bertone et al. (2014); Carrazza et al. (2015); Bertone (2018). In future developments, we also intend to interface with EKO Candido et al. (2022); Hekhorn and Magni (2023) to further expand our evolution technology.

One could argue that our treatment is incomplete because of the omission of light- and bottom-quark channels. To the best of our knowledge, no calculation currently exists for the collinear fragmentation of a nonconstituent quark into a triply heavy baryon. As a result, in our two-step evolution framework, light and bottom quarks are not assigned any initial-scale FFs and instead emerge dynamically through DGLAP evolution. However, drawing from analogies with NRQCD analyses of the color-singlet pseudoscalar Braaten and Yuan (1993); Braaten et al. (1993a); Artoisenet and Braaten (2015); Zhang et al. (2019); Zheng et al. (2021a, b) and vector Braaten and Yuan (1993); Braaten et al. (1993a); Zheng et al. (2019b) charmonia, these channels are expected to be strongly suppressed compared to those initiated by gluons and charm quarks.

Refer to caption
Refer to caption
Figure 3: Factorization-scale dependence of the OMG3Q1.0 FFs describing the VFNS fragmentation of the Ω3c{\rm\Omega}_{3c} baryon (left), compared with the corresponding TQ4Q1.1 FFs Celiberto and Gatto (2025); Celiberto (2025a) portraying the T4c(2++)T_{4c}(2^{++}) tetraquark (right), evaluated at z=0.5zz=0.5\simeq\langle z\rangle.

The plot on the left of Fig. 3 displays the factorization scale dependence of the OMG3Q1.0 NLO functions that describe collinear VFNS fragmentation into Ω3c{\rm\Omega}_{3c} baryons. For comparison, the panel on the right displays the corresponding μF\mu_{F} evolution of the TQ4Q1.1 NLO functions Celiberto and Gatto (2025), which describe the collinear VFNS fragmentation into the tensor T4cT_{4c} resonance (2++2^{++}), widely regarded as the leading candidate for the X(6900)X(6900) tetraquark Aaij et al. (2020). For conciseness, both plots are evaluated at a representative value of the momentum fraction, z=0.5zz=0.5\simeq\langle z\rangle, which approximates the typical region where FFs provide dominant contributions to semi-hard final states Celiberto et al. (2016a, 2017); Bolognino et al. (2018a); Celiberto (2021a); Celiberto et al. (2021a, b); Celiberto and Fucilla (2022a); Celiberto (2022b, 2023b).

A direct comparison of the results in Fig. 3 reveals that the (anti)charm-to-Ω3c{\rm\Omega}_{3c} fragmentation channel emerges as the dominant contribution. It significantly outweighs the light-parton (not shown due to their negligible size) and bottom-quark channels across the entire μF\mu_{F} range considered. Moreover, it exceeds the gluon contribution by approximately two orders of magnitude. This behavior contrasts with the case of the T4c(2++)T_{4c}(2^{++}) state, where the gluon FF dominates over the charm FF, with ratios ranging from a factor of 5 up to an order of magnitude. The prevailing role of the charm-initiated FF in Ω3c{\rm\Omega}_{3c} production directly reflects the initial-scale hierarchy between the two channels, as discussed earlier (see Fig. 2). While the gluon FF increases with μF\mu_{F} due to its feeding from timelike splittings, the charm FF exhibits a smoother trend and receives no comparable enhancement. This explains why the initial five-order hierarchy between the charm and gluon FFs is reduced to roughly two orders of magnitude after DGLAP evolution.

Despite the charm FF being substantially larger in magnitude than its gluon counterpart, the latter still plays a key role in predicting Ω3c{\rm\Omega}_{3c} (semi-)inclusive production rates at hadron colliders. This relevance stems from the overwhelming dominance of the gluon parton distribution function (PDF) over quark PDFs, which enhances the contribution of both the [gggg][gg\to gg] partonic subprocess and the gluon-initiated fragmentation. As a result, the simultaneous inclusion of initial-scale inputs for both the charm and gluon channels in our OMG3Q1.0 determinations strengthens the robustness and reliability of the proposed methodology.

Finally, we observe that the gluon-to-Ω3c{\rm\Omega}_{3c} FF exhibits a mild increase with μF\mu_{F}, while its gluon-to-T4cT{4c} counterpart displays a slow decrease. Both behaviors are smoothly varying with respect to the factorization scale, a property that carries significant phenomenological implications. Indeed, it has recently been shown that gluon FFs characterized by a smooth μF\mu_{F}-dependence act as effective “stabilizers” in high-energy resummed observables sensitive to the semi-inclusive production of singly Celiberto et al. (2021a, b) and multiply Celiberto and Fucilla (2022a); Celiberto (2022b); Celiberto and Papa (2024) heavy-flavored hadrons.

This remarkable feature has been referred to as the natural stability Celiberto (2023c) of high-energy resummation (see Section 4). The natural stability arising from the heavy flavor fragmentation will constitute the cornerstone of our forthcoming phenomenological analysis (see Section 5).

4 High-energy resummation at work

The first part of this Section (4.1) provides a concise overview of recent phenomenological advances in the exploration of the semi-hard regime of QCD. The second part (4.2) details the construction of the NLL/NLO+{\rm NLL/NLO^{+}} hybrid factorization framework and its adaptation to the semi-inclusive Ω3c{\rm\Omega}_{3c} plus jet hadroproduction process.

4.1 The semi-hard regime of QCD

The production of heavy-flavored hadrons offers a valuable perspective on high-energy QCD, where energy logarithms become large enough to challenge perturbative expansions. The Balitsky-Fadin-Kuraev-Lipatov (BFKL) approach Fadin et al. (1975); Kuraev et al. (1977); Balitsky and Lipatov (1978) addresses this by systematically resumming energy logarithms to all orders, covering both leading logarithmic (LL) terms αsnln(s)n\alpha_{s}^{n}\ln(s)^{n} and next-to-leading logarithmic (NLL) corrections αsn+1ln(s)n\alpha_{s}^{n+1}\ln(s)^{n}.

In the BFKL framework, cross sections are computed as transverse-momentum convolutions involving a universal NLO Green’s function Fadin and Lipatov (1998); Ciafaloni and Camici (1998) and process-dependent, singly off-shell emission functions, also known as forward impact factors. These functions incorporate collinear elements such as PDFs and FFs, leading to a hybrid-factorization formalism that merges high-energy and collinear QCD dynamics.

Over the years, the BFKL resummation has been applied to a variety of processes, including Mueller-Navelet jets Mueller and Navelet (1987); Ducloué et al. (2013); Colferai and Niccoli (2015); Celiberto et al. (2015a, b, 2016b); Celiberto (2017); Caporale et al. (2018); de León et al. (2021); Celiberto and Papa (2022); Baldenegro et al. (2024), di-hadron Celiberto et al. (2016a, 2017); Celiberto (2017); Celiberto et al. (2020); Celiberto (2022a) and hadron-jet systems Bolognino et al. (2018a, 2019a, 2019b); Celiberto (2021a); Celiberto et al. (2020); Mohammed (2022); Celiberto (2023b), multijets Caporale et al. (2016, 2017a); Celiberto (2016); Caporale et al. (2017b); Celiberto (2017), forward-Higgs Hentschinski et al. (2021); Celiberto et al. (2022a, 2021c); Mohammed (2022); Celiberto and Papa (2023); Celiberto et al. (2023, 2024b, 2024c, 2024d, 2022b, 2024e), Drell-Yan Celiberto et al. (2018a); Golec-Biernat et al. (2018), and heavy-flavored emissions Celiberto et al. (2018b); Boussarie et al. (2018); Bolognino et al. (2019c, d); Celiberto et al. (2021a, b); Celiberto and Fucilla (2022a); Celiberto (2023a, c); Bolognino et al. (2023); Celiberto (2022b); Celiberto and Fucilla (2022b); Celiberto (2024d). Single forward production rates have provided insights into small-xx gluon dynamics via the UGD, studied at HERA Anikin et al. (2011); Besse et al. (2013); Bolognino et al. (2018b, c, 2019e, 2020); Celiberto (2019); Bolognino (2021); Łuszczak et al. (2022), the EIC Bolognino et al. (2021a, 2022a); Bolognino (2021); Bolognino et al. (2022b, c). This has subsequently enabled the development of resummed PDFs Ball et al. (2018); Abdolmaleki et al. (2018); Bonvini and Giuli (2019); Silvetti and Bonvini (2023); Silvetti (2023); Rinaudo (2024) and improved low-xx TMDs Bacchetta et al. (2020, 2024a); Celiberto (2021b); Bacchetta et al. (2022a, b, c, d, e, f); Celiberto (2022c); Bacchetta et al. (2024b).

In the context of heavy-hadron emissions, processes like Λc{\rm\Lambda}_{c} Celiberto et al. (2021a) and bb-hadron production Celiberto et al. (2021b) have revealed mechanisms to address challenges in natural-scale descriptions of semi-hard processes. Unlike light-particle emissions, which suffer from large NLL corrections and nonresummed threshold effects Ducloué et al. (2014); Caporale et al. (2014); Bolognino et al. (2018a); Celiberto (2021a), heavy-flavored hadrons exhibit a natural stabilization trend Celiberto (2023c), driven by collinear VFNS fragmentation.

These findings have inspired further studies, including the development of VFNS DGLAP-evolving FFs based on NRQCD inputs Braaten et al. (1993a); Zheng et al. (2019b); Braaten and Yuan (1993); Chang and Chen (1992); Braaten et al. (1993b); Ma (1994); Zheng et al. (2019a, 2022); Feng et al. (2022b); Feng and Jia (2023), extending from vector quarkonia Celiberto and Fucilla (2022a); Celiberto (2023a) to Bc(1S0)B_{c}(^{1}S_{0}) and Bc(3S1)B_{c}(^{3}S_{1}) mesons Celiberto (2022b, 2024d). Natural stability has also opened a portal to exotic matter, serving as a phenomenological playground to study the fragmentation of the leading power doubly Celiberto and Papa (2024); Celiberto and Gatto (2025) or fully heavy tetraquarks Celiberto et al. (2024a); Celiberto and Gatto (2025); Celiberto (2025a) and pentacharms Celiberto (2025d).

Refer to caption
Figure 4: Pictorial representation of semi-inclusive hadroproduction of a Ω3Q{\rm\Omega}_{3Q} baryon plus a jet within the hybrid factorization. Red blobs depict collinear PDFs. The singly off-shell coefficient function embodied in the hadron (jet) emission function is portrayed by the green (violet) oval. The orange composite arrow represents the inclusive formation of a Ω3Q{\rm\Omega}_{3Q} baryon. The large blue oval blob stands for the high-energy Green’s function resumming secondary gluon emissions in the tt-channel. Diagram was realized with the JaxoDraw 2.0 code Binosi et al. (2009).

4.2 Differential cross section at NLL/NLO+

As an application to hadron-collider phenomenology, we examine the following semi-inclusive reaction (see Fig. 4)

p(Pa)p(Pb)Ω3c(q1,y1,ϕ1)+𝒳+jet(q2,y2,ϕ2),{\rm p}(P_{a})\,{\rm p}(P_{b})\to{{\rm\Omega}_{3c}}(q_{1},y_{1},\phi_{1})+{\cal X}+{\rm jet}(q_{2},y_{2},\phi_{2})\;, (20)

where a triply-charmed baryon Ω3c{\rm\Omega}_{3c} is detected with four-momentum q1q_{1}, rapidity y1y_{1}, and azimuthal angle ϕ1\phi_{1}. In addition, a light jet is tagged with four-momentum q2q_{2}, rapidity y2y_{2}, and azimuthal angle ϕ2\phi_{2}. Both objects feature high transverse momenta, such that |q1,2|ΛQCD|\vec{q}_{1,2}|\gg{\rm\Lambda}_{\rm QCD}, and a large rapidity distance, ΔYy1y2\Delta Y\equiv y_{1}-y_{2}. An undetected final-state gluon system, 𝒳{\cal X}, is produced inclusively. We express the final-state transverse momenta in the Sudakov-vector basis, defined by the parent protons’ momenta, Pa,bP_{a,b}, where Pa,b2=0P_{a,b}^{2}=0 and (PaPb)=s/2({P_{a}}\cdot{P_{b}})=s/2. This leads to the decomposition

q1,2=x1,2Pa,bq1,22x1,2sPb,a+q1,2,q_{1,2}=x_{1,2}\,P_{a,b}-\frac{q_{1,2\perp}^{2}}{x_{1,2}s}\,P_{b,a}+q_{{1,2\perp}}\;, (21)

with q1,22q1,2 2q_{1,2\perp}^{2}\equiv-\vec{q}_{1,2}^{\,2}. In the center-of-mass frame, the following relations hold between the final-state longitudinal momentum fractions, rapidities, and transverse momenta:

x1,2=|q1,2|se±y1,2,dy1,2=±dx1,2x1,2,x_{1,2}=\frac{|\vec{q}_{1,2}|}{\sqrt{s}}e^{\pm y_{1,2}}\;,\qquad{\rm d}y_{1,2}=\pm\frac{dx_{1,2}}{x_{1,2}}\;, (22)

and therefore

ΔYy1y2=lnx1x2s|q1||q2|.\qquad\Delta Y\equiv y_{1}-y_{2}=\ln\frac{x_{1}x_{2}s}{|\vec{q}_{1}||\vec{q}_{2}|}\;. (23)

