Dynamical mechanisms for deuteron production at mid-rapidity in relativistic heavy-ion collisions from SIS to RHIC energies
Abstract
The understanding of the mechanisms for the production of weakly bound clusters, such as a deuteron , in heavy-ion reactions at mid-rapidity is presently one of the challenging problems which is also known as the “ice in a fire” puzzle. In this study we investigate the dynamical formation of deuterons within the Parton-Hadron-Quantum-Molecular Dynamics (PHQMD) microscopic transport approach and advance two microscopic production mechanisms to describe deuterons in heavy-ion collisions from SIS to RHIC energies: kinetic production by hadronic reactions and potential production by the attractive potential between nucleons. Differently to other studies, for the “kinetic” deuterons we employ the full isospin decomposition of the various , channels and take into account the finite-size properties of the deuteron by means of an excluded volume condition in coordinate space and by the projection onto the deuteron wave function in momentum space. We find that considering the quantum nature of the deuteron in coordinate and momentum space reduces substantially the kinetic deuteron production in a dense medium as encountered in heavy-ion collisions. If we add the “potential” deuterons by applying an advanced Minimum Spanning Tree (aMST) procedure, we obtain good agreement with the available experimental data from SIS energies up to the top RHIC energy.
pacs:
12.38MhI Introduction
Quantum Chromodynamics (QCD), the theory describing the strong interaction between quarks and gluons, the elementary components of hadrons, owns important features, which have not yet been understood. To study this QCD matter under extreme conditions of temperature and density is the primary purpose of Heavy-Ion Collisions (HICs) at ultra-relativistic energies Busza et al. (2018); Jacak (2021), which are performed at the BNL Relativistic Heavy-ion Collider (RHIC) and at the CERN Large Hadron Collider (LHC). At very high energies the energy deposited during the initial stage of the collisions creates an almost net-baryon free hot medium consisting of deconfined quarks and gluons, which is called Quark-Gluon Plasma (QGP). On the theoretical side the knowledge of the QCD phase diagram, describing the pressure as a function of temperature and baryon chemical potential , is limited to the region of high and almost zero . There QCD calculations on lattices (lQCD) Borsanyi et al. (2014); Bazavov et al. (2014) predict a smooth crossover between the QGP phase and a gas of hadrons.
The ongoing Beam Energy Scan (BES) program at RHIC, as well as the future experiments at the Nuclotron based Ion Collider (NICA) and at the Facility for Antiproton and Ion Research (FAIR), under construction in Dubna and Darmstadt, respectively, will extend the study of strongly interacting matter to lower collision energies. The aim is to explore the QCD phase diagram at high net-baryon density and to search for the existence of a Critical End Point (CEP) at the end of a first-order phase boundary at non-zero , predicted by effective theories Asakawa and Yazaki (1989); Stephanov (2004).
To explore this transition from QCD matter to hadrons the study of the production of light nuclei, such as , , , and hypernuclei, is an important issue because the production of composite clusters depends on the correlations and fluctuations of the nucleons. The interest in light nuclei comes from both, experiments and theory.
From the experimental side, the observation of light nuclei began with the first heavy-ion experiments at the Bevalac accelerator Westfall et al. (1976); Gutbrod et al. (1976); Nagamiya et al. (1981) (after some low statistics bubble chamber data Grunder et al. (1971)). It continued at AGS Ahle et al. (2000a, b); Bennett et al. (1998); Saito et al. (1994); Armstrong et al. (2000),
the GSI SIS facility Reisdorf et al. (2010) and the SPS collider Anticic et al. (2004, 2016). Nowadays, the measurements of light nuclei and hypernuclei at mid-rapidity represent an important research program for the STAR collaboration Adam et al. (2019) at RHIC and for the ALICE collaboration Adam et al. (2016); Acharya et al. (2017) at LHC.
At low (SIS) beam energies between 30% and 50% of protons are bound in deuterons, tritons and , while this fraction decreases with increasing beam energy to values around 1% at the LHC (see, for instance, Ref. Dönigus et al. (2022)).
Collective variables, like the directed or the elliptic flow, are different for clusters and nucleons, indicating that clusters test different phase-space regions than nucleons.
Therefore, bound nucleons represent an interesting probe to study the dynamics of heavy-ion collisions.
From the theoretical side, the reason is even more fundamental since the mechanism of cluster formation in nucleus-nucleus collisions is not well understood. The deuterons with a binding energy of MeV appear to be fragile objects compared to the average kinetic energy of hadrons which surround them. At freeze-out, when the QGP is converted into hadrons, the kinetic freeze-out parameters indicate a temperature of MeV. It is surprising that they can survive in such an environment without being destroyed by collisions with the surrounding hadrons. Hence, it is puzzling that light nuclei are observed in central HICs at mid-rapidity at all, and it is even more puzzling that their multiplicity is well described by statistical model calculations Andronic et al. (2011). This observation has been portrayed as “ice cubes in a fire” Bratkovskaya et al. (2022), or “snowballs in hell” Oliinychenko et al. (2019). The presence of light clusters one may consider as a hint that they do not come from the same phase-space regions as nucleons, which makes them interesting for the study of the reaction dynamics. The formation of light nuclei at mid-rapidity at beam energies above 2 AGeV has been modeled by three main approaches:
-
i)
In the statistical model hadrons at mid-rapidity are assumed to be emitted from a common thermal source, which is characterized by the temperature , the chemical potential and a fixed volume Cleymans and Satz (1993); Andronic et al. (2013). All three quantities are determined by fitting the multiplicity of a multitude of hadrons. Surprisingly, the observed cluster multiplicities are also described with the same fit variables , and Andronic et al. (2011, 2018). The assumption of the statistical model approach is that the hadronic expansion of the system does not change the number of clusters.
-
ii)
In the coalescence approach it is assumed that a proton and a neutron form a deuteron if their distance in momentum and coordinate space is smaller than the coalescence parameters Sombun et al. (2019); Butler and Pearson (1963). This distance is measured when the last nucleon of the pair undergoes its last elastic or inelastic collision. Several variations of the coalescence model are being used. Some of them project the phase-space distribution function of the nucleons to the Wigner density of the relative coordinates of the nucleons in the deuteron. This distribution is usually approximated by a Gaussian form Scheibl and Heinz (1999); Zhu et al. (2015). However, the coalescence approach neglects that a deuteron is a bound object, which cannot be formed by a simple “fusion” of two nucleons, since it would violate the energy-momentum conservation. The formation of a deuteron is only possible if it interacts with the environment by a potential or via scattering processes. Nevertheless, this approach reproduces well the -spectrum of deuterons, as well as their multiplicity for a large range of beam energies. For most recent studies on deuteron production with the coalescence approach we refer to Refs. Sun et al. (2018); Zhao et al. (2020); Kittiratpattana et al. (2022).
-
iii)
The Minimum Spanning Tree (MST) approach has been originally advanced in Ref. Aichelin (1991) to study fragments which come from the projectile and target region and later it has also been employed to study mid-rapidity clusters Aichelin et al. (2020). It assumes that at the end of the heavy-ion reaction two nucleons are part of a cluster if their distance is smaller than a radius which is of the order of the range of the nucleon-nucleon interaction. As investigated in a successive study Gläßel et al. (2022), this model reproduces well the and spectra not only for deuterons, but also for all clusters, observed at mid-rapidity, in the energy range from AGS to top RHIC.
Recently, a fourth approach has been advanced. In Refs. Oliinychenko et al. (2019, 2021); Staudenmaier et al. (2021) it has been claimed that deuterons can also be created by elementary collisions: , , . Based on Danielewicz and Bertsch (1991); Kapusta (1980); Siemens and Kapusta (1979), where the production (disintegration) of deuterons by (nucleon catalysis) was studied at low energy HICs, it has been proposed in Ref. Oliinychenko et al. (2019) that at relativistic HICs the pion catalyis, i.e. , becomes more dominant at mid-rapidity due to the large abundance of pions. To demonstrate this, inelastic scatterings and the inverse processes have been implemented in the transport approach SMASH Weil et al. (2016), which describes the hadronic stage of HICs. In the study Oliinychenko et al. (2019) the catalysis reactions have been approximated as simple two-step processes of and where is a fictitious dibaryon resonance with mass and width determined by fitting the experimental total inclusive cross section for inelastic scattering. With this approach the deuteron multiplicity and -spectra at mid-rapidity could be reproduced for LHC Pb+Pb collisions TeV and for Au+Au collisions in the energy range of the RHIC BES ( GeV) Oliinychenko et al. (2021). Later, in Ref. Staudenmaier et al. (2021), the numerical artifact of employing the intermediate state has been replaced by the treatment of multi-particle reactions within the covariant rate formalism, firstly developed in Ref. Cassing (2002). In both studies the deuteron was treated as a point-like particle.
In this work we revise and improve two of the above mentioned dynamical processes for deuteron production in HICs, the “kinetic” production by hadronic collisions and the “potential” mechanism, where bound nucleons form deuterons and heavier clusters by potential interactions, and combine them to obtain a comprehensive approach for the description of the experimental measurements at mid-rapidity. For this study we use the Parton-Hadron-Quantum-Molecular Dynamics (PHQMD) transport approach Aichelin et al. (2020).
Concerning the first approach, we include, in contradistinction to Oliinychenko et al. (2019, 2021); Staudenmaier et al. (2021), all possible isospin channels for reactions which enhances the production rate compared to the case. Following Ref. Sun et al. (2021), in this “kinetic” mechanism we also take into account the distribution of the relative momentum of the two nucleons inside a deuteron. Regarding the second approach, we overcome the problem discussed in Ref. Aichelin et al. (2020) that a choice had to be made at which time the cluster analysis with MST takes place. We will show that an asymptotic distribution of stable clusters, which are also “bound” in the sense that they have negative binding energies , can be obtained, independent of the time when the clusters are identified. In order to do so, we present a novel cluster recognition procedure based on the MST algorithm used in point iii), which is further developed in order to trace the entire dynamical evolution of the baryons which are bound into a stable cluster. It is the purpose of this paper to show that the combination of such an advanced MST (aMST) approach and the production of deuterons by collisions gives a very good description of the total multiplicity, and the spectra of deuterons from SIS ( GeV) up to the highest RHIC energy ( GeV).
This paper is organized as follows: After the introduction given in Sec. I, in Sec. II.A we recall the basic ideas of the PHQMD transport approach. The identification of deuterons bound by potential interaction by means of the MST clusterization algorithm is the subject of Sec. II.B. In particular, after discussing in Sec. II.B.1 the basis of the original MST model employed in previous PHQMD studies (see Ref. Aichelin et al. (2020)), we present in Sec. II.B.2 our new “advanced” MST (aMST) approach. The theoretical formulation of the main hadronic reactions for the production of “kinetic” deuterons is the topic of Sec. III. In Sec. IV we test such deuteron reactions in a “box” and verify their correct numerical implementation by comparing with analytic rate results. In Sec. V we investigate the main physical effects of production and disintegration of deuterons by hadronic reactions in heavy-ion simulations within the PHQMD approach. The details on how the two “kinetic” and “potential” mechanisms are combined within the PHQMD framework are reported at the end of this section. In Sec. VI we confront our final results with combined kinetic and potential deuterons with the existing experimental data for rapidity and transverse momentum distributions in HICs from invariant center-of-mass collision energies of GeV to GeV. Finally, we outline our conclusions in Section VII.
II Model description
II.1 PHQMD
The Parton-Hadron-Quantum Molecular Dynamics (PHQMD) has been recently conceived as a new type of microscopic transport approach which combines the characteristics of baryon propagation from the Quantum Molecular Dynamics (QMD) model Aichelin (1991); Aichelin et al. (1987, 1988); Hartnack et al. (1998) and the dynamical properties and interactions in and out of equilibrium of hadronic and partonic degrees of freedom of the Parton-Hadron-String-Dynamics (PHSD) approach Cassing and Bratkovskaya (2008); Cassing (2009); Cassing and Bratkovskaya (2009); Bratkovskaya et al. (2011); Linnyk et al. (2016).
In this section we provide a short summary of these two building blocks. For more details of the PHQMD model we refer to Ref. Aichelin et al. (2020).
I. In QMD the baryons are described by single-particle wave functions of Gaussian form with a time independent width. The Wigner density of each particle is obtained by a Fourier transformation of the density matrix. Then, the n-body Wigner density is constructed by the direct product of the single-particle Wigner densities and its propagation is determined by a generalized Ritz variational principle Feldmeier (1990). Contrary to mean-field approaches, where the n-body phase-space correlations are integrated out and the dynamics is reduced to a single-particle propagation in a mean-field potential, in QMD these correlations are preserved and the fluctuations not suppressed. This allows to investigate the dynamical formation of clusters, which are correlations between nucleons.
In PHQMD a baryon is represented by the single-particle Wigner density, which is given by
(1) |
the Gaussian width is taken as fm2.
The QMD equations of motion (EoMs) for the centroids are similar to those of the Hamilton equations for a classical particle Aichelin (1991)
(2) |
where the difference lies in the fact that on the right hand side of both equations the expectation value of the quantum Hamiltonian of the many-body system appears. We note in passing that for a non-Gaussian form of the wave function the time evolution equations are different. The Hamiltonian is the sum of the kinetic energies of the particles and of their (density dependent) two-body interaction. The expectation value can be written as
(3) |
The potential interaction consists of two parts: the Coulomb interaction and a local density dependent Skyrme potential . The expectation value of the Coulomb interaction can be calculated analytically. The expectation value of the Skyrme part contains terms and , where is the local density. Their weights, as well as the exponent , are tuned to the Equation of State (EoS) of infinite nuclear matter MeV, where fm-3 is the saturation density at zero temperature. This fixes two of the three parameters. In PHQMD two parameter sets have been introduced, which yield a “soft” and a “hard” EoS, respectively. For details on the realization of the QMD dynamics and the impact of different EoS on bulk and cluster observables we refer to Ref. Aichelin et al. (2020). For bulk and strangeness particle production in PHQMD with a “hard” and a “soft” EoS at low energy HICs and the comparison with other transport models see also Ref. Reichert et al. (2022). In our present study of the deuteron production mechanisms, we employ the PHQMD in its “hard” EoS setup like in Ref. Gläßel et al. (2022).
Finally, we want to stress two aspects: i) in this study, as in the previous PHQMD publications, we employ a “static”, i.e. momentum independent, Skyrme potential. A form of the Skyrme interaction which contains also a momentum dependent part will be reserved for future work.
ii) The PHQMD approach aims to provide a dynamical description of cluster formation in HICs at low-energy, as well as at relativistic energies. The discussed QMD model uses non-relativistic quantum wave functions.
The relativistic formulation of QMD dynamics as a n-body theory has been developed in Refs. Sorge et al. (1989); Marty and Aichelin (2013) by introducing extra constraints in order to reduce the 8N-dimensional phase-space to the (6N + 1)-dimensional phase-space in which particle trajectories can be defined.
However, this method is computationally extremely expensive, requiring the inversion of a matrix of size in each time step and thus, it is presently not applicable for high statistics simulations.
Therefore, in PHQMD the original framework of QMD is extended to relativistic energies by introducing a modified single-particle Wigner density for each nucleon :
(4) | |||
which accounts for the Lorentz contraction of the nucleus in the beam -direction in coordinate and momentum space by including , where is the velocity of projectile and target in the computational frame, which is the center-of-mass system of the heavy-ion collision. Accordingly, the interaction density modifies as
(5) | |||||
We refer again to Ref. Aichelin et al. (2020) and Ref. Aichelin (1991) for a more detailed discussion and the explicit formulas.
We note that in PHQMD the nuclei are initialized in their rest frame with the Gaussian distributions Eq. (1). The Lorentz squeezing of nuclei by the gamma factor is done after the initialization of the nuclei in their rest frame, so it does not affect the initialization.
II. As in the PHSD (Parton-Hadron-String Dynamics) approach Cassing and Bratkovskaya (2008); Cassing (2009); Cassing and Bratkovskaya (2009); Bratkovskaya et al. (2011); Linnyk et al. (2016), in PHQMD the strongly interacting medium is described by off-shell hadrons and off-shell massive quasi-particles representing the deconfined quarks and gluons of the QGP phase, which is created if the local energy density is larger than a critical value of GeV/fm3. The propagation of these off-shell degrees of freedom, including their spectral functions, is based on the Kadanoff-Baym transport theory Kadanoff and Baym (1962) in first-order gradient expansion from which the Cassing-Juchem generalized off-shell transport equations Cassing and Juchem (2000a, b); Cassing (2009) in test-particle representation are derived (see Ref. Cassing (2021)). The hadronic part is taken from the early development of the Hadron-String-Dynamics (HSD) approach (see Ref. Cassing and Bratkovskaya (1999) for a detailed description of the baryon, meson and resonance species implemented).
The elementary baryon-baryon (), meson-baryon () and meson-meson () reactions for multi-particle production are realized according to the Lund string model Nilsson-Almqvist and Stenlund (1987) via the event generators FRITIOF 7.02 Nilsson-Almqvist and Stenlund (1987); Andersson et al. (1993) and PYTHIA 7.4 Sjostrand et al. (2006). Both generators are “tuned” for a better description of the experimental data for elementary collisions at intermediate energies Kireyeu et al. (2020).
The partonic part, which describes the QGP phase, follows the description of the so-called Dynamical Quasi-Particle Model (DQPM) Cassing (2007a, b); Peshier and Cassing (2005). In the DQPM quarks and gluons are represented by massive, strongly interacting quasi-particles. They have spectral functions whose pole positions and widths are defined by the real and imaginary terms of parton self-energies. The parton masses and widths are functions of the temperature (and in the most recent extension Soloveva et al. (2020) also of the baryon-chemical potential ) through an effective coupling constant, which is fixed by fitting lQCD results from Refs. Aoki et al. (2009); Borsanyi et al. (2014); Bazavov et al. (2014); Cheng et al. (2008); Borsanyi et al. (2015). These DQPM partons are evolved with their self-energies according to the same off-shell transport equations and scatter by microscopically computed cross sections.
We recall that in PHQMD only the propagation of mesons and partons relies on the PHSD approach, while the baryons evolve according to the QMD dynamics. However, it is always possible to run PHQMD in the “(P)HSD-mode” by switching the baryon propagation back to the mean-field dynamics of HSD. Again we refer to Ref. Aichelin et al. (2020) for a description and detailed studies.