In a purely collinear factorization approach, the LO differential cross section for these reactions can be written as a one-dimensional convolution involving the on-shell hard factor, proton PDFs, and Ω3c{\rm\Omega}_{3c} FFs:

dσ[collinear]LOdx1dx2d2q1d2q2=m,n=q,q¯,g01dxa01dxbfm(xa)fn(xb)\displaystyle\frac{{\rm d}\sigma^{\rm LO}_{\rm[collinear]}}{{\rm d}x_{1}{\rm d}x_{2}{\rm d}^{2}\vec{q}_{1}{\rm d}^{2}\vec{q}_{2}}=\sum_{m,n=q,{\bar{q}},g}\int_{0}^{1}{\rm d}x_{a}\!\!\int_{0}^{1}{\rm d}x_{b}\ f_{m}\left(x_{a}\right)f_{n}\left(x_{b}\right)
×x11dζζDmΩ3c(x1ζ)dσ^m,n(s^)dx1dx2d2q1d2q2,\displaystyle\quad\times\,\int_{x_{1}}^{1}\frac{{\rm d}\zeta}{\zeta}\,D^{{\rm\Omega}_{3c}}_{m}\left(\frac{x_{1}}{\zeta}\right)\frac{{\rm d}{\hat{\sigma}}_{m,n}\left(\hat{s}\right)}{{\rm d}x_{1}{\rm d}x_{2}{\rm d}^{2}\vec{q}_{1}{\rm d}^{2}\vec{q}_{2}}\;, (24)

where the indices m,nm,n sum over all partons except the (anti)top quark, which does not participate in hadronization. For simplicity, the explicit dependence on the factorization scale μF\mu_{F} is omitted in Eq. (4.2). The functions fm,n(xa,b,μF)f_{m,n}\left(x_{a,b},\mu_{F}\right) represent the proton collinear PDFs, while the DmΩ3c(x1/ζ,μF)D^{{\rm\Omega}_{3c}}_{m}\left(x_{1}/\zeta,\mu_{F}\right) functions are the Ω3c{\rm\Omega}_{3c} collinear FFs. The variables xa,bx_{a,b} correspond to the longitudinal momentum fractions of the parent partons, and ζ\zeta is the momentum fraction of the outgoing parton fragmenting into the Ω3c{\rm\Omega}_{3c} particle. Lastly, dσ^m,n(s^){\rm d}\hat{\sigma}_{m,n}\left(\hat{s}\right) is the partonic hard factor, where s^xaxbs\hat{s}\equiv x_{a}x_{b}s is the partonic center-of-mass energy squared.

Conversely, the differential cross section within our hybrid high-energy and collinear factorization approach is expressed as a transverse-momentum convolution involving the BFKL Green’s function and the two singly off-shell emission functions. We rewrite the cross section as a Fourier series of the azimuthal-angle coefficients, 𝒞k 0{\cal C}_{k\,\geq\,0}. We write

(2π)2dσdy1dy2d|q1|d|q2|dϕ1dϕ2=[𝒞0+2k=1cos(kϕ)𝒞k],\frac{(2\pi)^{2}\,{\rm d}\sigma}{{\rm d}y_{1}{\rm d}y_{2}{\rm d}|\vec{q}_{1}|{\rm d}|\vec{q}_{2}|{\rm d}\phi_{1}{\rm d}\phi_{2}}\!=\!\left[{\cal C}_{0}+2\!\sum_{k=1}^{\infty}\!\cos(k\phi)\,{\cal C}_{k}\right]\,, (25)

where ϕϕ1ϕ2π\phi\equiv\phi_{1}-\phi_{2}-\pi.

In the MS¯\overline{\rm MS} renormalization scheme and using BFKL, we obtain (see Ref. Caporale et al. (2013) for more details)

𝒞kNLL/NLO+=eΔYs+dνeΔYα¯s(μR)χNLO(k,ν){\cal C}_{k}^{\rm NLL/NLO^{+}}=\frac{e^{\Delta Y}}{s}\int_{-\infty}^{+\infty}\!\!\!{\rm d}\nu\,e^{{\Delta Y}\bar{\alpha}_{s}(\mu_{R})\chi^{\rm NLO}(k,\nu)} (26)
×αs2(μR){Ω3cNLO(k,ν,|q1|,x1)[JNLO(k,ν,|q2|,x2)]\times\,\alpha_{s}^{2}(\mu_{R})\biggl{\{}{\cal E}_{{\rm\Omega}_{3c}}^{\rm NLO}(k,\nu,|\vec{q}_{1}|,x_{1})[{\cal E}_{J}^{\rm NLO}(k,\nu,|\vec{q}_{2}|,x_{2})]^{*}
+αs2(μR)ΔYβ04πχ(k,ν)[ln(|q1||q2|)+i2ddνlnΩ3cJ]}.+\left.\alpha_{s}^{2}(\mu_{R})\Delta Y\frac{\beta_{0}}{4\pi}\,\chi(k,\nu)\left[\ln\left(|\vec{q}_{1}||\vec{q}_{2}|\right)+\frac{i}{2}\,\frac{{\rm d}}{{\rm d}\nu}\ln\frac{{\cal E}_{{\rm\Omega}_{3c}}}{{\cal E}_{J}^{*}}\right]\right\}.

In this equation, α¯s(μR)αs(μR)Nc/π\bar{\alpha}_{s}(\mu_{R})\equiv\alpha_{s}(\mu_{R})N_{c}/\pi is the QCD running coupling, with Nc=3N_{c}=3 representing the number of colors, and β0=11Nc/32nf/3\beta_{0}=11N_{c}/3-2n_{f}/3 being the first coefficient of the QCD β\beta-function. We choose a two-loop running coupling setup with αs(mZ)=0.118\alpha_{s}\left(m_{Z}\right)=0.118, and the dynamic number of flavors, nfn_{f}.

The high-energy kernel in the exponent of Eq. (26) captures the resummation of energy logarithms at NLL accuracy:

χNLO(k,ν)=χ(k,ν)+α¯sχ^(k,ν),\displaystyle\chi^{\rm NLO}(k,\nu)=\chi(k,\nu)+\bar{\alpha}_{s}\hat{\chi}(k,\nu)\;, (27)

where χ(k,ν)\chi(k,\nu) is the eigenvalue of the LO kernel:

χ(k,ν)=2γE2Re{ψ(1+k2+iν)},\displaystyle\chi\left(k,\nu\right)=-2\gamma_{\rm E}-2\,{\rm Re}\left\{\psi\left(\frac{1+k}{2}+i\nu\right)\right\}\,, (28)

with γE\gamma_{\rm E} being the Euler–Mascheroni constant and ψ(z)Γ(z)/Γ(z)\psi(z)\equiv\Gamma^{\prime}(z)/\Gamma(z) the logarithmic derivative of the Gamma function. The function χ^(k,ν)\hat{\chi}(k,\nu) in Eq. (27) represents the NLO correction to the high-energy kernel:

χ^\displaystyle\hat{\chi} (k,ν)=χ¯(k,ν)+β08Ncχ(k,ν)\displaystyle\left(k,\nu\right)=\bar{\chi}(k,\nu)+\frac{\beta_{0}}{8N_{c}}\chi(k,\nu)
×\displaystyle\times {χ(k,ν)+10/3+2ln[(μR2/(|q1||q2|)]},\displaystyle\left\{-\chi(k,\nu)+10/3+2\ln\left[\left(\mu_{R}^{2}/(|\vec{q}_{1}||\vec{q}_{2}|\right)\right]\right\}\;,

with the characteristic χ¯(k,ν)\bar{\chi}(k,\nu) function calculated in Ref. Kotikov and Lipatov (2000).

The two quantities

Ω3c,JNLO(k,ν,|q1,2|,x1,2)=Ω3c,J+αs(μR)^Ω3c,J\displaystyle{\cal E}_{{\rm\Omega}_{3c},J}^{\rm NLO}(k,\nu,|\vec{q}_{1,2}|,x_{1,2})={\cal E}_{{\rm\Omega}_{3c},J}+\alpha_{s}(\mu_{R})\,\hat{\cal E}_{{\rm\Omega}_{3c},J} (30)

represent the NLO emission functions for the rare baryon (Ω3c{\rm\Omega}_{3c}) and the light jet (JJ), respectively. These objects are calculated in Mellin space and projected onto the eigenfunctions of the LO kernel. For the Ω3c{\rm\Omega}_{3c} emission function, we adopt the NLO calculation from Ref. Ivanov and Papa (2012a). Although originally tailored for light-flavored hadron production, this computation aligns with our VFNS approach for heavy baryons, provided that transverse momenta exceed the heavy-quark thresholds relevant for DGLAP evolution.

At the LO level, the Ω3c{\rm\Omega}_{3c} emission function reads

Ω3c(k,ν,|q1|,x1)=δc|q1|2iν1x1dζζx^12iν×[ρcfg(ζ)DgΩ3c(x^)+m=q,q¯fm(ζ)DmΩ3c(x^)],\begin{split}{\cal E}_{{\rm\Omega}_{3c}}(k,\nu,|\vec{q}_{1}|,x_{1})=\delta_{c}\,|\vec{q}_{1}|^{2i\nu-1}\int_{x}^{1}\frac{{\rm d}\zeta}{\zeta}\;\hat{x}^{1-2i\nu}\\ \times\,\Big{[}\rho_{c}f_{g}(\zeta)D_{g}^{{\rm\Omega}_{3c}}\left(\hat{x}\right)+\sum_{m=q,\bar{q}}f_{m}(\zeta)D_{m}^{{\rm\Omega}_{3c}}\left(\hat{x}\right)\Big{]}\;,\end{split} (31)

where x^=x/ζ\hat{x}=x/\zeta, δc=2CF/CA\delta_{c}=2\sqrt{C_{F}/C_{A}}, and ρc=CA/CF\rho_{c}=C_{A}/C_{F}. Here, CF=(Nc21)/(2Nc)C_{F}=(N_{c}^{2}-1)/(2N_{c}) and CA=NcC_{A}=N_{c} are the Casimir factors associated with gluon emissions from quarks and gluons, respectively. The complete NLO formula for Ω3cNLO{\cal E}_{{\rm\Omega}_{3c}}^{\rm NLO} is available in Ref. Ivanov and Papa (2012a).

The LO jet emission function is expressed as

cJ(k,ν,|q2|,x)=δc|q2|2iν1[ρcfg(x)+n=q,q¯fn(x)].\!\!c_{J}(k,\nu,|\vec{q}_{2}|,x)=\delta_{c}|\vec{q}_{2}|^{2i\nu-1}\!\Big{[}\rho_{c}f_{g}(x)+\!\!\!\sum_{n=q,\bar{q}}f_{n}(x)\Big{]}\,. (32)

The corresponding NLO correction can be obtained by combining Eq. (36) of Ref. Caporale et al. (2013) with Eqs. (4.19) and (4.20) of Ref. Colferai and Niccoli (2015). These calculations are grounded in results from Refs. Ivanov and Papa (2012a, b), optimized for numerical analyses. They employ a small-cone jet selection function Furman (1982); Aversa et al. (1989) based on a cone-type algorithm Colferai and Niccoli (2015).

Equations (26), (31), and (32) clearly demonstrate the implementation of our hybrid high-energy and collinear factorization scheme. In this approach, the cross section is expressed through the BFKL formalism, with the Green’s function and emission functions as key components. The Green’s function handles the resummation of large logarithmic contributions in the high-energy limit, while the emission functions encapsulate the PDFs and FFs, effectively bridging collinear factorization with high-energy dynamics.

The ‘++’ superscript in the 𝒞kNLL/NLO+{\cal C}_{k}^{\rm NLL/NLO^{+}} label signifies that the expression of azimuthal coefficients in Eq. (26) incorporates corrections beyond the NLL accuracy. These enhancements stem from two key sources: the exponentiated NLO corrections to the high-energy kernel and the interplay between the NLO corrections to the impact factors. As a result, the azimuthal coefficients provide a more refined representation, capturing intricate effects essential for precise predictions in processes where both high-energy and collinear logarithms are influential.