As already stated, in PHQMD the collision integral, which encodes the main scattering/dissipative processes of hadrons and partons, is adopted from the PHSD model. The main hadronic reactions have been implemented for many observables, like strangeness, dileptons, photons, heavy quarks, etc. (cf. examples in the reviews Linnyk et al. (2016); Bleicher and Bratkovskaya (2022)). It contains also in-medium effects, such as a dynamical modification of vector meson spectral functions by collisional broadening Bratkovskaya and Cassing (2008), and the modification of strange degrees of freedom in line with many-body G-matrix calculations Cassing et al. (2003); Song et al. (2021), as well as chiral symmetry restoration via the Schwinger mechanism for the string decay Cassing et al. (2016); Palmese et al. (2016) in a dense medium. The important and pioneering development in the PHSD is related to the formulation and development of the theoretical formalism in order to realize the detailed balance for reactions based on covariant rates Cassing (2002). This formalism has been implemented in PHSD in Ref. Cassing (2002) for the description of baryon-antibaryon annihilation of and the inverse reaction of multi-meson fusion to pairs; an extension of this first study accounting for all baryon-antibaryon combinations in PHSD has been presented in Ref. Seifert and Cassing (2018). We also mention that the implementation of detailed balance on the level of reactions is realized for the main channels of strangeness production/absorption by baryons () and pions Song et al. (2021).
One of the main goals of this work is to extend this formalism to the and processes, which are relevant for the production and disintegration of deuterons. Therefore, we dedicate a separate section to the detailed description of this formalism and its application to deuteron reactions.
II.2 The “advanced” MST approach (aMST)
II.2.1 The original MST approach
The Minimum Spanning Tree (MST) method has been employed in the PHQMD transport approach in Ref. Aichelin et al. (2020) to identify clusters at different stages of the dynamical evolution of the system. We stress that MST is a cluster recognition procedure, not a “cluster building” mechanism, since PHQMD propagates baryons, not clusters. The possibility to trace back in time the baryons, which combine to clusters due to their potential interaction, allows to investigate more quantitatively the nature of cluster formation, and to answer some fundamental questions, for example how clusters can survive the dense medium Gläßel et al. (2022).
The principle of MST in its original version described in Ref. Aichelin (1991) is to collect nucleons which are close in coordinate space. At a given time a snapshot of the positions and momenta of all nucleons is recorded and the MST clusterization algorithm is applied: two nucleons and are considered as “bound” to a deuteron or to any larger cluster if they fulfill the condition
(6) |
where on the left hand side the positions are boosted in the center-of-mass of the pair. The maximum distance between cluster nucleons fm corresponds roughly to the range of the (attractive) potential. Additional momentum cuts do not change the result because the trajectories of baryons, which are not bound, diverge. Therefore the formation of baryons in MST is a consequence of the attractive potential interaction. A nucleon belongs to a cluster if it is “bound” with at least one nucleon of that cluster, i.e. if there exists a nucleon for which the condition Eq. (6) is fulfilled. Recently, MST has been developed to an independent tool, which can be coupled to any theoretical transport approach or to any theoretical framework for detector calibration Kireyeu (2021).
We want to stress the fact that MST does not lead to the formation of real deuterons. It rather marks only whether a given nucleon is a part of a deuteron at times . During the time step each nucleon continues to propagate according to the QMD EoMs Eq. (2). One can also identify whether in the subsequent time steps the same nucleons form a deuteron and at each time step the binding energy of the deuteron can be calculated. In particular, the binding energy of the produced cluster of size is calculated in its center-of-mass (rest) frame as ; where () is the energy (mass) of the th nucleon of the cluster in the rest frame of the cluster. For its calculation the energy and momentum of nucleons are boosted into this frame from the calculational frame. Even if there are no elastic collisions between one of the cluster nucleons and a third hadron the cluster binding energy can change its sign. This is due to the fact that for the propagation the forces between the nucleons are calculated at the same time in the computational frame. On the other hand, to calculate the binding energy in MST one has to take the positions of the baryons after Lorentz boost into the cluster center-of-mass. However, in this frame the baryons have different times, in principle should be corrected but can hardly be done in practice. The larger the factor between the computational frame and the cluster center-of-mass frame, the more these time differences in the cluster center-of-mass system become important.
In order to overcome this problem of the semi-classical approach, we recall that in our previous study Gläßel et al. (2022) we calculated cluster observables at the “physical time” , which accounts for the time dilatation between the cluster rest frame and the center-of mass system of heavy-ion reaction: , where is the cluster “freeze-out” time at mid-rapidity and is the rapidity of the center-of-mass of the cluster in the calculational frame, the center-of-mass system of the heavy-ion reaction. We called the “physical time” because it marks the identical times in the rest systems. The time was determined such that we could reproduce the total experimental multiplicities of the clusters at mid-rapidity. Also we verified that the choice of affects only the multiplicity and neither the form of the rapidity distribution nor that of the transverse momentum distribution.
II.2.2 “Advanced” MST
This numerical artifact can be surmounted by freezing the internal degrees of freedom of the cluster when it is not anymore in contact (neither by collisions nor by potential interactions) with fellow hadrons, which are not part of the cluster. This freezing can be applied to the “collision history” file which contains the positions and momenta of the baryons as a function of time. Therefore, it does not influence the dynamics of the reaction. This so-called stabilization procedure works as follows and the results are presented in Fig. 1:
-
1)
Nucleons can be part of a cluster only after they have had their last elastic or inelastic collision. At each time step the positions and momenta of all nucleons are recorded and clusters are identified by means of the MST algorithm. This is the standard MST method shown as dashed lines in Fig. 1, for (green line) and (red).
-
2)
Clusters have to have a negative binding energy . Applying this selection on the clusters identified by MST after point 1), the result is shown by dash-dotted lines in Fig. 1. Shortly after the collision starts until kinetic freeze-out, unbound nucleons with quite different momenta can be found at the same position in coordinate space. If time proceeds their trajectories diverge and they do not form a cluster anymore. Indeed, only if the clusters are bound, the nucleons are forced to stay together. Therefore, at late times each dashed line joins the corresponding dash-dotted line.
-
3)
PHQMD conserves energy strictly and the cluster nucleons are maximally separated from the other nucleons with a MST radius of 4 fm. Due to the non-relativistic Skyrme potential, the time shift between the nucleons in the cluster center-of-mass system (where the binding energy is calculated) and an approximation used to extrapolate the interaction density to the relativistic case Aichelin (1991), it may happen that the sign of the binding energy (which is tiny as compared to the total energy of the cluster) changes from negative to positive between the time and the next time step (although the cluster is composed of the same nucleons) and the cluster starts to disintegrate. This disintegration is artificial and, therefore, we preserve the cluster by freezing the internal degrees of freedom.
-
4)
Due to the semi-classical nature of our approach, it may happen that a “bound” cluster with at time step spontaneously disintegrates between and , because the available kinetic energy is given to one nucleon which then can leave the cluster. In a quantum system, where the energy of the ground state is larger than in a semi-classical system (because the wave function cannot have zero momentum), this process is much less probable. Therefore, we consider this evaporation as artificial and restore the cluster of the previous time step . The result, if we include 3) and 4), is shown by the full lines in Fig. 1. We see that at large times the fragment yield becomes stable. Due to the larger factor the freeze-out of the internal cluster degrees of freedom is important at high beam energies. At SIS energies it is almost negligible.

In Fig. 1 the multiplicity of (green lines) and (red lines) clusters at mid-rapidity, , from PHQMD simulations of central Au+Au collisions are shown as a function of time for three different collision energies; from upper to lower panel: a) GeV, b) GeV, c) GeV. The dashed lines correspond to the clusters identified by the original MST as in Ref. Aichelin et al. (2020) according to the description “1)” from above. The dash-dotted lines denote those clusters, which are effectively bound having a binding energy , corresponding to case “2)” from above. The solid lines show the cluster yield obtained with the advanced MST (aMST), i.e. employing after the MST identification the stabilization procedure according to the points “3)” and “4)”. It is clearly visible that MST and aMST give the same cluster multiplicity at low beam energies (top panel) while at higher energies (center and bottom panel) - where the relativistic effects discussed above play a major role - aMST stabilizes the cluster multiplicity. Therefore, it is not anymore necessary to define a time at which the cluster analysis is performed. This represents a remarkable improvement of our previous study in Ref. Gläßel et al. (2022) where we still had to select such a time. This procedure can be applied to any type of cluster of any size, including light nuclei and hypernuclei. In this study we present only the results for deuterons. The study of hypernuclei and heavy clusters we reserve for a future publication.
III Kinetic approach for deuteron reactions
Collision Integral
As described, the collision processes involving the formation and the breakup of a deuteron are implemented in PHQMD by means of a covariant rate formalism which has been firstly developed in Ref. Cassing (2002) and applied within the PHSD microscopic approach in order to study the baryon-antibaryon production by multi-meson fusion Seifert and Cassing (2018); Seifert (2018). Following the steps of Ref. Cassing (2002), we start by writing the covariant Boltzmann transport equation for the single phase-space distribution function of an on-shell hadron
(7) |
using the notation and with the on-shell condition ( is the rest mass). The left hand side of Eq. (7) contains only the free streaming term because we neglect for simplicity any mean-field interaction. On the right hand side the collision integral encodes all the multi-particle transitions which involve the hadron i either in the initial or in the final state. Hence, can be written as a sum over all scattering processes with increasing number of participant particles
(8) |
Each collision term accounts for a particular forward scattering process () with -particles in the initial state and -particles in the final state, as well as for the corresponding backward reaction (). The forward and backward reactions can be grouped together and in collision theory one usually distinguishes a gain and loss term. Therefore, the on-shell collision term becomes
(9) |
In Eq. (III) there are integrals over the initial and final momenta of all particles, excluding the tagged hadron i (the deuteron) whose momentum is denoted as .