By discarding all NLO contributions in Eq. (26), one retrieves the pure LL limit of the angular coefficients, thus having

𝒞kLL/LO=eΔYs+dνeΔYα¯s(μR)χ(k,ν)×αs2(μR)Ω3c(k,ν,|q1|,x1)[J(k,ν,|q2|,x2)].\begin{split}&{\cal C}_{k}^{\rm LL/LO}=\frac{e^{\Delta Y}}{s}\int_{-\infty}^{+\infty}{\rm d}\nu\,e^{{\Delta Y}\bar{\alpha}_{s}(\mu_{R})\chi(k,\nu)}\\[5.12128pt] &\hskip 4.26773pt\times\,\alpha_{s}^{2}(\mu_{R})\,{\cal E}_{{\rm\Omega}_{3c}}(k,\nu,|\vec{q}_{1}|,x_{1})[{\cal E}_{J}(k,\nu,|\vec{q}_{2}|,x_{2})]^{*}\;.\end{split} (33)

Then, to get a meaningful comparison between high-energy resummed predictions and those derived from a purely collinear, DGLAP-inspired framework, it becomes crucial to evaluate observables within both our hybrid factorization scheme and fixed-order computations. Unfortunately, current limitations mean that no numerical tools exist to calculate fixed-order distributions at NLO for inclusive semi-hard hadron plus jet production. To address this gap and quantify the influence of high-energy resummation on DGLAP predictions, we adopt an alternative strategy.

Our methodology, originally developed to investigate the angular distributions of Mueller-Navelet Celiberto et al. (2015a, b) and hadron-jet Celiberto (2021a) azimuthal distributions, involves truncating the high-energy expansion at the NLO level. This approach allows us to simulate the high-energy behavior as it would appear in a purely NLO calculation. Specifically, we achieve this by limiting the expansion of the azimuthal coefficients to 𝒪(αs3){\cal O}(\alpha_{s}^{3}), effectively constructing a high-energy fixed-order (HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}}) approximation. This approximation provides a practical framework for our phenomenological studies, facilitating a systematic comparison of the BFKL resummation effects with the high-energy regime of fixed-order predictions. The HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}} angular coefficients, computed within the MS¯\overline{\rm MS} renormalization scheme, are given by

𝒞kHE-NLO+=eΔYs+dναs2(μR)\displaystyle{\cal C}_{k}^{{\rm HE}\text{-}{\rm NLO}^{+}}=\frac{e^{\Delta Y}}{s}\int_{-\infty}^{+\infty}{\rm d}\nu\,\alpha_{s}^{2}(\mu_{R})
×[1+α¯s(μR)ΔYχ(k,ν)]\displaystyle\hskip 14.22636pt\times\,\left[1+\bar{\alpha}_{s}(\mu_{R})\Delta Y\chi(k,\nu)\right] (34)
×Ω3cNLO(k,ν,|q1|,x1)[JNLO(k,ν,|q2|,x2)].\displaystyle\hskip 14.22636pt\times\,{\cal E}_{{\rm\Omega}_{3c}}^{\rm NLO}(k,\nu,|\vec{q}_{1}|,x_{1})[{\cal E}_{J}^{\rm NLO}(k,\nu,|\vec{q}_{2}|,x_{2})]^{*}\;.

In our analysis, the factorization scale (μF\mu_{F}) and renormalization one (μR\mu_{R}) are set to natural energy values determined by the kinematics of the final state. We define μF=μRμ𝒩\mu_{F}=\mu_{R}\equiv\mu_{\cal N}, where the natural scale is μ𝒩=mΩ3c,+|q1|\mu_{\cal N}=m_{{\rm\Omega}_{3c},\perp}+|\vec{q}_{1}|. Here, mΩ3c,=mΩ3c2+|q1|2m_{{\rm\Omega}_{3c},\perp}=\sqrt{m_{{\rm\Omega}_{3c}}^{2}+|\vec{q}_{1}|^{2}} represents the transverse mass of the fully heavy baryon, and we set mΩ3c=3mcm_{{\rm\Omega}_{3c}}=3m_{c}. The transverse mass of the light jet is equal to its transverse momentum, |q2||\vec{q}_{2}|.

Although the emission of two particles naturally involves two distinct energy scales, we simplify the analysis by combining these into a single reference scale (μ𝒩\mu_{\cal N}), defined as the sum of the transverse masses of both particles. This choice aligns with strategies employed in other precision QCD calculations and codes, such as those in Refs. Alioli et al. (2010); Campbell et al. (2012); Hamilton et al. (2013). It enables a consistent comparison of our results with predictions from different approaches while adhering to common conventions in QCD phenomenology.

To investigate the impact of MHOUs, we vary both μF\mu_{F} and μR\mu_{R} within a range of μ𝒩/2\mu_{\cal N}/2 to 2μ𝒩2\mu_{\cal N}, controlled by the parameter CμC_{\mu}. This method allows us to evaluate the sensitivity of our predictions to scale variations, providing a robust estimation of theoretical uncertainties.

5 From HL-LHC to FCC

All numerical results presented in this work were obtained using the Python+Fortran JETHAD multimodular interface Celiberto (2021a, 2022a, 2023a, 2024b). For the proton PDFs, we used the NNPDF4.0 NLO determination Ball et al. (2021, 2022a), accessed through LHAPDF v6.5.5 Buckley et al. (2015).

The uncertainty bands in the plots reflect both the impact of MHOUs and errors arising from multidimensional numerical integrations, which were kept consistently below 1% by the JETHAD integrators.

5.1 Rapidity distributions

The first production rate matter of our phenomenological analysis is the rapidity distribution, namely the cross section differential in the rapidity interval, ΔY=y1y2\Delta Y=y_{1}-y_{2}, between the baryon and the jet

dσ(ΔY,s)dΔY=y1miny1maxdy1y2miny2maxdy2δ(y1y2ΔY)×|q1|min|q1|maxd|q1||q2|min|q2|maxd|q2|𝒞0[accuracy],\begin{split}\frac{{\rm d}\sigma(\Delta Y,s)}{{\rm d}\Delta Y}&=\int_{y_{1}^{\rm min}}^{y_{1}^{\rm max}}\!\!\!\!\!{\rm d}y_{1}\int_{y_{2}^{\rm min}}^{y_{2}^{\rm max}}\!\!\!\!\!{\rm d}y_{2}\,\,\delta(y_{1}-y_{2}-\Delta Y)\\[2.84544pt] &\times\,\int_{|\vec{q}_{1}|^{\rm min}}^{|\vec{q}_{1}|^{\rm max}}\!\!\!\!\!{\rm d}|\vec{q}_{1}|\int_{|\vec{q}_{2}|^{\rm min}}^{|\vec{q}_{2}|^{\rm max}}\!\!\!\!\!{\rm d}|\vec{q}_{2}|\,\,{\cal C}_{0}^{\rm[accuracy]}\;,\end{split} (35)

with 𝒞0{\cal C}_{0} being the k=0k=0, ϕ\phi-averaged angular coefficient given in Section 4.2. In this context, the ‘[accuracy]{\rm[accuracy]}’ superscript includes NLL/NLO+{\rm NLL/NLO^{+}}, LL/LO{\rm LL/LO}, or HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}}.

The transverse momenta of the heavy baryon span from 6060 to 120120 GeV, while those of the jet range from 5050 to 120120 GeV. These ranges align with current and prospective analyses of jets and hadrons at the LHC Khachatryan et al. (2016, 2021). Using asymmetric windows for the observed transverse momenta enhances the prominence of the high-energy signal over the fixed-order background Celiberto et al. (2015a, b); Celiberto (2021a).

Our choice of rapidity intervals aligns with established criteria in ongoing LHC studies. Baryon detections, restricted to the barrel calorimeter as in the CMS experiment Chatrchyan et al. (2012), are limited to the rapidity range |y1|<2.5|y_{1}|<2.5. For jets, which can also be tagged in the endcap regions Khachatryan et al. (2016), we consider a broader range of rapidity of |y2|<4.7|y_{2}|<4.7.

Figure 5 presents our resummed predictions for Ω3c{\rm\Omega}_{3c} plus jet ΔY\Delta Y distributions at s=14\sqrt{s}=14 TeV (HL-LHC, left panel) and s=100\sqrt{s}=100 TeV (FCC, right panel). To facilitate meaningful comparisons with prospective experimental measurements, we adopt uniformly spaced ΔY\Delta Y bins of fixed width, ΔY=0.5\Delta Y=0.5. Ancillary panels beneath the main plots show ratios of the LL/LO{\rm LL/LO} and HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}} predictions to the NLL/NLO+{\rm NLL/NLO^{+}} baseline.

The resulting cross sections are encouraging, spanning roughly from 10210^{2} pb down to 10210^{-2} pb. Notably, a more than tenfold increase is observed when moving from HL-LHC to FCC energies. We emphasize that the vertical scales of the two panels differ accordingly. All ΔY\Delta Y spectra display a monotonic decreasing behavior with increasing ΔY\Delta Y. This trend reflects the interplay of two competing mechanisms. The partonic hard factor grows with energy (and thus with ΔY\Delta Y), in line with expectations from high-energy resummation. Conversely, this growth is counteracted by suppression due to the collinear convolution with PDFs and FFs in the emission functions (see Eqs. (31) and (32)).

Refer to caption
Refer to caption
Figure 5: ΔY\Delta Y differential distribution for semi-inclusive Ω3c{\rm\Omega}_{3c} plus jet detections at s=14\sqrt{s}=14 TeV (HL-LHC, left) and 100100 TeV (nominal FCC, right). NNPDF4.0 NLO proton PDFs Ball et al. (2021, 2022a) are used in combination with OMG3Q1.0 NLO heavy-baryon FFs Celiberto (2025b). Ancillary panels below the main plots show the ratio of LL/LO{\rm LL/LO} and HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}} predictions to NLL/NLO+{\rm NLL/NLO^{+}}. Uncertainty bands account for the combined effects of MHOUs and numerical phase-space integration.
Refer to caption
Refer to caption
Figure 6: ΔY\Delta Y differential distribution for semi-inclusive T4c(2++)T_{4c}(2^{++}) plus jet detections at s=14\sqrt{s}=14 TeV (HL-LHC, left) and 100100 TeV (nominal FCC, right). NNPDF4.0 NLO proton PDFs Ball et al. (2021, 2022a) are used in combination with TQ4Q1.1 NLO heavy-tetraquark FFs Celiberto and Gatto (2024a); Celiberto (2025c). Ancillary panels below the main plots display the ratio of LL/LO{\rm LL/LO} and HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}} predictions to NLL/NLO+{\rm NLL/NLO^{+}}. Uncertainty bands reflect the combined effect of MHOUs and numerical integration over phase space.

Figure 6 shows the corresponding ΔY\Delta Y rates for the tensor T4c(2++)T_{4c}(2^{++}) resonance, the most favored candidate for the X(6900)X(6900) state Aaij et al. (2020). These plots allow for a direct comparison between two different scenarios of fully heavy hadron production—one baryonic, the other tetraquark—in the semi-hard regime of QCD. Although both analyses share the same kinematic cuts and explore the same theoretical accuracies, the behavior of the observables is markedly different in terms of both intensity and sensitivity to the collider energy.

The Ω3c{\rm\Omega}_{3c} channel exhibits significantly larger cross sections, compared to the much lower rates observed for T4cT_{4c}. Moreover, the increase in rate from 14 TeV to 100 TeV is more pronounced for the baryon, amounting to over two orders of magnitude. This confirms that Ω3c{\rm\Omega}_{3c} production benefits more strongly from higher center-of-mass energies, and highlights the dual phenomenological character of the process at high energy.

On one hand, the LL uncertainty bands for Ω3c{\rm\Omega}_{3c} are more widely separated from the NLL bands than in the tetraquark case. They approach the full resummed prediction only gradually as s\sqrt{s} increases, indicating that the emergence of predictive stability under MHOUs and higher-order corrections requires higher energies for the baryon case. This is in contrast to T4cT_{4c}, where LL and NLL bands are already closely aligned at 14 TeV, suggesting a milder sensitivity to high-energy dynamics. In this sense, both HL-LHC and FCC offer relevant but distinct opportunities for future studies of Ω3c{\rm\Omega}_{3c} plus jet emissions: the former highlights the onset of resummation effects, while the latter enables their full manifestation.

On the other hand, the [Ω3c+jet][{\rm\Omega}_{3c}+\text{jet}] process turns out to be much more effective than its tetraquark counterpart at discriminating the resummed signal from its fixed-order limit. This is a particularly important and novel result, as rapidity-interval distributions are generally considered less suited to highlighting resummation structures than transverse-momentum spectra. Yet in this case, the semi-hard production of a rare, fully heavy baryon proves to be remarkably sensitive to high-energy dynamics, setting it apart as a unique and promising probe of QCD behavior in the forward region. This distinctive feature underscores the relevance of Ω3c{\rm\Omega}_{3c} studies as a benchmark channel for future experimental and theoretical investigations.

5.2 Transverse-momentum distributions

The second production rate matter of our phenomenological analysis is the transverse-momentum distribution

dσ(|q1|,s)d|q1|=ΔYminΔYmaxdΔYy1miny1maxdy1y2miny2maxdy2×δ(y1y2ΔY)|q2|min|q2|maxd|q2|𝒞0[accuracy].\begin{split}&\frac{{\rm d}\sigma(|\vec{q}_{1}|,s)}{{\rm d}|\vec{q}_{1}|}=\int_{\Delta Y^{\rm min}}^{\Delta Y^{\rm max}}{\rm d}\Delta Y\int_{y_{1}^{\rm min}}^{y_{1}^{\rm max}}{\rm d}y_{1}\int_{y_{2}^{\rm min}}^{y_{2}^{\rm max}}{\rm d}y_{2}\\[2.84544pt] &\hskip 15.6491pt\times\,\delta(y_{1}-y_{2}-\Delta Y)\int_{|\vec{q}_{2}|^{\rm min}}^{|\vec{q}_{2}|^{\rm max}}\!\!{\rm d}|\vec{q}_{2}|\,\,{\cal C}_{0}^{\rm[accuracy]}\;.\end{split} (36)

Being differential in |q1||\vec{q}_{1}|, this observable permits us to focus on the transverse-momentum spectrum of Ω3c{\rm\Omega}_{3c}, while the jet transverse momentum is integrated in the range 40GeV<|q2|<120GeV40~{}\text{GeV}<|\vec{q}_{2}|<120~{}\text{GeV}, and ΔY\Delta Y stays in a given bin of length ΔYmaxΔYmin\Delta Y^{\rm max}-\Delta Y^{\rm min}. The rapidity windows for both baryon and jet detections remain consistent with previous criteria.

Examining the |q1||\vec{q}_{1}| spectrum sets the groundwork for investigating potential links between the NLL/NLO+{\rm NLL/NLO^{+}} factorization approach and other theoretical frameworks. Allowing |q1||\vec{q}_{1}| to vary from 1010 to 100GeV100~{}\text{GeV} enables the exploration of a broad kinematic domain where additional resummation effects may become relevant. Recent studies on high-energy Higgs Celiberto et al. (2021c) and heavy-jet Bolognino et al. (2021b) tagging have demonstrated that our hybrid methodology is effective within the moderate transverse momentum regime, particularly when |q1||\vec{q}_{1}| and |q2||\vec{q}_{2}| are of comparable magnitude.

In contrast, higher values of |q1||\vec{q}_{1}| are expected to amplify threshold logarithms in the high-momentum domain (|q1||q2|max|\vec{q}_{1}|\gg|\vec{q}_{2}|^{\rm max}), while soft logarithms become significant in the low-momentum region (|q1||q2|min|\vec{q}_{1}|\ll|\vec{q}_{2}|^{\rm min}). Therefore, analyzing |q1||\vec{q}_{1}| distributions not only serves as validation for our framework, but also effectively prepares for the integration of additional resummation techniques.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Transverse-momentum distribution for semi-inclusive Ω3c{\rm\Omega}_{3c} plus jet detections at s=14\sqrt{s}=14 TeV (HL-LHC, left) and 100100 TeV (nominal FCC, right). The rapidity interval lies within the range 2<ΔY<42<\Delta Y<4 (upper panels) or 4<ΔY<64<\Delta Y<6 (lower panels). NNPDF4.0 NLO proton PDFs Ball et al. (2021, 2022a) are used in combination with OMG3Q1.0 NLO heavy-baryon FFs Celiberto (2025b). Ancillary panels below the main plots show the ratio of LL/LO{\rm LL/LO} and HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}} predictions to NLL/NLO+{\rm NLL/NLO^{+}}. Uncertainty bands account for the combined effects of MHOUs and numerical phase-space integration.

We present in Fig. 7 the transverse-momentum distributions of the Ω3c{\rm\Omega}{3c} baryon, produced in semi-inclusive Ω3c{\rm\Omega}{3c} plus jet events at s=14\sqrt{s}=14 TeV (HL-LHC, left) and 100100 TeV (nominal FCC, right). Results are shown for two representative rapidity intervals, 2<ΔY<42<\Delta Y<4 (top panels) and 4<ΔY<64<\Delta Y<6 (bottom panels), with each spectrum evaluated in uniform |q1||\vec{q}_{1}| bins of 1010 GeV. As expected, the overall cross sections are higher in the FCC configuration due to the extended phase space reach, and decrease when ΔY\Delta Y spans over larger values, as expected.

The |q1||\vec{q}_{1}| distributions exhibit a smooth and monotonic decrease as |q1||\vec{q}_{1}| increases, across both collider energies and ΔY\Delta Y bins. No pronounced peak is observed; instead, the cross section steadily falls from the lowest transverse-momentum values onward. This behavior is consistent with expectations from collinear QCD at moderate |q1||\vec{q}_{1}|, where the hybrid resummation framework captures the dominant logarithmic enhancements.

At 1414 TeV, the NLL/NLO+{\rm NLL/NLO^{+}} predictions show a sharper drop, reflecting stronger suppression at high |q1||\vec{q}_{1}| due to limited phase space and reduced parton luminosities. At 100100 TeV, the distributions become broader and flatter, extending more visibly into the high-|q1||\vec{q}_{1}| region thanks to enhanced partonic fluxes. In all panels, the separation between the NLL signal and the two background approximations (LL and high-energy fixed order) becomes more pronounced with increasing |q1||\vec{q}_{1}|, especially at FCC energies. This confirms the value of transverse-momentum observables as effective probes for assessing the stability and discriminatory power of high-energy resummation in rare baryon production.

In particular, the growing discrepancy between the NLL/NLO+{\rm NLL/NLO^{+}} and HE-NLO+{\rm HE}\mbox{-}{\rm NLO^{+}} curves at larger |q1||\vec{q}_{1}| is not unexpected. While BFKL resummation is well-suited to configurations where |q1||q2||\vec{q}_{1}|\simeq|\vec{q}_{2}|, the regime |q1||q2||\vec{q}_{1}|\gg|\vec{q}_{2}| falls outside its natural domain of applicability. In this region, large logarithms of DGLAP type, together with potential threshold logarithms, become relevant, indicating the need to enhance our theoretical framework through alternative and dedicated resummation strategies tailored to these kinematic limits.

Furthermore, the behavior of the LL/LO{\rm LL/LO} versus NLL/NLO+{\rm NLL/NLO^{+}} ratio reflects a complex interplay of competing NLO effects. While jet functions typically receive negative NLO corrections Bartels et al. (2002); Ivanov and Papa (2012b); Colferai and Niccoli (2015), the hadron side behaves differently: the CggC_{gg} coefficient is corrected positively, whereas other terms yield negative shifts Ivanov and Papa (2012a). Such opposing trends can partially cancel, depending on the kinematic region, modulating the LL/LO{\rm LL/LO} over NLL/NLO+{\rm NLL/NLO^{+}} behavior. For instance, in cascade-baryon plus jet production, the ratio often exceeds unity Celiberto (2023b), unlike in doubly charmed tetraquark production Celiberto and Papa (2024), where the gap is milder. These variations highlight how process-dependent emission dynamics shape the relative weight of LL versus NLL contributions.

To conclude, the fair stability of our predictions under MHOU variations, as illustrated in Figs. 7, positions the transverse-momentum spectrum of rare-baryon plus jet events as a sensitive probe of high-energy QCD mechanisms. This robustness, rooted in the VFNS fragmentation of Ω3c{\rm\Omega}_{3c}, remains evident not only at current LHC scales but also at the higher energies envisioned for the FCC.

6 Final remarks

We explored the leading-power fragmentation of fully charmed baryons, focusing on the rare Ω3c{\rm\Omega}_{3c} sector at present and future hadron colliders. To this aim, we released a new set of hadron-structure-oriented collinear FFs in a VFNS, labeled OMG3Q1.0. Our formulation builds on a diquark-inspired proxy model, which parallels the standard structure commonly employed in the fragmentation modeling of both singly heavy hadrons and heavy quarkonia. In this picture, FFs are modeled as convolutions between perturbative SDCs, describing the splitting of an outgoing high-energy parton into the valence Fock state, and a purely nonperturbative component that encapsulates the hadronization process into the observed hadron.

A key feature of our method lies in the simultaneous inclusion of both charm quark and gluon channels at the initial scale. This dual channel structure makes Ω3c{\rm\Omega}_{3c} baryons an ideal testing ground for the HF-NRevo scheme Celiberto (2024a, 2025e, 2025f). Indeed, the presence of heavy partonic thresholds necessitates a dedicated DGLAP evolution strategy that treats the charm and gluon channels consistently. By implementing a two-step evolution process—analytic below and numerical above the gluon threshold—we ensure full control over the matching between partonic sectors and the proper enforcement of threshold effects.

We evaluated the phenomenological implications of our approach using the (sym)JETHAD framework. In particular, we investigated the semi-inclusive production of Ω3c{\rm\Omega}_{3c} plus jet systems within a hybrid NLL/NLO+{\rm NLL/NLO^{+}} high-energy resummation scheme. Our results, validated across HL-LHC and FCC energy scales, exhibit a progressively increasing natural stability Celiberto (2023c), directly stemming from the collinear VFNS fragmentation of rare Ω{\rm\Omega} baryons. This stability reinforces the robustness of our formalism and underscores the predictive power of our FF determinations.

Remarkably, our phenomenological analysis has revealed the emergence of novel features. Among these, the rapidity-interval distributions stand out for their unprecedented ability to discriminate between the NLL resummed signal and the high-energy NLO background. Such discriminating power has not been previously observed in the context of two-object semi-inclusive hadroproduction within the semi-hard regime of QCD. These findings point to a rich, nontrivial, and compelling interplay between heavy-flavor fragmentation and high-energy resummation, which opens promising avenues for targeted future studies.

Looking ahead, our goal is to enhance our framework by extending the HF-NRevo scheme with systematic uncertainty quantification, potentially related to MHOU effects Kassabov et al. (2023); Harland-Lang and Thorne (2019); Ball and Pearson (2021); McGowan et al. (2023); Ball et al. (2024a); Pasquini et al. (2023). Moreover, the inclusion of additional fragmentation channels, such as those initiated by light and bottom quarks, will allow for a more complete implementation of threshold-sensitive evolution, offering further insight into the production mechanisms of rare baryons.

Heavy-flavor fragmentation remains one of the most promising intersections between hadronic structure and precision QCD. Its relevance is amplified when studying exotic systems with complex internal configurations, such as Ω3c{\rm\Omega}_{3c}. The unique interplay of perturbative and nonperturbative dynamics required to describe such objects deepens our understanding of hadron formation mechanisms and opens new avenues for exploring strong-interaction phenomena at high energy.

An intriguing prospect concerns the potential of rare Ω\rm\Omega baryons to serve as indirect probes of intrinsic charm content in the proton Brodsky et al. (1980, 2015); Ball et al. (2022b); Guzzi et al. (2023). Due to their fully charmed valence structure, states like Ω3c{\rm\Omega}_{3c} provide a unique opportunity to isolate and enhance the sensitivity to initial charm contributions in high-energy collisions. While their production is largely influenced by gluon-initiated processes (especially at small and moderate Bjorken-xx), certain kinematic regions, such as forward rapidities and intermediate-to-large momentum fractions, may enhance charm-initiated channels, thereby exposing potential valence-like components in the proton wavefunction Ball et al. (2024b).

The simultaneous presence of charm and gluon initial-scale FF inputs in our OMG3Q1.0 determinations ensures that such channels are consistently treated within the HF-NRevo scheme. Despite the statistically challenging nature of Ω3c{\rm\Omega}_{3c} production, any anomalous enhancement in data compared to gluon-driven predictions may hint at intrinsic charm dynamics. In this sense, rare Ω\rm\Omega baryons could complement more conventional observables, such as [γ+c][\gamma+c] or [J/ψ+c][J/\psi+c] systems Flore et al. (2020), in mapping the charm content of the nucleon and in exploring its possible connection to the structure of exotic hadrons.

Our findings strengthen the case for rare baryons as precision and exploratory tools in QCD studies. They pave the way for future experimental searches and provide a solid theoretical baseline for shedding light on the core structure hadrons, particularly in view of the far-reaching potential offered by forthcoming high-energy collider programs Chapon et al. (2022); Accardi et al. (2025); Abdul Khalek et al. (2022a, b); Hentschinski et al. (2023); Amoroso et al. (2022); Abir et al. (2023); Allaire et al. (2024); Anchordoqui et al. (2022); Feng et al. (2023); Adachi et al. (2022); Balazs et al. (2025); Attié et al. (2025); Arbuzov et al. (2021); Accettura et al. (2023, 2024a, 2024b); Black et al. (2024); Accettura et al. (2025); Accardi et al. (2024).

Acknowledgments

The author thanks Alessandro Pilloni and Marco Bonvini for fruitful conversations. The present study received support from the Atracción de Talento Grant n. 2022-T1/TIC-24176 of the Comunidad Autónoma de Madrid (Spain).