The quantity is called transition amplitude and is related to the square of the scattering matrix for the transition where and denote a particular set of allowed discrete quantum numbers for the particles (except the hadron i) in the initial and final states. The function guarantees the energy and momentum conservation in each individual reaction. Finally, the single-particle distribution functions of the hadrons appear, in particular the functions and for the forward/loss term and the function for the backward/gain term . We assign the arbitrary sign to distinguish between the gain and the loss term such that in our study the reaction which leads to the production of a deuteron (playing the role of the tagged hadron i) is associated to the backward/gain term, while the inverse reaction where the deuteron is destroyed, is associated to the forward/loss term. This choice agrees with the original formulation in Ref. Cassing (2002). In the collision integral of Eq. (III) we have also neglected the quantum statistical corrections, i.e. the Pauli-blocking or Bose-enhancement factors, which multiply the product of the phase-space distribution functions, as well as any anti-symmetrization procedure in the transition amplitude for the fermions involved in the reactions. Only the statistical factor , which counts the number of identical particles (either bosons or fermions) in the initial and final states, survives. The expression Eq. (III) can be straightforwardly generalized to the off-shell case by including an additional integration over the energy of each single hadron weighted by the associated spectral function (cf. Ref. Cassing (2002)). In PHQMD/PHSD such an off-shell version of the collision integral is adopted for the dynamical interaction of other baryons and mesons which are propagated with self-consistent off-shell transport equations.
The covariant collision number for the scattering process is the number of forward reaction events in the covariant 4-volume = and, therefore, the covariant collision rate is obtained by dividing the loss term in Eq.(III) by the energy , followed by the integration over the momentum and and a summation over the quantum numbers of the tagged hadron i in the initial state of the transition,
(10) |
Similarly, the covariant collision rate of the backward reaction is obtained from the gain term of Eq. (III)
(11) |
The transition amplitude in Eq. (III) and Eq. (III) is the same because of the equivalence of the scattering matrix under the detailed balance condition for forward and backward processes. Therefore, both expressions can be analytically or numerically solved knowing the dependence of the transition amplitude on the kinematic variables. This has been suggested in Ref. Cassing (2002) where, in particular, it has been shown that the collision probabilities of forward and backward reactions can be written in terms of the corresponding many-body phase-space integrals (cf. Appendix C) if one assumes that the transition amplitude is a function only of the invariant energy of the collision . We apply this procedure for deuteron reactions.
The goal of this work is to implement in the PHQMD transport approach the following deuteron reactions: i) the elastic and reactions, as well as inelastic reactions; ii) the inelastic reactions and with all pion species and .
Employing the covariant expressions in Eq. (III) and (III) with and , the collision rate for the elastic and inelastic reactions can be written as follows,
(12) |
where is the total cross section for a two-to-two scattering process which is defined from the transition amplitude by the well known definition
(13) |
with the flux factor related the (relativistic) relative velocity of the incident on-shell particles of masses and
(14) |
In Eq. (III) the sum is performed over the quantum numbers involved in the reaction and it includes also the statistical factor which we absorb in the cross section.
The PHQMD/PHSD collision integral for the deuteron reactions is solved numerically by dividing the coordinate space in a grid of cells of volume and sampling the on-shell single-particle distribution function at each time step by means of the test-particle ansatz Cassing (2009)
(15) |
where and are, respectively, the position and the momentum of the particle at time . By inserting Eq. (15) in Eq. (12) we obtain the collision probability for the reactions in the unit volume and unit time
(16) |
which depends on through and . Employing sufficiently small values of and we solve numerically the collisions for the deuterons by the stochastic method, i.e. by calculating the invariant energy of each possible reaction and then the associated probability which is confronted with a random number between 0 and 1. To calculate for the inelastic process we use parametrized expressions for the total cross section which are reported in the Appendix A.
Now we describe the inelastic reactions and and, in particular, how the backward reaction can be fully implemented within the covariant rate formalism adopted in our PHQMD approach. On the one hand, this is physically motivated by the fact that these are the dominant reactions for the production of deuterons in HICs due to their large cross sections, , compared to the sub-dominant channel with . On the other hand, it provides an effective method to describe reactions with more than 2 particles in the entrance channel, which cannot be formulated in terms of cross sections as in Eq. (III).
For the forward reaction, the breakup of deuterons by an incident or , the definition of the covariant collision rate follows straightforwardly and is given by Eq. (III) with , and ,
(17) |
where is the total inelastic cross section for either the or the scattering process, which is defined similarly to Eq. (III) with an extra integration over the momentum of the third particle in the final state. The sum over the internal quantum numbers appearing in Eq. (17) is also absorbed in . In Appendix A we provide the parametrized expressions of such inelastic cross sections as a function of and we describe in detail how they are obtained from the experimental measurements of the total inclusive cross section for and collisions. Employing again the test-particle ansatz in Eq. (17), we derive the collision probability for the forward reaction
(18) |
which is a function of , and we sample stochastically the collisions in the unit volume and the unit time for each PHQMD/PHSD parallel ensemble event. When a collision occurs, we construct the final state of three particles in the center-of-mass of the incident pair by means of standard kinematic routines Kajantie and Karimaeki (1971); Byckling and Kajantie (1971).
The covariant rate for the backward and reactions follows from Eq. (III), but we cannot write it in terms of a cross section. However, what is important for us is to obtain a collision probability in order to apply the stochastic method. With the assumption Cassing (2002) the transition amplitude can be moved outside the integration over the momenta of the two particles in the final state. As a result, these integrations can be combined with the function into the two-body phase-space (cf. Appendix C), so that we can write as intermediate step,
(19) |
with the sum running over the quantum numbers and taking into account also the statistical factor for identical particles. Next, we introduce the of the forward process by inverting its definition from the transition amplitude and use again the condition to isolate the three-body phase-space of the initial particles. Hence,
(20) |
where and denote the factors coming from the sum over the spin and isospin quantum numbers in the transition matrix. For the spin contribution we have
(21) |
where in reactions the particles are ordered as follows,
(22) |
For the isospin part a separate discussion is given at the end of the section.
Combining Eq. (III) and Eq. (20) with the test-particle ansatz, we finally obtain the collision probability for the backward process in the unit volume and the unit time ,
where in the second line we employ the collision probability for the forward reaction from Eq. (III). We notice that on the right hand side of Eq. (20) and (III) the energy of the produced particles and appear. That means that in our numerical implementation we have to sample the possible kinematics of the final state before the collisions take place. If the collision occurs, we reconstruct the kinematics of the emitted particles in the center-of-mass of the three interacting initial particles according to our previous sampling. In this sense, we can implement the forward and the backward reactions consistently within the same stochastic model.
In Ref. Oliinychenko et al. (2019) the deuteron reactions and have been implemented in the SMASH transport approach for relativistic HICs. To do this numerically a fictitious resonance was introduced and the process was divided into a two 2-to-2 steps.
Later on, in Ref. Staudenmaier et al. (2021) the same multi-particle production mechanisms have been described according to the covariant rate formalism of Ref. Cassing (2002). In particular, Eq. (6) of Ref. Staudenmaier et al. (2021) shows the same probability for the stochastic treatment of the 3-to-2 process as the one we have just derived in Eq. (III). Comparing both expression we can make some comments:
i) The two- and three-body phase-spaces and appear in both equations as a function of and particle masses. For we employ the well known analytic expression, while for we adopt the parametrization of Ref. Seifert (2018). We collect all formula in Appendix D. In Ref. Staudenmaier et al. (2021) it is done similarly, so we do not expect any discrepancy due to this part.
ii) The probability is proportional to the total cross section for the inverse 2-to-3 process. For deuteron disintegration into 3 particles by and scattering we employ a parametrization of the cross section as a function of , as reported in Appendix A, which is fitted to the available experimental data in the peak region. At high we let our cross section to tend to zero because the 2-to-3 phase-space closes and other inelastic processes with final particles open. In Ref. Staudenmaier et al. (2021) cross sections from Ref. Oliinychenko et al. (2019) are used which differ only for a constant behavior at high . We investigated the possible difference arising from the different asymptotic behavior of the cross sections and we did not find any impact on deuteron yields and the -spectra.
iii) Our spin factor is the same as the one in Ref. Staudenmaier et al. (2021), while the isospin coefficient does not appear there. This represents the novelty of our work.
In Ref. Staudenmaier et al. (2021) as from Ref. Oliinychenko et al. (2019) the -catalysis is considered only for the channel where there is no charge difference between the initial and the final pion, i.e. . We extend the deuteron production to all possible channels which fulfill the conservation of total isospin. We list all implemented channels in the table 1.
-1 | ||
0 | ||
0 | ||
1 | ||
1 | ||
1 | ||
2 | ||
2 | ||
3 |
Since the deuteron has isospin zero, the state is a state with defined isospin 1 provided by the pion. In general, a three particle state has not a definite value of total isospin (i.e. it is not an eigenstate of this quantum number), rather it is formed by a superposition of eigenstates. Therefore, for each channel of the table the represents the projection of the state on the isospin 1 state. We perform the calculation in detail in Appendix C. For the inverse reaction the initial state has total isospin 1. The total cross section describes the reaction of to any of the possible channels. In order to correctly evaluate the disintegration reaction, we weight the transition to one specific channel with the corresponding isospin factor calculated in Appendix C. Similarly we calculate the isospin factors for the -catalysis, where in this case there are no multiple channels available, i.e. does not account for other channels with respect to .