Data availability

The LHAPDF version of OMG3Q1.0 collinear FFs Celiberto (2025b) for fully charmed Ω\rm\Omega baryons are publicly available. They can be downloaded from the following url: https://github.com/FGCeliberto/Collinear_FFs/.

References

  • Bjorken (1985) J. D. Bjorken, AIP Conf. Proc. 132, 390 (1985).
  • Fleck and Richard (1989) S. Fleck and J. M. Richard, Prog. Theor. Phys. 82, 760 (1989).
  • Martynenko (2008) A. P. Martynenko, Phys. Lett. B 663, 317 (2008), 0708.2033.
  • Martynenko and Trunin (2014) A. P. Martynenko and A. M. Trunin, Phys. Rev. D 89, 014004 (2014), 1308.3998.
  • Karliner and Rosner (2014) M. Karliner and J. L. Rosner, Phys. Rev. D 90, 094007 (2014), 1408.5877.
  • Yoshida et al. (2015) T. Yoshida, E. Hiyama, A. Hosaka, M. Oka, and K. Sadato, Phys. Rev. D 92, 114029 (2015), 1510.01067.
  • Ebert et al. (2002) D. Ebert, R. N. Faustov, V. O. Galkin, and A. P. Martynenko, Phys. Rev. D 66, 014008 (2002), hep-ph/0201217.
  • Roberts and Pervin (2008) W. Roberts and M. Pervin, Int. J. Mod. Phys. A 23, 2817 (2008), 0711.2492.
  • Chen and Wu (2011) Y.-Q. Chen and S.-Z. Wu, JHEP 08, 144 (2011), [Erratum: JHEP 09, 089 (2011)], 1106.0193.
  • Padmanath et al. (2014) M. Padmanath, R. G. Edwards, N. Mathur, and M. Peardon, Phys. Rev. D 90, 074504 (2014), 1307.7022.
  • Brown et al. (2014) Z. S. Brown, W. Detmold, S. Meinel, and K. Orginos, Phys. Rev. D 90, 094507 (2014), 1409.0497.
  • Meinel (2012) S. Meinel, Phys. Rev. D 85, 114510 (2012), 1202.1312.
  • Flynn et al. (2003) J. M. Flynn, F. Mescia, and A. S. B. Tariq (UKQCD), JHEP 07, 066 (2003), hep-lat/0307025.
  • Shah et al. (2016) Z. Shah, K. Thakkar, and A. K. Rai, Eur. Phys. J. C 76, 530 (2016), 1609.03030.
  • Shah and Rai (2017) Z. Shah and A. K. Rai, Eur. Phys. J. A 53, 195 (2017).
  • Bagan et al. (1994) E. Bagan, H. G. Dosch, P. Gosdzinsky, S. Narison, and J. M. Richard, Z. Phys. C 64, 57 (1994), hep-ph/9403208.
  • Chang et al. (2006) C.-H. Chang, C.-F. Qiao, J.-X. Wang, and X.-G. Wu, Phys. Rev. D 73, 094022 (2006), hep-ph/0601032.
  • Gomshi Nobary and Sepahvand (2006) M. A. Gomshi Nobary and R. Sepahvand, Nucl. Phys. B 741, 34 (2006), hep-ph/0508115.
  • Gershtein et al. (2000) S. S. Gershtein, V. V. Kiselev, A. K. Likhoded, and A. I. Onishchenko, Phys. Rev. D 62, 054021 (2000).
  • Aaij et al. (2015a) R. Aaij et al. (LHCb), Phys. Rev. Lett. 115, 072001 (2015a), 1507.03414.
  • Aaij et al. (2019) R. Aaij et al. (LHCb), Phys. Rev. Lett. 122, 222001 (2019), 1904.03947.
  • Aaij et al. (2021a) R. Aaij et al. (LHCb), Sci. Bull. 66, 1278 (2021a), 2012.10380.
  • Brambilla et al. (2020) N. Brambilla, S. Eidelman, C. Hanhart, A. Nefediev, C.-P. Shen, C. E. Thomas, A. Vairo, and C.-Z. Yuan, Phys. Rept. 873, 1 (2020), 1907.07583.
  • Esposito et al. (2017) A. Esposito, A. Pilloni, and A. D. Polosa, Phys. Rept. 668, 1 (2017), 1611.07920.
  • Lebed et al. (2017) R. F. Lebed, R. E. Mitchell, and E. S. Swanson, Prog. Part. Nucl. Phys. 93, 143 (2017), 1610.04528.
  • Faustov and Galkin (2022) R. N. Faustov and V. O. Galkin, Phys. Rev. D 105, 014013 (2022), 2111.07702.
  • Mathur and Padmanath (2019) N. Mathur and M. Padmanath, Phys. Rev. D 99, 031501 (2019), 1807.00174.
  • Francis et al. (2019) A. Francis, R. J. Hudspith, R. Lewis, and K. Maltman, Evidence for charm-bottom tetraquarks and the mass dependence of heavy-light tetraquark states from lattice QCD (2019), 1810.10550.
  • Eichten and Quigg (1994) E. J. Eichten and C. Quigg, Phys. Rev. D 49, 5845 (1994), hep-ph/9402210.
  • Godfrey and Isgur (1985) S. Godfrey and N. Isgur, Phys. Rev. D 32, 189 (1985).
  • Ali et al. (2017) A. Ali, J. S. Lange, and S. Stone, Prog. Part. Nucl. Phys. 97, 123 (2017), 1706.00610.
  • Aaij et al. (2020) R. Aaij et al. (LHCb), Sci. Bull. 65, 1983 (2020), 2006.16957.
  • Karliner and Rosner (2020) M. Karliner and J. L. Rosner, Phys. Rev. D 102, 094016 (2020), 2008.05993.
  • Pineda (2012) A. Pineda, Prog. Part. Nucl. Phys. 67, 735 (2012), 1111.0165.
  • Celiberto et al. (2024a) F. G. Celiberto, G. Gatto, and A. Papa, Eur. Phys. J. C 84, 1071 (2024a), 2405.14773.
  • Celiberto and Gatto (2025) F. G. Celiberto and G. Gatto, Phys. Rev. D 111, 034037 (2025), 2412.10549.
  • Celiberto (2025a) F. G. Celiberto (2025a), 2504.03949.
  • Llanes-Estrada et al. (2012) F. J. Llanes-Estrada, O. I. Pavlova, and R. Williams, Eur. Phys. J. C 72, 2019 (2012), 1111.7087.
  • Wei et al. (2017) K.-W. Wei, B. Chen, N. Liu, Q.-Q. Wang, and X.-H. Guo, Phys. Rev. D 95, 116005 (2017), 1609.02512.
  • Yang et al. (2020) G. Yang, J. Ping, P. G. Ortega, and J. Segovia, Chin. Phys. C 44, 023102 (2020), 1904.10166.
  • Gómez-Rocha et al. (2023) M. Gómez-Rocha, J. More, and K. Serafin, Few Body Syst. 64, 44 (2023), 2305.06728.
  • Najjar et al. (2024) Z. R. Najjar, K. Azizi, and H. R. Moshfegh, Eur. Phys. J. C 84, 612 (2024), 2402.14348.
  • de Arenaza et al. (2024) N. M. de Arenaza, J. J. Gálvez-Viruet, and F. J. Llanes-Estrada (2024), 2407.07232.
  • Apollinari et al. (2015) G. Apollinari, O. Brüning, T. Nakamoto, and L. Rossi, CERN Yellow Rep. pp. 1–19 (2015), 1705.08830.
  • Benedikt et al. (2025a) M. Benedikt et al. (FCC) (2025a), 2505.00272.
  • Benedikt et al. (2025b) M. Benedikt et al. (FCC) (2025b), 2505.00274.
  • Benedikt et al. (2025c) M. Benedikt et al. (FCC) (2025c), 2505.00273.
  • Abe et al. (2010) T. Abe et al. (Belle-II) (2010), 1011.0352.
  • Aaboud et al. (2016) M. Aaboud et al. (ATLAS), Phys. Rev. D 94, 032005 (2016), 1604.07773.
  • Hayrapetyan et al. (2024a) A. Hayrapetyan et al. (CMS), JINST 19, P11021 (2024a), 2410.17038.
  • Aaij et al. (2021b) R. Aaij et al. (LHCb), JHEP 02, 023 (2021b), 2010.09437.
  • Braaten et al. (1995) E. Braaten, K.-m. Cheung, S. Fleming, and T. C. Yuan, Phys. Rev. D 51, 4819 (1995), hep-ph/9409316.
  • Braaten and Yuan (1993) E. Braaten and T. C. Yuan, Phys. Rev. Lett. 71, 1673 (1993), hep-ph/9303205.
  • Braaten et al. (1993a) E. Braaten, K.-m. Cheung, and T. C. Yuan, Phys. Rev. D 48, 4230 (1993a), hep-ph/9302307.
  • Braaten et al. (1994) E. Braaten, M. A. Doncheski, S. Fleming, and M. L. Mangano, Phys. Lett. B 333, 548 (1994), hep-ph/9405407.
  • Braaten et al. (1993b) E. Braaten, K.-m. Cheung, and T. C. Yuan, Phys. Rev. D 48, R5049 (1993b), hep-ph/9305206.
  • Kiselev et al. (1994) V. V. Kiselev, A. K. Likhoded, and M. V. Shevlyagin, Phys. Lett. B 332, 411 (1994), hep-ph/9408407.
  • Anselmino et al. (1993) M. Anselmino, E. Predazzi, S. Ekelin, S. Fredriksson, and D. B. Lichtenberg, Rev. Mod. Phys. 65, 1199 (1993).
  • Ebert et al. (1996) D. Ebert, T. Feldmann, C. Kettner, and H. Reinhardt, Z. Phys. C 71, 329 (1996), hep-ph/9506298.
  • Moosavi Nejad and Sartipi Yarahmadi (2016) S. M. Moosavi Nejad and P. Sartipi Yarahmadi, Eur. Phys. J. A 52, 315 (2016), 1609.07422.
  • Caswell and Lepage (1986) W. E. Caswell and G. P. Lepage, Phys. Lett. B 167, 437 (1986).
  • Bodwin et al. (1995) G. T. Bodwin, E. Braaten, and G. P. Lepage, Phys. Rev. D 51, 1125 (1995), [Erratum: Phys.Rev.D 55, 5853 (1997)], hep-ph/9407339.
  • Cho and Leibovich (1996a) P. L. Cho and A. K. Leibovich, Phys. Rev. D 53, 150 (1996a), hep-ph/9505329.
  • Cho and Leibovich (1996b) P. L. Cho and A. K. Leibovich, Phys. Rev. D 53, 6203 (1996b), hep-ph/9511315.
  • Bodwin et al. (2005) G. T. Bodwin, E. Braaten, and J. Lee, Phys. Rev. D 72, 014004 (2005), hep-ph/0504014.
  • Adamov and Goldstein (1997) A. D. Adamov and G. R. Goldstein, Phys. Rev. D 56, 7381 (1997), hep-ph/9706491.
  • Yang (2002) J.-J. Yang, Phys. Rev. D 65, 094035 (2002).
  • Gomshi Nobary and Sepahvand (2005) M. A. Gomshi Nobary and R. Sepahvand, Phys. Rev. D 71, 034024 (2005), hep-ph/0406148.
  • Moosavi Nejad and Delpasand (2017) S. M. Moosavi Nejad and M. Delpasand, Eur. Phys. J. A 53, 174 (2017).
  • Moosavi Nejad (2017) S. M. Moosavi Nejad, Phys. Rev. D 96, 114021 (2017).
  • Delpasand and Moosavi Nejad (2019) M. Delpasand and S. M. Moosavi Nejad, Phys. Rev. D 99, 114028 (2019).
  • Cacciari and Greco (1994) M. Cacciari and M. Greco, Nucl. Phys. B 421, 530 (1994), hep-ph/9311260.
  • Buza et al. (1998) M. Buza, Y. Matiounine, J. Smith, and W. L. van Neerven, Eur. Phys. J. C 1, 301 (1998), hep-ph/9612398.
  • Cacciari and Catani (2001) M. Cacciari and S. Catani, Nucl. Phys. B 617, 253 (2001), hep-ph/0107138.
  • Mitov and Moch (2006) A. Mitov and S.-O. Moch, Nucl. Phys. B 751, 18 (2006), hep-ph/0604160.
  • An et al. (2019) H.-T. An, Q.-S. Zhou, Z.-W. Liu, Y.-R. Liu, and X. Liu, Phys. Rev. D 100, 056004 (2019), 1905.07858.
  • Ortiz-Pacheco and Bijker (2023) E. Ortiz-Pacheco and R. Bijker, Phys. Rev. D 108, 054014 (2023), 2307.04939.
  • Liu et al. (2024) C.-L. Liu, W.-X. Zhang, and D. Jia, Chin. Phys. C 48, 103110 (2024), 2403.13456.
  • Chen et al. (2016) H.-X. Chen, W. Chen, X. Liu, and S.-L. Zhu, Phys. Rept. 639, 1 (2016), 1601.02092.
  • Wang and Xu (2018) W. Wang and J. Xu, Phys. Rev. D 97, 093007 (2018), 1803.01476.
  • Celiberto (2025b) F. G. Celiberto, OMG3Q1.0: OMeGa baryons with 3 heavy Quarks VFNS FFs (2025b), URL {https://github.com/FGCeliberto/Collinear_FFs/}.