IV Box validation and analytic results
We first study deuteron reactions in the static “box” framework where we can compare the results from our stochastic multi-particle approach with expectations from so-called rate equations. Rate equations differ from the transport approach because they involve the solution of chemical rates of a kinetically equilibrated gas. They have been used for example in Ref. Pan and Pratt (2014) to study the dynamical evolution of baryon-antibaryon annihilation and regeneration by solving fugacity equations or in Ref. Neidig et al. (2022) where the time evolution of light cluster abundancies has been investigated in an expanding medium. In this sense they represent an alternative approach to the covariant rate formalism of Refs. Cassing (2002); Seifert and Cassing (2018).
In a static medium at equilibrium with temperature the rate equations can be taken as analytic reference to verify the correct implementation of the numerical collision criteria. Here we follow a one-by-one comparison with Section B of the work done by the SMASH group in Ref. Staudenmaier et al. (2021) and check the agreement of our results.
As a model we consider the -catalysis reaction with no isospin factors. Using the same notation of Ref. Staudenmaier et al. (2021) we introduce the fugacities for the particle species involved in reactions. Without isospin factor the initial and final pion have the same charge. Therefore, we can see immediately that the number of pions remains constant. Hence, we can write the system of rate equations for and as follows
(24) |
in units of fm-1 and denoting the time derivative as . In Eq. (27) the sum runs over all pions which are initialized in the system according to an equilibrium density at given temperature times a constant fugacity
(25) |
The factors are the spin degenerancies with values
(26) |
Finally, is the cross section for deuteron breakup by an incident pion reported in Appendix A and the thermal average is calculated using the formula
(27) | ||||
which generalizes the expression in Ref. Cannoni (2014) for different particle masses. The time evolution of the nucleon and deuteron density can be directly calculated from the fugacities by
where are the densities at equilibrium at temperature . We set as initial values 0.12 fm-3 and . Provided with these initial conditions Eq. (24) is a system of coupled first order ordinary differential equations (ODEs) which can be solved applying Runge-Kutta methods.

We prepare a cubic box of volume (10 fm)3 in which particles are initially distributed uniformly in coordinate space with a density fm-3 for nucleons and a density fm-3 for pions and in momentum space according to a Boltzmann distribution with temperature . Then, we divide the box volume into unit cells (2.5 fm)3 where deuteron reactions are sampled numerically at each time step fm. We set the parameters and in order to fulfill the main conditions of the stochastic method Lang et al. (1993); Xu and Greiner (2005). In particular, in each unit cell there are sufficient particles to perform collisions with probabilities Eq. (18) and Eq. (III) which must be always smaller than unity.
In Fig. 2 we show the evolution of particles densities as function of time in a static medium at temperature GeV due to the reactions . The labels and the colors in the plot identify the different particles species: red for nucleons , green for deuterons and orange for pions. The solid lines are the solutions from the system of Eq. (24) for nucleons and deuterons using our parametrized form for the cross section plotted in Appendix A. The dashed black line is the expectation value for the deuterons derived with the same rate equations, but employing the parametrized cross section taken from the SMASH study Oliinychenko et al. (2019). The symbols represent the results obtained from box simulations. In particular, for deuterons we show two cases: i) the case where the reactions are solved numerically by means of the multi-particle stochastic approach in both directions (filled circles); ii) the second case where the forward channel is perfomed by means of the geometric criterium where the deuteron collides with a pion and it is disintegrated into final system if it fulfills the condition
(28) |
Here is the distance of closest approach as defined in Ref. Kodama et al. (1984) (see also Ref. Hirano et al. (2013)). The geometric criterium is used to describe many reactions in the original PHSD collision integral, which is also employed within PHQMD. As follows from Fig. 2, both methods - stochastic and geometric criterium - give the same equilibrium values for reactions. We note that more details of the box simulations for the other deuteron reactions, i.e. and , are reported in Appendix B.

In Fig. 3 the detailed balance condition is verified by checking the differential collision rate
as a function of the invariant energy for each implemented scattering process: inelastic (upside down triangles and solid lines); inelastic (squares with dash-dotted lines); inelastic . The symbols, lines and colors for each channel are described in the figure legend.
As follows from Fig. 3, the reaction rate for the forward direction is equal to the reaction rate of the backward direction for all channels. Thus, detailed balance is fulfilled in our calculations for all isospin channels.
The static box calculations show that:
i) the numerically computed densities of protons and deuterons are in a good agreement with analytical results for the equilibrium values;
ii) the stochastic and geometrical methods for reactions agree;
iii) at equilibrium the detailed balance is verified for and reactions.
This ensures the validity of our implementation of the and reactions for deuteron production and absorption within the static box study. We note also that our box results agree with the SMASH calculations Staudenmaier et al. (2021) when considering the same isospin reaction channel with the same cross section.
After the box tests all deuteron reactions are implemented in the PHQMD framework. In particular, the deuteron production by and reactions are sampled stochastically within each PHQMD/PHSD parallel ensemble, while the inverse processes , and the sub-dominant and elastic reactions are performed by means of the geometric criterium described above, in order to speed up the computations.
In contrast to the “box” model, for the stochastic method in realistic HICs with PHQMD we simulate the phase-space evolution of the fireball on an expanding 3D-grid which we divide into cells of volume where fm and fm and the longitudinal expansion of the fireball is taken into account through the factor , where is the velocity of a projectile or target in the cm frame. Inside each cell there are sufficient particles to sample stochastically the deuteron reactions at each time-step .
Moreover, the time-step is initially increasing with time as in order to let particles in each cell evolve smoothly at the beginning of the nucleus-nucleus collision. However, we employ the condition fm/c at later times in order to keep the collision probability below unity.
V Kinetic deuterons in HICs

We start this section by showing in Fig. 4 the collision rates for all inelastic processes for deuteron production (solid green lines) and disintegration (dashed red lines) implemented in PHQMD for two different HIC systems. The top panels (I) show the reaction rates for Au+Au collisions at AGeV (impact parameter fm), while the bottom panels (II) show the reaction rates for Pb+Pb collisions at AGeV (impact parameter fm). Confronting (I) and (II) we clearly see that at the lower collision energy the formation and breakup of deuterons is mainly driven by the channel, involving only nucleons, while at higher energies the reaction becomes dominant because pions are more abundant. The two-body inelastic reaction (right panels) has a much lower rate compared to three-body inelastic (left panel) and (middle panel) channels because the cross section is smaller.
V.1 Effect of charge exchange reactions
Before showing our final results, we study the impact of the different isospin channels on deuteron production in relativistic HICs. The reaction is important only in the case where the pion catalysis is dominant compared to the nucleon catalysis. Therefore, we select Au+Au central collisions at the energy GeV to study the production of deuterons through all the implemented reactions.

Moreover, for this system we can also compare our results with those obtained recently by the SMASH collaboration Staudenmaier et al. (2021) where for deuteron production by multi-particle reactions the same covariant rate formalism is applied. This substitutes the previous SMASH work, where the channel was simulated numerically by a sequence of two processes passing through the formation of a fictitious resonance (Ref. Oliinychenko et al. (2019) for the LHC energy, Oliinychenko et al. (2021) for RHIC BES).
In SMASH studies only the kind of reactions , where the isospin degrees of freedom are not taken into account, were considered within the pion catalysis.
However, isospin conservation allows for two types of pion catalyzed reactions:
, in which the charge is conserved and the charge exchange reaction .
As discussed in the previous section, a goal of this work is to study the impact of including all the charge exchange reaction channels on the production of “kinetic” deuterons in HICs.
In Fig. 5 the mid-rapidity () multiplicity of deuterons in Au+Au collisions at GeV for a fixed impact parameter is shown as a function of time. The full black circle is the STAR measurement Adam et al. (2019) and the full black line is the PHQMD result if we include both types of catalyzed channels.
It is useful to compare our results with those of SMASH in which only the channel (and similar for and ) is included and displayed as red dashed line (taken from Fig. (4)a in Ref. Staudenmaier et al. (2021)). If we omit the charge exchange reaction and retain only what is also employed by SMASH, we find the dash-dotted orange line.
Our result gives a slightly smaller number of deuterons and shows also a different time behavior. The small difference of the order of 10% is not surprising because SMASH and PHQMD are completely different transport approaches. In SMASH, a hydrodynamical evolution of the fireball is followed by a particlization at the hypersurface of an energy density of GeV/fm3. Then particles propagate without potential interaction (cascade) and collide in the hadronic phase until chemical and kinetic freeze-out is achieved. Therefore, in SMASH the kinetic production of deuterons is limited to the latest stages and, more important, it is not affected by the potential interactions among nucleons. On the contrary, in PHQMD the deuteron reactions are embedded in a transport environment in which the baryons are propagated after their creation by the QMD equations, which include potential interactions. If the local energy density exceeds the critical value GeV/fm3, the hadrons dissolve into the partons, which follow the description of the Dynamical Quasi-Particle Model (DQPM) implemented within the PHSD framework (see Sec. II and references therein). Deuteron formation is therefore only possible in regions in which the energy density is smaller than , where hadrons are the degrees of freedom of the system.
Such a different description of the expanding system makes it difficult to disentangle the differences of these two approaches. We mention that, just for test purposes, we have implemented in PHQMD an additional energy density cut for deuteron production, GeV/fm3, mimicking the transition from the hydro to the hadronic phase as in the SMASH approach, however, the results are similar within statistical uncertainties.
Comparing the full black and the orange dot-dashed curve we observe that the charge exchange reaction increases the deuteron yield by 50% (for the isospin factors we refer to Appendix C) at this beam energy and brings the complete calculation outside of the experimental error bars. To complete our study, we made a similar check for collisions at lower energies. In particular, we confirm that at the energy of the GSI-SIS accelerator AGeV, i.e. GeV, where the production of deuterons occurs mainly by (see the collision rate in Fig. 4) the contribution of catalyzed deuteron production is negligible, while at GeV, the energy of the STAR FiXed Target (FXT) experiment, the charge exchange channel increases the deuteron production by 20%.
V.2 Modelling of Finite-size of deuteron
In dense nuclear matter the binding energy of a deuteron is reduced and becomes eventually positive because (for a deuteron at rest and a zero temperature environment) the quantum states with the lowest energy are occupied by protons and neutrons up to the Fermi momentum which is related to the density of nuclear matter by
(29) |
Therefore, only the momentum components above the Fermi surface can contribute to the deuteron binding energy and the expectation value of the deuteron hamiltonian with respect to the pair wavefunction is given by
(30) |
where is the binding energy of the deuteron in nuclear matter. If increases becomes positive and the deuteron becomes unbound. The value of the nuclear density at which the deuteron binding energy vanishes is known as Mott density Röpke et al. (1982); Röpke and Schulz (1988). However, only the case of low-density cold () infinite nuclear matter can be calculated analytically. In the hot fireball ( MeV), created in relativistic HICs at mid-rapidity, deuterons are in addition destroyed by collisions with fellow particles (mostly pions), which scatter with a thermal transverse momentum , i.e. which is much larger than the deuteron binding energy.
Excluded-volume
In collision integrals the final-state particles are considered as point-like particles. In vacuum this is the proper description but in matter, where the final-state hadrons are surrounded by other hadrons, modifications are necessary if the produced particles have a finite extension.
A deuteron with a rms radius of about fm cannot be formed if between the and the other hadrons are located. One possibility to take this into account is to include in our covariant rate formalism an excluded volume condition. As discussed in Sec. III, for the dominant production channels and the probability that the collision occurs and the deuteron is formed is given by Eq. (III). We include the excluded volume condition in our calculation in the following way. When, according to the collision rate, a deuteron should be produced at time , we compute the position and momentum of the center-of-mass of the “candidate” deuteron . Subsequently, we loop over all hadrons, which exist at that time , and for each particle we check the following condition
(31) |
where the parameter is the radius of the excluded volume. The particle positions, and , are calculated in the center-of-mass frame of the candidate deuteron. In order to produce a deuteron at the final state of the reaction the condition Eq. (31) must be fulfilled by all the surveyed particles. Otherwise, the candidate deuteron is considered not formed and the system is restored to the initial state, as if the participant hadrons had never scattered. Thus, like for the MST condition for cluster formation Eq. (6), we use also here a geometrical criterium to take into account the finite extension of the deuteron. However, both criteria work differently: i) in MST and can only be baryons, while for the excluded-volume we account for all spectator hadrons in the fireball, i.e. those particles which do not participate in the initial stage of the deuteron reaction; ii) the excluded-volume condition, which excludes deuteron formation if a third hadron is too close, works oppositely to the MST clustering where two baryons form a cluster if they are sufficiently close. The excluded radius in Eq. (31) is related to the root-mean-square (rms) radius of the deuteron by
(32) |
where is the Deuteron Wave-Function (DWF) in coordinate space, which is obtained by solving the radial Schrödinger equation using a phenomenological parametrization of the nucleon-nucleon potential , which correctly reproduces its ground state properties. In particular, the function takes into account the fact that the deuteron ground state is a mixture of a - and a -states, with assigned real functions and respectively, and it is normalized so that the total probability to find the deuteron in one of the two states is one,
(33) |
More specifically, we employ the DFW parametrization from Ref. McGee (1967), where the functions and can be expressed as discrete superposition of Hankel functions. Similar calculations were performed by the Paris group Lacombe et al. (1980, 1981) using their own parametrization of the potential.


In Fig. 6 the DWF from Ref. McGee (1967) is represented by the solid blue line.
The dashed blue line is the result employing the Paris potential from Ref. Lacombe et al. (1980, 1981). Both are showing the same behavior. Inserting this function in Eq. (32) and solving the integral numerically we find fm (red vertical line).
In Fig. 7 we study the impact of excluded-volume condition on deuteron production for central ( fm) Au+Au collisions at GeV. The full circle is the measured value of the STAR experiment Adam et al. (2019) at mid-rapidity. The lines represent the time evolution of the deuteron yield in PHQMD for the rapidity range . The black solid line is the result if all production channels are included, which has already been shown in Fig. 5. The red lines are the results if we include the excluded-volume condition. Here we present the results for two excluded volume radius parameter, (red thick solid line) and (red dashed line).
As seen from Fig. 7, the inclusion of the excluded volume condition has a large impact on the formation of deuterons - at the considered energy it reduces their abundance at mid-rapidity by a factor of about 3. This is due to the high density of hadrons at mid-rapidity in the initial phase of the reaction.
The final abundance of deuterons depends on . Two choices of the excluded radius, = 1.8 fm and = 2.1 fm - two values around the rms radius of the deuteron - give a difference of the final number of deuterons of 15%.
Momentum projection
As we have seen, the excluded-volume condition models the fact that the deuteron is an extended object in coordinate space with a root-mean-square radius determined from Eq. (32), where is the square of its ground state wave function represented in Fig. 6 with the colored lines. The square of the relative momentum of the bound pair can be obtained from the deuteron wave function represented in momentum space, which can be derived by taking the Fourier transforms of the and state components. The square of the DWF in momentum space , calculated using the Paris potential, Ref. Lacombe et al. (1981), is presented in Fig. 8 as a solid red line and its integral is normalized to unity. Using the uncertainty principle, we can calculate the expected relative momentum of the bound pair
(34) |
For fm, we obtain GeV a value very close to the value calculated using the DWF in Fig. 8 which is about GeV.
The probability amplitude that a proton and a neutron collide and form a deuteron is given by where is the DWF, whose square is shown in Fig. 8, while is the wave function of the relative momentum of a proton and a neutron just after the collision occurred. In both cases, the relative momentum is calculated in the center-of-mass frame of the deuteron. This indicates that a proton and a neutron have the highest chance to form a deuteron in the region where their wave function overlaps most with , which happens if their relative momentum is of the order of 0.1 GeV.

The covariant collision rate for reactions derived in Eq. (III) assumes that the transition amplitude depends only on the center-of-mass energy, . For the deuteron case, this is an oversimplified assumption because nucleons with a very large relative momentum have a small probability to form a deuteron. In order to relax this assumption, we assume that the momentum transfer of the third body is small and that, therefore, in the and reactions, the initial relative momentum of the nucleons is close to that of the two nucleons bound in a deuteron. This allows to determine the probability that a deuteron is produced in these reactions by weighting the initial relative momentum with (which is normalized to unity). Consequently, a pair with a smaller relative momentum in its center-of-mass system has a higher chance to produce a deuteron than a pair with a larger relative momentum.
In the transport calculations we employ a Monte Carlo procedure to decide whether a deuteron is produced or not. If three nucleons are in the entrance channel we randomly determine which of the possible pairs is considered as a possible deuteron candidate. Here we study the effect of this momentum projection on the deuteron production in HICs. In Fig. 9 we display the time evolution of the number of deuterons at mid-rapidity in PHQMD simulations for central Au+Au collisions at GeV, as compared to STAR data at mid-rapidity (black point). The black solid line and red solid line are those from Fig. 7 and represent the deuteron yield including all possible channels, respectively without and with the excluded-volume condition with a parameter fm. The dashed blue line is the result applying the momentum projection only. One can see that momentum projection strongly suppresses the deuteron production at the initial stage of Au+Au collisions where dominantly the collisions of energetic nucleons take place. Moreover, we find that the momentum projection and the excluded-volume condition give for large times the same suppression at mid-rapidity. They both lead to a strong suppression of deuteron formation at mid-rapidity at the time of 10-20 fm, due to the presence of the dense medium populated by many particles (especially pions) which can exist in the volume occupied by the deuteron. At later times deuteron production becomes important but asymptotically the production is only 30% of that without projection.





As a next step, we investigate the case where the two finite-size effects, namely the excluded volume in coordinate space and the momentum projection in momentum space, are simultaneously applied to the production of kinetic deuterons. This scenario is shown in Fig. (9) as the thick dashed green line with full squares. One can see that the inclusion of both conditions produces a suppression which is about a factor 2 stronger than the case where the excluded volume or the momentum projection are applied individually - which means an overall suppression factor of 6 with respect to the case where no finite-size effects are considered (Fig. (9) solid black line).