  • Celiberto and Gatto (2024a) F. G. Celiberto and G. Gatto, TQ4Q1.1: TetraQuarks with 4 heavy Quarks VFNS FFs (2024a), URL {https://github.com/FGCeliberto/Collinear_FFs/}.
  • Celiberto and Gatto (2024b) F. G. Celiberto and G. Gatto, TQHL1.1: TetraQuarks with Heavy and Light flavor collinear VFNS FFs (2024b), URL {https://github.com/FGCeliberto/Collinear_FFs/}.
  • Celiberto (2025c) F. G. Celiberto, TQ4Q1.1: axial-vector TetraQuarks with 4 heavy Quarks VFNS FFs (2025c), URL {https://github.com/FGCeliberto/Collinear_FFs/}.
  • Celiberto (2025d) F. G. Celiberto (2025d), 2502.11136.
  • Mele and Nason (1991) B. Mele and P. Nason, Nucl. Phys. B 361, 626 (1991), [Erratum: Nucl.Phys.B 921, 841–842 (2017)].
  • Celiberto (2024a) F. G. Celiberto, in 58th Rencontres de Moriond on QCD and High Energy Interactions (2024a), 2405.08221.
  • Celiberto (2025e) F. G. Celiberto, PoS DIS2024, 168 (2025e), 2406.10779.
  • Celiberto (2025f) F. G. Celiberto, Acta Phys. Polon. Supp. 18, 1 (2025f), 2412.05661.
  • Buckley et al. (2015) A. Buckley, J. Ferrando, S. Lloyd, K. Nordström, B. Page, M. Rüfenacht, M. Schönherr, and G. Watt, Eur. Phys. J. C 75, 132 (2015), 1412.7420.
  • Celiberto (2021a) F. G. Celiberto, Eur. Phys. J. C 81, 691 (2021a), 2008.07378.
  • Celiberto (2022a) F. G. Celiberto, Phys. Rev. D 105, 114008 (2022a), 2204.06497.
  • Celiberto (2023a) F. G. Celiberto, Universe 9, 324 (2023a), 2305.14295.
  • Celiberto (2024b) F. G. Celiberto, Symmetry 16, 550 (2024b), 2403.15639.
  • Celiberto (2024c) F. G. Celiberto, Particles 7, 502 (2024c), 2405.09526.
  • Cacciari et al. (1997) M. Cacciari, M. Greco, S. Rolli, and A. Tanzini, Phys. Rev. D 55, 2736 (1997), hep-ph/9608213.
  • Jaffe and Randall (1994) R. L. Jaffe and L. Randall, Nucl. Phys. B 412, 79 (1994), hep-ph/9306201.
  • Kniehl et al. (2005) B. A. Kniehl, G. Kramer, I. Schienbein, and H. Spiesberger, Eur. Phys. J. C 41, 199 (2005), hep-ph/0502194.
  • Helenius and Paukkunen (2018) I. Helenius and H. Paukkunen, JHEP 05, 196 (2018), 1804.03557.
  • Helenius and Paukkunen (2023) I. Helenius and H. Paukkunen, JHEP 07, 054 (2023), 2303.17864.
  • Mele and Nason (1990) B. Mele and P. Nason, Phys. Lett. B 245, 635 (1990).
  • Rijken and van Neerven (1996) P. J. Rijken and W. L. van Neerven, Phys. Lett. B 386, 422 (1996), hep-ph/9604436.
  • Blumlein and Ravindran (2006) J. Blumlein and V. Ravindran, Nucl. Phys. B 749, 1 (2006), hep-ph/0604019.
  • Melnikov and Mitov (2004) K. Melnikov and A. Mitov, Phys. Rev. D 70, 034027 (2004), hep-ph/0404143.
  • Mitov (2005) A. Mitov, Phys. Rev. D 71, 054021 (2005), hep-ph/0410205.
  • Biello and Bonino (2024) C. Biello and L. Bonino, Eur. Phys. J. C 84, 1192 (2024), 2407.07623.
  • Kartvelishvili et al. (1978) V. Kartvelishvili, A. Likhoded, and V. Petrov, Phys. Lett. B 78, 615 (1978).
  • Bowler (1981) M. G. Bowler, Z. Phys. C 11, 169 (1981).
  • Peterson et al. (1983) C. Peterson, D. Schlatter, I. Schmitt, and P. M. Zerwas, Phys. Rev. D 27, 105 (1983).
  • Andersson et al. (1983) B. Andersson, G. Gustafson, and B. Soderberg, Z. Phys. C 20, 317 (1983).
  • Collins and Spiller (1985) P. D. B. Collins and T. P. Spiller, J. Phys. G 11, 1289 (1985).
  • Colangelo and Nason (1992) G. Colangelo and P. Nason, Phys. Lett. B 285, 167 (1992).
  • Georgi (1990) H. Georgi, Phys. Lett. B 240, 447 (1990).
  • Eichten and Hill (1990) E. Eichten and B. R. Hill, Phys. Lett. B 234, 511 (1990).
  • Grinstein (1992) B. Grinstein, Ann. Rev. Nucl. Part. Sci. 42, 101 (1992).
  • Neubert (1994) M. Neubert, Phys. Rept. 245, 259 (1994), hep-ph/9306320.
  • Cacciari and Greco (1997) M. Cacciari and M. Greco, Physical Review D 55, 7134–7143 (1997), ISSN 1089-4918, URL http://dx.doi.org/10.1103/PhysRevD.55.7134.
  • Thacker and Lepage (1991) B. A. Thacker and G. P. Lepage, Phys. Rev. D 43, 196 (1991).
  • Leibovich (1997) A. K. Leibovich, Phys. Rev. D 56, 4412 (1997), hep-ph/9610381.
  • Grinstein (2000) B. Grinstein, Int. J. Mod. Phys. A 15, 461 (2000), hep-ph/9811264.
  • Krämer (2001) M. Krämer, Prog. Part. Nucl. Phys. 47, 141 (2001), hep-ph/0106120.
  • Brambilla et al. (2004) N. Brambilla et al. (Quarkonium Working Group) (2004), hep-ph/0412158.
  • Lansberg (2005) J.-P. Lansberg, Phd thesis (2005), hep-ph/0507175.
  • Lansberg (2020) J.-P. Lansberg, Phys. Rept. 889, 1 (2020), 1903.09185.
  • Alekhin et al. (2010) S. Alekhin, J. Blumlein, S. Klein, and S. Moch, Phys. Rev. D 81, 014032 (2010), 0908.2766.
  • Fleming et al. (2012) S. Fleming, A. K. Leibovich, T. Mehen, and I. Z. Rothstein, Phys. Rev. D 86, 094012 (2012), 1207.2578.
  • Kang et al. (2014) Z.-B. Kang, Y.-Q. Ma, J.-W. Qiu, and G. Sterman, Phys. Rev. D 90, 034006 (2014), 1401.0923.
  • Echevarria (2019) M. G. Echevarria, JHEP 10, 144 (2019), 1907.06494.
  • Boer et al. (2023) D. Boer, J. Bor, L. Maxia, C. Pisano, and F. Yuan, JHEP 08, 105 (2023), 2304.09473.
  • Zheng et al. (2019a) X.-C. Zheng, C.-H. Chang, T.-F. Feng, and X.-G. Wu, Phys. Rev. D 100, 034004 (2019a), 1901.03477.
  • Zheng et al. (2022) X.-C. Zheng, C.-H. Chang, and X.-G. Wu, JHEP 05, 036 (2022), 2112.10520.
  • Celiberto and Fucilla (2022a) F. G. Celiberto and M. Fucilla, Eur. Phys. J. C 82, 929 (2022a), 2202.12227.
  • Celiberto (2022b) F. G. Celiberto, Phys. Lett. B 835, 137554 (2022b), 2206.09413.
  • Celiberto (2024d) F. G. Celiberto, Eur. Phys. J. C 84, 384 (2024d), 2401.01410.
  • Aaij et al. (2015b) R. Aaij et al. (LHCb), Phys. Rev. Lett. 114, 041801 (2015b), 1411.3104.
  • Aaij et al. (2017) R. Aaij et al. (LHCb), Phys. Rev. Lett. 118, 052002 (2017), [Erratum: Phys.Rev.Lett. 119, 169901 (2017)], 1612.05140.
  • Aad et al. (2023) G. Aad et al. (ATLAS), Phys. Rev. Lett. 131, 151902 (2023), 2304.08962.
  • Hayrapetyan et al. (2024b) A. Hayrapetyan et al. (CMS), Phys. Rev. Lett. 132, 111901 (2024b), 2306.07164.
  • Zhang and Ma (2020) H.-F. Zhang and Y.-Q. Ma (2020), 2009.08376.
  • Zhu (2021) R. Zhu, Nucl. Phys. B 966, 115393 (2021), 2010.09082.
  • Feng et al. (2022a) F. Feng, Y. Huang, Y. Jia, W.-L. Sang, X. Xiong, and J.-Y. Zhang, Phys. Rev. D 106, 114029 (2022a), 2009.08450.
  • Suzuki (1977) M. Suzuki, Phys. Lett. B 71, 139 (1977).
  • Suzuki (1986) M. Suzuki, Phys. Rev. D 33, 676 (1986).
  • Amiri and Ji (1987) F. Amiri and C.-R. Ji, Phys. Lett. B 195, 593 (1987).
  • Moosavi Nejad and Amiri (2022) S. M. Moosavi Nejad and N. Amiri, Phys. Rev. D 105, 034001 (2022), 2110.15251.
  • Celiberto and Papa (2024) F. G. Celiberto and A. Papa, Phys. Lett. B 848, 138406 (2024), 2308.00809.
  • Bai et al. (2024) X.-W. Bai, F. Feng, C.-M. Gan, Y. Huang, W.-L. Sang, and H.-F. Zhang, JHEP 09, 002 (2024), 2404.13889.
  • Farashaeian and Moosavi Nejad (2024a) R. Farashaeian and S. M. Moosavi Nejad, Eur. Phys. J. A 60, 65 (2024a).
  • Farashaeian and Moosavi Nejad (2024b) R. Farashaeian and S. M. Moosavi Nejad, Eur. Phys. J. A 60, 143 (2024b).
  • Gell-Mann (1964) M. Gell-Mann, Phys. Lett. 8, 214 (1964).
  • Maiani et al. (2005) L. Maiani, F. Piccinini, A. D. Polosa, and V. Riquer, Phys. Rev. D 71, 014028 (2005), hep-ph/0412098.
  • Jaffe and Wilczek (2003) R. L. Jaffe and F. Wilczek, Phys. Rev. Lett. 91, 232003 (2003), hep-ph/0307341.
  • Guo et al. (2013) F.-K. Guo, C. Hidalgo-Duque, J. Nieves, and M. P. Valderrama, Phys. Rev. D 88, 054014 (2013), 1305.4052.
  • De Sanctis et al. (2016) M. De Sanctis, J. Ferretti, R. Magaña Vsevolodovna, P. Saracco, and E. Santopinto, Few Body Syst. 57, 1177 (2016), 1608.00387.
  • Bacchetta et al. (2008) A. Bacchetta, F. Conti, and M. Radici, Phys. Rev. D78, 074010 (2008), 0807.0323.
  • Bacchetta et al. (2010) A. Bacchetta, M. Radici, F. Conti, and M. Guagnelli, Eur. Phys. J. A45, 373 (2010), 1003.1328.
  • Bacchetta et al. (2020) A. Bacchetta, F. G. Celiberto, M. Radici, and P. Taels, Eur. Phys. J. C 80, 733 (2020), 2005.02288.
  • Bacchetta et al. (2024a) A. Bacchetta, F. G. Celiberto, and M. Radici, Eur. Phys. J. C 84, 576 (2024a), 2402.17556.
  • Chakrabarti et al. (2023) D. Chakrabarti, P. Choudhary, B. Gurjar, R. Kishore, T. Maji, C. Mondal, and A. Mukherjee, Phys. Rev. D 108, 014009 (2023), 2304.09908.
  • Banu et al. (2024) K. Banu, A. Mukherjee, A. Pawar, and S. Rajesh, Phys. Rev. D 110, 054009 (2024), 2406.00271.
  • Nzar and Hoodbhoy (1995) M. Nzar and P. Hoodbhoy, Phys. Rev. D 51, 32 (1995), hep-ph/9502349.
  • Ma et al. (2002) B.-Q. Ma, I. Schmidt, J. Soffer, and J.-J. Yang, Phys. Rev. D 65, 034004 (2002), hep-ph/0110029.
  • Falk et al. (1994) A. F. Falk, M. E. Luke, M. J. Savage, and M. B. Wise, Phys. Rev. D 49, 555 (1994), hep-ph/9305315.
  • Maiani et al. (2015) L. Maiani, A. D. Polosa, and V. Riquer, Phys. Lett. B 749, 289 (2015).
  • Faustov et al. (2020) R. N. Faustov, V. O. Galkin, and E. M. Savchenko, Phys. Rev. D 102, 114030 (2020), 2009.13237.
  • Faustov et al. (2021) R. N. Faustov, V. O. Galkin, and E. M. Savchenko, Universe 7, 94 (2021), 2103.01763.
  • Faustov et al. (2022) R. N. Faustov, V. O. Galkin, and E. M. Savchenko, Symmetry 14, 2504 (2022), 2210.16015.
  • Martynenko and Saleev (1996) A. P. Martynenko and V. A. Saleev, Phys. Lett. B 385, 297 (1996), hep-ph/9604259.
  • Gomshi Nobary and Sepahvand (2007) M. A. Gomshi Nobary and R. Sepahvand, Phys. Rev. D 76, 114006 (2007), 0711.0187.
  • Delpasand and Moosavi Nejad (2020) M. Delpasand and S. M. Moosavi Nejad, Eur. Phys. J. A 56, 56 (2020).
  • Bjorken (1978) J. D. Bjorken, Phys. Rev. D 17, 171 (1978).
  • Kinoshita (1986) K. Kinoshita, Prog. Theor. Phys. 75, 84 (1986).
  • (173) W. R. Inc., Mathematica, Version 14.2, champaign, IL, 2024, URL https://www.wolfram.com/mathematica.
  • Bertone et al. (2018) V. Bertone, N. P. Hartland, E. R. Nocera, J. Rojo, and L. Rottoli (NNPDF), Eur. Phys. J. C 78, 651 (2018), 1807.03310.
  • Mertig et al. (1991) R. Mertig, M. Bohm, and A. Denner, Comput. Phys. Commun. 64, 345 (1991).
  • Shtabovenko et al. (2016) V. Shtabovenko, R. Mertig, and F. Orellana, Comput. Phys. Commun. 207, 432 (2016), 1601.01167.
  • Shtabovenko et al. (2020) V. Shtabovenko, R. Mertig, and F. Orellana, Comput. Phys. Commun. 256, 107478 (2020), 2001.04407.
  • Gomshi Nobary (1994) M. A. Gomshi Nobary, J. Phys. G 20, 65 (1994).
  • Celiberto et al. (2016a) F. G. Celiberto, D. Yu. Ivanov, B. Murdaca, and A. Papa, Phys. Rev. D 94, 034013 (2016a), 1604.08013.
  • Celiberto et al. (2017) F. G. Celiberto, D. Yu. Ivanov, B. Murdaca, and A. Papa, Eur. Phys. J. C 77, 382 (2017), 1701.05077.
  • Bolognino et al. (2018a) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, Eur. Phys. J. C 78, 772 (2018a), 1808.05483.
  • Celiberto et al. (2021a) F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, and A. Papa, Eur. Phys. J. C 81, 780 (2021a), 2105.06432.
  • Celiberto et al. (2021b) F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, Phys. Rev. D 104, 114007 (2021b), 2109.11875.
  • Bertone et al. (2014) V. Bertone, S. Carrazza, and J. Rojo, Comput. Phys. Commun. 185, 1647 (2014), 1310.1394.
  • Carrazza et al. (2015) S. Carrazza, A. Ferrara, D. Palazzo, and J. Rojo, J. Phys. G 42, 057001 (2015), 1410.5456.
  • Bertone (2018) V. Bertone, PoS DIS2017, 201 (2018), 1708.00911.
  • Candido et al. (2022) A. Candido, F. Hekhorn, and G. Magni, Eur. Phys. J. C 82, 976 (2022), 2202.02338.
  • Hekhorn and Magni (2023) F. Hekhorn and G. Magni (2023), 2306.15294.
  • Artoisenet and Braaten (2015) P. Artoisenet and E. Braaten, JHEP 04, 121 (2015), 1412.3834.
  • Zhang et al. (2019) P. Zhang, C.-Y. Wang, X. Liu, Y.-Q. Ma, C. Meng, and K.-T. Chao, JHEP 04, 116 (2019), 1810.07656.
  • Zheng et al. (2021a) X.-C. Zheng, Z.-Y. Zhang, and X.-G. Wu, Phys. Rev. D 103, 074004 (2021a), 2101.01527.
  • Zheng et al. (2021b) X.-C. Zheng, X.-G. Wu, and X.-D. Huang, JHEP 07, 014 (2021b), 2105.14580.
  • Zheng et al. (2019b) X.-C. Zheng, C.-H. Chang, and X.-G. Wu, Phys. Rev. D 100, 014005 (2019b), 1905.09171.
  • Celiberto (2023b) F. G. Celiberto, Eur. Phys. J. C 83, 332 (2023b), 2208.14577.
  • Celiberto (2023c) F. G. Celiberto, Acta Phys. Polon. Supp. 16, 41 (2023c), 2211.11780.
  • Fadin et al. (1975) V. S. Fadin, E. Kuraev, and L. Lipatov, Phys. Lett. B 60, 50 (1975).
  • Kuraev et al. (1977) E. Kuraev, L. Lipatov, and V. S. Fadin, Sov. Phys. JETP 45, 199 (1977).
  • Balitsky and Lipatov (1978) I. Balitsky and L. Lipatov, Sov. J. Nucl. Phys. 28, 822 (1978).
  • Fadin and Lipatov (1998) V. S. Fadin and L. N. Lipatov, Phys. Lett. B 429, 127 (1998), hep-ph/9802290.
  • Ciafaloni and Camici (1998) M. Ciafaloni and G. Camici, Phys. Lett. B 430, 349 (1998), hep-ph/9803389.
  • Mueller and Navelet (1987) A. H. Mueller and H. Navelet, Nucl. Phys. B 282, 727 (1987).
  • Ducloué et al. (2013) B. Ducloué, L. Szymanowski, and S. Wallon, JHEP 05, 096 (2013), 1302.7012.
  • Colferai and Niccoli (2015) D. Colferai and A. Niccoli, JHEP 04, 071 (2015), 1501.07442.
  • Celiberto et al. (2015a) F. G. Celiberto, D. Yu. Ivanov, B. Murdaca, and A. Papa, Eur. Phys. J. C 75, 292 (2015a), 1504.08233.
  • Celiberto et al. (2015b) F. G. Celiberto, D. Yu. Ivanov, B. Murdaca, and A. Papa, Acta Phys. Polon. Supp. 8, 935 (2015b), 1510.01626.
  • Celiberto et al. (2016b) F. G. Celiberto, D. Yu. Ivanov, B. Murdaca, and A. Papa, Eur. Phys. J. C 76, 224 (2016b), 1601.07847.
  • Celiberto (2017) F. G. Celiberto, Phd thesis, Università della Calabria and INFN-Cosenza (2017), 1707.04315.
  • Caporale et al. (2018) F. Caporale, F. G. Celiberto, G. Chachamis, D. Gordo Gómez, and A. Sabio Vera, Nucl. Phys. B 935, 412 (2018), 1806.06309.
  • de León et al. (2021) N. B. de León, G. Chachamis, and A. Sabio Vera, Eur. Phys. J. C 81, 1019 (2021), 2106.11255.
  • Celiberto and Papa (2022) F. G. Celiberto and A. Papa, Phys. Rev. D 106, 114004 (2022), 2207.05015.
  • Baldenegro et al. (2024) C. Baldenegro, G. Chachamis, M. Kampshoff, M. Klasen, G. J. Milhano, C. Royon, and A. Sabio Vera, Phys. Rev. D 110, 114027 (2024), 2406.10681.
  • Celiberto et al. (2020) F. G. Celiberto, D. Yu. Ivanov, and A. Papa, Phys. Rev. D 102, 094019 (2020), 2008.10513.
  • Bolognino et al. (2019a) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, PoS DIS2019, 049 (2019a), 1906.11800.
  • Bolognino et al. (2019b) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, Acta Phys. Polon. Supp. 12, 773 (2019b), 1902.04511.
  • Mohammed (2022) M. M. A. Mohammed, Phd thesis, Università della Calabria and INFN-Cosenza (2022), 2204.11606.
  • Caporale et al. (2016) F. Caporale, F. G. Celiberto, G. Chachamis, D. Gordo Gómez, and A. Sabio Vera, Nucl. Phys. B 910, 374 (2016), 1603.07785.
  • Caporale et al. (2017a) F. Caporale, F. G. Celiberto, G. Chachamis, D. Gordo Gómez, and A. Sabio Vera, Eur. Phys. J. C 77, 5 (2017a), 1606.00574.
  • Celiberto (2016) F. G. Celiberto, Frascati Phys. Ser. 63, 43 (2016), 1606.07327.
  • Caporale et al. (2017b) F. Caporale, F. G. Celiberto, G. Chachamis, D. Gordo Gómez, and A. Sabio Vera, Phys. Rev. D 95, 074007 (2017b), 1612.05428.
  • Hentschinski et al. (2021) M. Hentschinski, K. Kutak, and A. van Hameren, Eur. Phys. J. C 81, 112 (2021), [Erratum: Eur. Phys. J. C 81, 262 (2021)], 2011.03193.
  • Celiberto et al. (2022a) F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, JHEP 08, 092 (2022a), 2205.02681.
  • Celiberto et al. (2021c) F. G. Celiberto, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, Eur. Phys. J. C 81, 293 (2021c), 2008.00501.
  • Celiberto and Papa (2023) F. G. Celiberto and A. Papa (2023), 2305.00962.
  • Celiberto et al. (2023) F. G. Celiberto, L. Delle Rose, M. Fucilla, G. Gatto, and A. Papa, in 57th Rencontres de Moriond on QCD and High Energy Interactions (2023), 2305.05052.
  • Celiberto et al. (2024b) F. G. Celiberto, L. Delle Rose, M. Fucilla, G. Gatto, and A. Papa, PoS RADCOR2023, 069 (2024b), 2309.11573.
  • Celiberto et al. (2024c) F. G. Celiberto, L. Delle Rose, M. Fucilla, G. Gatto, and A. Papa, PoS EPS-HEP2023, 390 (2024c), 2310.16967.
  • Celiberto et al. (2024d) F. G. Celiberto, M. Fucilla, M. M. A. Mohammed, D. Yu. Ivanov, and A. Papa, PoS RADCOR2023, 091 (2024d), 2309.07570.
  • Celiberto et al. (2022b) F. G. Celiberto, M. Fucilla, M. M. A. Mohammed, and A. Papa, Phys. Rev. D 105, 114056 (2022b), 2205.13429.
  • Celiberto et al. (2024e) F. G. Celiberto, L. Delle Rose, M. Fucilla, G. Gatto, and A. Papa, JHEP 12, 061 (2024e), 2409.20354.
  • Celiberto et al. (2018a) F. G. Celiberto, D. Gordo Gómez, and A. Sabio Vera, Phys. Lett. B786, 201 (2018a), 1808.09511.
  • Golec-Biernat et al. (2018) K. Golec-Biernat, L. Motyka, and T. Stebel, JHEP 12, 091 (2018), 1811.04361.
  • Celiberto et al. (2018b) F. G. Celiberto, D. Yu. Ivanov, B. Murdaca, and A. Papa, Phys. Lett. B 777, 141 (2018b), 1709.10032.
  • Boussarie et al. (2018) R. Boussarie, B. Ducloué, L. Szymanowski, and S. Wallon, Phys. Rev. D 97, 014008 (2018), 1709.01380.
  • Bolognino et al. (2019c) A. D. Bolognino, F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, B. Murdaca, and A. Papa, PoS DIS2019, 067 (2019c), 1906.05940.
  • Bolognino et al. (2019d) A. D. Bolognino, F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, and A. Papa, Eur. Phys. J. C 79, 939 (2019d), 1909.03068.
  • Bolognino et al. (2023) A. D. Bolognino, F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, M. M. A. Mohammed, and A. Papa, Acta Phys. Polon. Supp. 16, 17 (2023), 2211.16818.
  • Celiberto and Fucilla (2022b) F. G. Celiberto and M. Fucilla, in 29th International Workshop on Deep-Inelastic Scattering and Related Subjects (2022b), 2208.07206.
  • Anikin et al. (2011) I. Anikin, A. Besse, D. Yu. Ivanov, B. Pire, L. Szymanowski, and S. Wallon, Phys. Rev. D 84, 054004 (2011), 1105.1761.
  • Besse et al. (2013) A. Besse, L. Szymanowski, and S. Wallon, JHEP 11, 062 (2013), 1302.1766.
  • Bolognino et al. (2018b) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, and A. Papa, Eur. Phys. J. C78, 1023 (2018b), 1808.02395.
  • Bolognino et al. (2018c) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, and A. Papa, Frascati Phys. Ser. 67, 76 (2018c), 1808.02958.