We studied the impact of finite-size effects on kinetic deuterons at mid-rapidity for different collision systems and found that the amount of suppression is quite similar at all collision energies. Only for top RHIC energy we have noticed a larger factor - about an order of magnitude between the case without and with both effects - which we could explain by the higher density of particles (pions) at the initial stages which makes the excluded volume condition more effective. It is also interesting to study this effect in different rapidity intervals.
In Fig. (10) we present the rapidity distributions of kinetic deuterons from PHQMD simulations in central nucleus-nucleus collisions for four different collision systems, which are reported in the legend. The color coding is the same as in Fig. (9). At the lowest energy, GeV, corresponding to AGeV, where projectile and target decelerate almost completely and form a mid-rapidity source, the maximum of proton as well as the maximum of the deuteron distribution is peaked at mid-rapidity. At the higher energies projectile and target pass each other and the proton as well as the deuteron distributions have a minimum at mid-rapidity.
It is remarkable that the excluded volume and the momentum projection approach leads at all beam energies to an almost identical rapidity distribution around mid-rapidity. Only towards the edge of the rapidity interval, which we investigated here, momentum projection leads to a larger suppression of the deuteron yield because at this rapidity the relative momentum of the nucleons is larger.
The suppression of the deuterons is always of the order of 3 at mid-rapidity. At finite rapidities the momentum projection gives always a larger suppression than the excluded volume approach. If we apply the excluded volume and momentum projection simultaneously we obtain an additional suppression, which is, however, at mid-rapidity small compared to the suppression due to the individual application of one of these finite-size corrections. This is a sign that the relative distance and momentum of the proton-neutron pair are correlated. The form of is close to the one obtained for momentum projection only and is shallower than the distribution without finite-size effects.
We studied also the slope of the -spectra of the kinetic deuterons and found that it is not changed by finite-size effects.
Before moving to the results, we want to mention that:
i) currently, the kinetic deuterons, which are created in collisions, are treated in PHQMD as point-like particles which stream freely until they eventually disintegrate due to collisions with pions or nucleons. This means that they have no potential interaction with other nucleons;
ii) the nucleons of the kinetic deuterons do not enter into the MST algorithm, otherwise they would be double counted. One could argue that nothing prevents a deuteron to interact with a surrounding nucleon and gets bound into a larger cluster. However, this cannot happen when applying the excluded volume condition, as the presence of another nucleon close by would not allow the formation of the kinetic deuteron itself.
VI Results

In this section we compare, from SIS to top RHIC energies, the rapidity and transverse momentum distribution of deuterons, calculated with PHQMD, with the experimental data. As mentioned in the previous sections we consider two sources of deuteron production:
-
•
Deuterons produced by collisions (kinetic deuterons)
Kinetic deuterons can be produced by the inelastic reactions , and . We include all possible charge exchange channels in the catalysis. The quantum properties of deuterons are modeled through the three finite-size corrections, discussed in the last section:-
I)
by the excluded-volume condition choosing the radius fm;
-
II)
by the momentum projection on the DWF ;
-
III)
by taking into account simultaneously I+II.
-
I)
-
•
Deuterons produced by potential interaction (potential deuterons)
The deuterons, which are produced due to the potential interactions between baryons are identified by the ”advanced” Minimum Spanning Tree (aMST) algorithm during the fireball evolution and reconstructed as “bound” clusters () according to the stabilization procedure described in Sec. II B.
VI.1 Au+Au at SIS AGeV
We start by presenting in Fig. 11 the PHQMD results for central Au+Au collisions at AGeV ( GeV) the scaled rapidity distribution as function of , where is the projectile rapidity in the center-of-mass frame of the colliding nuclei. Kinetic deuterons are presented by a thin red line, potential deuterons by a dashed green line and the sum of both as a blue line. The three panels display three different approaches to finite-size effects in the kinetic production via , and . From left to right we display: I) deuterons obtained when applying the excluded-volume condition, II) deuterons obtained when the momentum-projection is employed and III) deuterons if both finite-size corrections are simultaneously considered. The results are compared with the FOPI experimental data Reisdorf et al. (2010), displayed as open circles.
We see in Fig. 11 that the aMST deuterons alone give less than 40% of the measured yield for all scenarios, while the contribution of the kinetic deuterons varies strongly: the scenario II with ”momentum projection” only gives the best description of the experimental data while an additional application of the excluded volume leads to a strong suppression of the deuteron production. At this energy a baryon rich, almost equilibrated fireball is created at mid-rapidity which makes both finite-size corrections very effective. Even together with the aMST deuterons, the deuteron yield is underpredicted.
We note that the underprediction of the cluster multiplicity at mid-rapidity for low beam energies has been already observed in non-relativistic IQMD calculations Le Fèvre et al. (2019). At this low energy the mid-rapidity region is very complex because it contains decelerated projectile and target nucleons as well as fireball nucleons and is characterized by a high baryon density. Indeed, it seems that in our approach some correlations, which contribute to deuteron formation, are absent. We think that further improvement of cluster recognition algorithm as well as improvement of the QMD dynamics (e.g. by using a momentum-dependent potential instead of simple Skyrme potential used in this study) might improve the situation.
VI.2 Au+Au at AGeV
The PHQMD results for the 10% most central Au+Au collisions at a laboratory energy AGeV, corresponding to GeV, are displayed in Figs. 12 - 14. This system will be explored by the future Compressed Baryonic Matter (CBM) experiment at GSI FAIR. Therefore, our results can be considered as predictions until experimental data will be available. However, it is possible to compare our results with the measurements performed at the AGS accelerator for the asymmetric Au+Pb collisions at the same beam energy.
We note, that in our previous study Gläßel et al. (2022) the multiplicity and -spectra of the light nuclei , , were presented in Au+Pb collisions for the same beam energy and compared to the same data from the E864 experiment Armstrong et al. (2000). As mentioned above the clusters were identified by the original MST approach described in Sec. II B.1. In that case the number of clusters was shown to decrease as a function of time due to the instabilities originating from the semi-classical nature of the QMD approach. Therefore, it was necessary to introduce a physical time for the identification of clusters which was around 50 fm/c for deuterons and 60 fm/c for tritons and .
In Fig. 12 the rapidity distributions of kinetic deuterons (solid red line), potential deuterons identified with aMST (dashed green line) and the sum of the two contributions (solid blue line) are compared to the data from the E864 collaboration Armstrong et al. (2000) (open circles). For the PHQMD results we apply the same selection in transverse momentum, GeV, as in the experiment. Similarly to Fig. 11, the different approaches to model the finite-size of the deuteron for the kinetic deuterons are separated in three different panels.
The excluded-volume condition (left panel I) and momentum projection (center panel II) give about the same amount of kinetic deuterons at mid-rapidity in agreement with the experimental yield, but they start to overestimate them at larger rapidities. When both effects are applied (right panel III), the shape of the rapidity distribution agrees better with the data points, even though the total yield is slightly underestimated. This is due to the
aMST deuterons, which have a shallower distribution.
In Fig. 13 we show the transverse momentum distribution of deuterons as function of for the same collision system and at three different rapidity intervals indicated in each plot.
The color coding is the same as in Fig. 12. We observe that the -slope of kinetic deuterons is insensitive to the modelling of finite-size effects and combined with the potential deuterons give a good description of the trend of the experimental data.



To conclude the analysis at this collision energy we calculate the covariant coalescence function which is defined by the formula
(35) |
for deuterons with momentum , where is the momentum of the free nucleon. we present in Fig. 14 as function of the transverse momentum and confront it with the experimental data from E864 collaboration Armstrong et al. (2000). We refer to Ref. Gläßel et al. (2022) for the details.









As in our previous calculations in Ref. Gläßel et al. (2022) the coalescence function of deuteron shows a quite flat behavior as function of which is also observed in the experimental data, apart from the strong increase at large in the interval . Moreover, the obtained from PHQMD simulations seems to be quite independent from the modelling of finite-size effects in the kinetic production. In particular, when either excluded-volume condition (solid lines) or -pair momentum projection (dotted lines) are applied separately, the results are practically the same, while when both effects are simultaneously taken into account (dashed lines) a smaller is observed due to the stronger suppression of kinetic deuterons at large rapidities. In Fig. 14 the PHQMD calculations refer to the sum of the contribution of kinetic and potential deuterons. The latter are identified via the advanced MST (aMST).

VI.3 SPS energies
We step up in energy and present the results of the PHQMD approach for heavy-ion collisions in the CERN SPS energy range. The rapidity distribution of kinetic and potential deuterons in central Pb+Pb collisions are shown in Fig. 15. The columns represent our three different options to model finite-size effects in the collisional deuteron production. From left to right we display: I) excluded-volume, II) momentum projection, III) both together. The color coding is the same as in Fig. 11. The rows collect results for the full energy range of the SPS facility; from top (a) to bottom (e) we see: 20, 30, 40, 80, 158 GeV per nucleon. The dots are the experimental data from the NA49 collaboration Anticic et al. (2016).
By comparing the first and the second columns, we observe that even though the excluded-volume and the momentum projection give a similar suppression of deuterons at mid-rapidity (Fig. 9), at target-projectile rapidity their effect on deuteron is quite different. There mostly spectator nucleons are localized whose density in central collisions is not very high. Their momentum distribution is close to the Fermi distribution because these baryons have not scattered or scatter with a small momentum transfer. Therefore, the excluded-volume prescription of deuteron production gives less suppression than the projection on the deuteron wave function. Combining both prescription lowers the deuteron production further (column III). The rapidity distribution of kinetic deuterons is narrower at mid-rapidity than the experimental data. If one adds the deuterons created by potential interactions (the sum is presented by the full blue line), which is wider than that of the kinetic deuterons, the experimental multiplicity as well as the rapidity distribution of deuterons, measured by the NA49 collaboration Anticic et al. (2016), is nicely reproduced.