  • Bolognino et al. (2019e) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, and A. Papa, Acta Phys. Polon. Supp. 12, 891 (2019e), 1902.04520.
  • Bolognino et al. (2020) A. D. Bolognino, A. Szczurek, and W. Schaefer, Phys. Rev. D 101, 054041 (2020), 1912.06507.
  • Celiberto (2019) F. G. Celiberto, Nuovo Cim. C42, 220 (2019), 1912.11313.
  • Bolognino (2021) A. D. Bolognino, Phd thesis, Calabria U. (2021), 2109.03033.
  • Łuszczak et al. (2022) A. Łuszczak, M. Łuszczak, and W. Schäfer, Phys. Lett. B 835, 137582 (2022), 2210.02877.
  • Bolognino et al. (2021a) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, A. Papa, W. Schäfer, and A. Szczurek, Eur. Phys. J. C 81, 846 (2021a), 2107.13415.
  • Bolognino et al. (2022a) A. D. Bolognino, F. G. Celiberto, D. Yu. Ivanov, and A. Papa, SciPost Phys. Proc. 8, 089 (2022a), 2107.12725.
  • Bolognino et al. (2022b) A. D. Bolognino, F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, A. Papa, W. Schäfer, and A. Szczurek, Rev. Mex. Fis. Suppl. 3, 0308109 (2022b), 2202.02513.
  • Bolognino et al. (2022c) A. D. Bolognino, F. G. Celiberto, D. Yu.. Ivanov, A. Papa, W. Schäfer, and A. Szczurek, in 29th International Workshop on Deep-Inelastic Scattering and Related Subjects (2022c), 2207.05726.
  • Ball et al. (2018) R. D. Ball, V. Bertone, M. Bonvini, S. Marzani, J. Rojo, and L. Rottoli, Eur. Phys. J. C78, 321 (2018), 1710.05935.
  • Abdolmaleki et al. (2018) H. Abdolmaleki et al. (xFitter Developers’ Team), Eur. Phys. J. C 78, 621 (2018), 1802.00064.
  • Bonvini and Giuli (2019) M. Bonvini and F. Giuli, Eur. Phys. J. Plus 134, 531 (2019), 1902.11125.
  • Silvetti and Bonvini (2023) F. Silvetti and M. Bonvini, Eur. Phys. J. C 83, 267 (2023), 2211.10142.
  • Silvetti (2023) F. Silvetti, Phd thesis, Rome U. (2023), 2403.20315.
  • Rinaudo (2024) A. Rinaudo, Phd thesis, Genoa U. (2024).
  • Celiberto (2021b) F. G. Celiberto, Nuovo Cim. C44, 36 (2021b), 2101.04630.
  • Bacchetta et al. (2022a) A. Bacchetta, F. G. Celiberto, M. Radici, and P. Taels, SciPost Phys. Proc. 8, 040 (2022a), 2107.13446.
  • Bacchetta et al. (2022b) A. Bacchetta, F. G. Celiberto, and M. Radici, PoS EPS-HEP2021, 376 (2022b), 2111.01686.
  • Bacchetta et al. (2022c) A. Bacchetta, F. G. Celiberto, and M. Radici, PoS PANIC2021, 378 (2022c), 2111.03567.
  • Bacchetta et al. (2022d) A. Bacchetta, F. G. Celiberto, and M. Radici, JPS Conf. Proc. 37, 020124 (2022d), 2201.10508.
  • Bacchetta et al. (2022e) A. Bacchetta, F. G. Celiberto, and M. Radici, Rev. Mex. Fis. Suppl. 3, 0308108 (2022e), 2206.07815.
  • Bacchetta et al. (2022f) A. Bacchetta, F. G. Celiberto, M. Radici, and A. Signori, in 29th International Workshop on Deep-Inelastic Scattering and Related Subjects (2022f), 2208.06252.
  • Celiberto (2022c) F. G. Celiberto, Universe 8, 661 (2022c), 2210.08322.
  • Bacchetta et al. (2024b) A. Bacchetta, F. G. Celiberto, and M. Radici, PoS EPS-HEP2023, 247 (2024b), 2310.19916.
  • Ducloué et al. (2014) B. Ducloué, L. Szymanowski, and S. Wallon, Phys. Rev. Lett. 112, 082003 (2014), 1309.3229.
  • Caporale et al. (2014) F. Caporale, D. Yu. Ivanov, B. Murdaca, and A. Papa, Eur. Phys. J. C 74, 3084 (2014), [Erratum: Eur.Phys.J.C 75, 535 (2015)], 1407.8431.
  • Chang and Chen (1992) C.-H. Chang and Y.-Q. Chen, Phys. Rev. D 46, 3845 (1992), [Erratum: Phys.Rev.D 50, 6013 (1994)].
  • Ma (1994) J. P. Ma, Phys. Lett. B 332, 398 (1994), hep-ph/9401249.
  • Feng et al. (2022b) F. Feng, Y. Jia, and D. Yang, Phys. Rev. D 106, 054030 (2022b), 2112.15569.
  • Feng and Jia (2023) F. Feng and Y. Jia, Chin. Phys. C 47, 033103 (2023), 1810.04138.
  • Binosi et al. (2009) D. Binosi, J. Collins, C. Kaufhold, and L. Theussl, Comput. Phys. Commun. 180, 1709 (2009), 0811.4113.
  • Caporale et al. (2013) F. Caporale, D. Yu. Ivanov, B. Murdaca, and A. Papa, Nucl. Phys. B 877, 73 (2013), 1211.7225.
  • Kotikov and Lipatov (2000) A. V. Kotikov and L. N. Lipatov, Nucl. Phys. B 582, 19 (2000), hep-ph/0004008.
  • Ivanov and Papa (2012a) D. Yu. Ivanov and A. Papa, JHEP 07, 045 (2012a), 1205.6068.
  • Ivanov and Papa (2012b) D. Yu. Ivanov and A. Papa, JHEP 05, 086 (2012b), 1202.1082.
  • Furman (1982) M. Furman, Nucl. Phys. B 197, 413 (1982).
  • Aversa et al. (1989) F. Aversa, P. Chiappetta, M. Greco, and J. P. Guillet, Nucl. Phys. B 327, 105 (1989).
  • Alioli et al. (2010) S. Alioli, P. Nason, C. Oleari, and E. Re, JHEP 06, 043 (2010), 1002.2581.
  • Campbell et al. (2012) J. M. Campbell, R. K. Ellis, R. Frederix, P. Nason, C. Oleari, and C. Williams, JHEP 07, 092 (2012), 1202.5475.
  • Hamilton et al. (2013) K. Hamilton, P. Nason, C. Oleari, and G. Zanderighi, JHEP 05, 082 (2013), 1212.4504.
  • Ball et al. (2021) R. D. Ball et al. (NNPDF), Eur. Phys. J. C 81, 958 (2021), 2109.02671.
  • Ball et al. (2022a) R. D. Ball et al. (NNPDF), Eur. Phys. J. C 82, 428 (2022a), 2109.02653.
  • Khachatryan et al. (2016) V. Khachatryan et al. (CMS), JHEP 08, 139 (2016), 1601.06713.
  • Khachatryan et al. (2021) V. Khachatryan et al. (CMS), JINST 16, P02010 (2021), 2011.01185.
  • Chatrchyan et al. (2012) S. Chatrchyan et al. (CMS), Phys. Lett. B 714, 136 (2012), 1205.0594.
  • Bolognino et al. (2021b) A. D. Bolognino, F. G. Celiberto, M. Fucilla, D. Yu. Ivanov, and A. Papa, Phys. Rev. D 103, 094004 (2021b), 2103.07396.
  • Bartels et al. (2002) J. Bartels, D. Colferai, and G. P. Vacca, Eur. Phys. J. C 24, 83 (2002), hep-ph/0112283.
  • Kassabov et al. (2023) Z. Kassabov, M. Ubiali, and C. Voisey, JHEP 03, 148 (2023), 2207.07616.
  • Harland-Lang and Thorne (2019) L. A. Harland-Lang and R. S. Thorne, Eur. Phys. J. C 79, 225 (2019), 1811.08434.
  • Ball and Pearson (2021) R. D. Ball and R. L. Pearson, Eur. Phys. J. C 81, 830 (2021), 2105.05114.
  • McGowan et al. (2023) J. McGowan, T. Cridge, L. A. Harland-Lang, and R. S. Thorne, Eur. Phys. J. C 83, 185 (2023), [Erratum: Eur.Phys.J.C 83, 302 (2023)], 2207.04739.
  • Ball et al. (2024a) R. D. Ball et al. (NNPDF), Eur. Phys. J. C 84, 517 (2024a), 2401.10319.
  • Pasquini et al. (2023) B. Pasquini, S. Rodini, and S. Venturini (MAP (Multi-dimensional Analyses of Partonic distributions)), Phys. Rev. D 107, 114023 (2023), 2303.01789.
  • Brodsky et al. (1980) S. J. Brodsky, P. Hoyer, C. Peterson, and N. Sakai, Phys. Lett. B 93, 451 (1980).
  • Brodsky et al. (2015) S. J. Brodsky, A. Kusina, F. Lyonnet, I. Schienbein, H. Spiesberger, and R. Vogt, Adv. High Energy Phys. 2015, 231547 (2015), 1504.06287.
  • Ball et al. (2022b) R. D. Ball, A. Candido, J. Cruz-Martinez, S. Forte, T. Giani, F. Hekhorn, K. Kudashkin, G. Magni, and J. Rojo (NNPDF), Nature 608, 483 (2022b), 2208.08372.
  • Guzzi et al. (2023) M. Guzzi, T. J. Hobbs, K. Xie, J. Huston, P. Nadolsky, and C. P. Yuan, Phys. Lett. B 843, 137975 (2023), 2211.01387.
  • Ball et al. (2024b) R. D. Ball, A. Candido, J. Cruz-Martinez, S. Forte, T. Giani, F. Hekhorn, G. Magni, E. R. Nocera, J. Rojo, and R. Stegeman (NNPDF), Phys. Rev. D 109, L091501 (2024b), 2311.00743.
  • Flore et al. (2020) C. Flore, J.-P. Lansberg, H.-S. Shao, and Y. Yedelkina, Phys. Lett. B 811, 135926 (2020), 2009.08264.
  • Chapon et al. (2022) E. Chapon et al., Prog. Part. Nucl. Phys. 122, 103906 (2022), 2012.14161.
  • Accardi et al. (2025) A. Accardi et al. (LHCspin) (2025), 2504.16034.
  • Abdul Khalek et al. (2022a) R. Abdul Khalek et al., Nucl. Phys. A 1026, 122447 (2022a), 2103.05419.
  • Abdul Khalek et al. (2022b) R. Abdul Khalek et al., in 2022 Snowmass Summer Study (2022b), 2203.13199.
  • Hentschinski et al. (2023) M. Hentschinski et al., Acta Phys. Polon. B 54, 2 (2023), 2203.08129.
  • Amoroso et al. (2022) S. Amoroso et al., Acta Phys. Polon. B 53, A1 (2022), 2203.13923.
  • Abir et al. (2023) R. Abir et al. (2023), 2305.14572.
  • Allaire et al. (2024) C. Allaire et al., Comput. Softw. Big Sci. 8, 5 (2024), 2307.08593.
  • Anchordoqui et al. (2022) L. A. Anchordoqui et al., Phys. Rept. 968, 1 (2022), 2109.10905.
  • Feng et al. (2023) J. L. Feng et al., J. Phys. G 50, 030501 (2023), 2203.05090.
  • Adachi et al. (2022) I. Adachi et al. (ILC International Development Team and ILC Community) (2022), 2203.07622.
  • Balazs et al. (2025) C. Balazs et al. (Linear Collider) (2025), 2503.24049.
  • Attié et al. (2025) D. Attié et al. (Linear Collider Vision) (2025), 2503.19983.
  • Arbuzov et al. (2021) A. Arbuzov et al., Prog. Part. Nucl. Phys. 119, 103858 (2021), 2011.15005.
  • Accettura et al. (2023) C. Accettura et al., Eur. Phys. J. C 83, 864 (2023), [Erratum: Eur.Phys.J.C 84, 36 (2024)], 2303.08533.
  • Accettura et al. (2024a) C. Accettura et al. (International Muon Collider), 2/2024 (2024a), 2407.12450.
  • Accettura et al. (2024b) C. Accettura et al. (MuCoL) (2024b), 2411.02966.
  • Black et al. (2024) K. M. Black et al., JINST 19, T02015 (2024), 2209.01318.
  • Accettura et al. (2025) C. Accettura et al. (International Muon Collider) (2025), 2504.21417.
  • Accardi et al. (2024) A. Accardi et al., Eur. Phys. J. A 60, 173 (2024), 2306.09360.