The transverse momentum distributions of deuterons at mid-rapidity are shown in Fig. 16 with the same color coding and panel structure as in Fig. 15. For each energy the rapidity interval for the -spectra is indicated in the right column and taken in correspondence with the NA49 experimental data Anticic et al. (2016). While the excluded-volume (left column) and the momentum projection (middle column) give roughly the same suppression of kinetic deuterons, both effects combined yield an additional suppression factor of 2, as already seen in the rapidity distribution. The form of the -spectra is for all three approaches to model the finite-size effects the same.


VI.4 RHIC BES energies
PHQMD is designed to describe clusters also at higher energies. Therefore, we can study the deuteron production in the full energy range of the RHIC Beam Energy Scan (BES). The STAR collaboration has measured in this energy region the multiplicity at mid-rapidity, as well as the mid-rapidity distribution Adam et al. (2019). Here we present the results for model III, which accounts for both finite-size effects: in coordinate by an excluded-volume with radius fm and in momentum space by the projection on the DWF. The kinetic deuterons are supplemented by the potential deuterons, calculated with the advanced MST (aMST) method. We start out by showing the excitation function of the deuteron yield at mid-rapidity, which has been already studied in Gläßel et al. (2022) for the standard MST deuteron recognition algorithm. It is displayed in Fig. 17 for central Au+Au collisions as function of . The black points represent the data at mid-rapidity, measured by the STAR experiment Adam et al. (2019). The lines are the PHQMD results for the same rapidity interval, . The color coding is the same as in Fig. 11. The combined PHQMD results are in quite good agreement with the data, giving slightly less deuterons at the lowest RHIC BES energy GeV and slightly more deuterons at the top RHIC energy, GeV, compared to experimental data points.
It is visible that at these energies the kinetic contribution is small compared to that of aMST (roughly by a factor of 3 less) and, therefore, the multiplicity is dominated by potential deuterons. The STAR collaboration has also measured the ratio at mid-rapidity and, hence, the fraction of mid-rapidity protons which is bound in the lightest cluster. It is shown in Fig. 18 as a function of the NN center-of-mass energy. Again we separate kinetic and potential deuterons. In this energy regime, as we have already seen, the kinetic deuterons contribute only around 30% to the total yield. It is remarkable and unexpected that the form of the excitation function for kinetic and potential deuterons is very similar. If we add kinetic and potential deuterons we overpredict above GeV this ratio by about 30% what represents almost exactly the contribution of kinetic deuterons.
In Fig. 19 we present the transverse momentum distribution of deuterons as function of for the same central Au+Au collisions and for the same mid-rapidity interval . The color and symbol coding is identical to the one used in Fig. 17. Again the combined deuteron yield overpredicts the experimental result by roughly 30%. One can observe that the -spectra of kinetic (thin solid red line) and aMST (dashed green line) deuterons have a quite similar shape and the total -spectra (thick solid blue line) is in good agreement with the measured spectra in the wide energy range of RHIC BES.



VII Conclusions
In this study we have investigated the production of deuterons in nucleus-nucleus collisions within the PHQMD microscopic transport approach. We have focused on the description of deuteron rapidity and transverse momentum distributions around mid-rapidity from SIS AGeV up to top RHIC GeV, covering essentially the whole energy range of relativistic HICs. In the PHQMD framework we have studied two possible mechanisms for the dynamical formation of deuterons - by collisions and by potential interaction. The results can be summarized as follows:
-
•
“Kinetic” mechanism:
i) We have implemented , and the sub-dominant by means of the covariant rate formalism Cassing (2002) in the PHQMD collision integral. The numerical implementation of the and reactions has been tested by the stationary “box” calculation in comparison with the solutions of the rate equations. Moreover, it has been verified that the detailed balance condition is fulfilled for each isospin channel.ii) Differently to the previous study by the SMASH group Staudenmaier et al. (2021), we have accounted in the main and reactions for all possible reaction channels which are allowed by the conservation of total isospin. We have found that the inclusion in the catalysis - which at high collision energies is more dominant than -catalysis due to the large pion abundance - of all charge exchange reactions enhances the production of “kinetic” deuterons by about 50% at mid-rapidity for GeV STAR BES energy, while at GeV, the energy of the STAR FXT experiment, the charge exchange channels increase the deuteron yield by 20%.
iii) The deuteron, being an extended quantum object in coordinate space with a small relative momentum cannot be produced as a point-like hadron. We have taken this into account by two approaches: I) an excluded-volume condition which suppresses the formation of deuterons in the presence of surrounding hadrons and II) by a projection of the relative momentum of the -pairs on the deuteron wave function. We have shown that the inclusion of each of these finite-size effects leads to a significant but similar suppression of deuterons at mid-rapidity. Applying both effects together the deuteron yield is suppressed by an additional factor of two.
-
•
“Potential” mechanism:
We have extended our study of deuteron formation by “potential” interaction between nucleons in Refs. Aichelin et al. (2020); Gläßel et al. (2022), where we had to determine a time at which the cluster recognition by the Minimum-Spanning-Tree (MST) approach has been performed. The newly developed “advanced” MST (aMST) method stabilizes the clusters, which in semi-classical approaches are not stable and whose stability suffers from the Lorentz transformation between the cluster center-of-mass system (where the binding energy is determined) and the nucleus-nucleus center-of-mass, i.e. the computational frame. In aMST clusters, which are bound, cannot disintegrate after the constituents had their last collision and are outside the range of the potential interaction of other clusters and nucleons. In aMST the clusters are stable and therefore no time has to be determined at which we analyse them. -
•
As found in our previous studies Gläßel et al. (2022); Kireyeu et al. (2022), the clusters - produced via potential mechanisms - are created after the fast hadrons have already escaped from the reaction zone, i.e. clusters remain in transverse direction closer to the center of the heavy-ion collision than free nucleons. The “kinetic” deuterons analyzed here follow the same tendency. Thus, since the “fire” is not at the same place as the “ice”, clusters can survive, what solves the “ice in the fire” puzzle.
We have found that the PHQMD approach with the two mechanisms of deuteron production, consistently combined, provides a good description of the large set of available experimental data from SIS Reisdorf et al. (2010), NA49 Anticic et al. (2016) and STAR Adam et al. (2019) collaborations. Finally, we mention also that our results for Au+Au collisions at AGS and RHIC BES energies are relevant for the future experiments which will be carried out at the FAIR and NICA facilities.
Acknowledgements
The authors acknowledge inspiring discussions with M. Bleicher, W. Cassing, H. Elfner, I. Grishmanovskii, C.-M. Ko, D. Oliinychenko, L. Oliva, T. Reichert, O. Soloveva, T. Song, J. Staudenmaier, J. Steinheimer, Io. Vassiliev. Furthermore, we acknowledge support by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) grant BL982-3 by the Russian Science Foundation grant 19-42-04101 and by the GSI-IN2P3 agreement under contract number 13-70. This study is part of a project that has received funding from the European Union’s Horizon 2020 research and innovation program under grant agreement STRONG – 2020 - No 824093. The computational resources have been provided by the Center for Scientific Computing (CSC) of the Goethe University and the ”Green Cube” at GSI, Darmstadt.
Appendix A: Cross sections
In this appendix we make a collection of scattering cross sections for the reactions of production and breakup with nucleons and pions which have been implemented in this work.
I. Elastic processes and are characterized by cross sections of the order of for scattering Norem (1971). We use the parametrization of elastic cross sections as function of invariant center-of-mass energy reported in Ref. Oh et al. (2009) (see Appendix there and reference therein). Deuterons can be inelastically produced in collision with projectile energy of the order of GeV and accompanied pion emission. Conversely, a projectile pion with beam energy hitting a deuteron target can be absorbed and breakup the deuteron into a final pair without pion emission in the final state. The complete reactions represent a two-body inelastic process of production and disintegration. The total cross section for this reaction in both directions have been extensively analyzed within isospin decomposition formalism VerWest and Arndt (1982); Arndt et al. (1997); Oh et al. (1997) and it has been proven that the detailed balance condition is fulfilled. This allows to relate cross section of forward and backward reactions as follows
(36) |
where and are respectively the nucleon and pion momentum computed in the center-of-mass frame of the particles pair, which by simple kinematics can be written in terms of masses and , respectively,
(37) |
for the nucleon momentum and
(38) |
for the pion momentum . In Fig. 20 the cross section for as function of , which we also take from Ref. Oh et al. (2009), is shown (red dashed line) in comparison with previous calculations VerWest and Arndt (1982) (black solid line) as a function of the kinetic energy of the projectile proton , where is the beam momentum, which is used to calculate the corresponding value of according to the formula
(39) |
in the laboratory frame where the target proton is at rest. The black points refer to some experimental measurements of in the range of the peak Anderson et al. (1974); Shimizu et al. (1982). Still in Fig. 20 the cross section for inverse process, i.e. inelastic two-body breakup by incident into pair with no pion emission in the final state, obtained from detailed balance condition Eq. (36) is shown with blue dash-dotted curve. We can use same arguments to implement the reaction , but for the case we have to account for an extra isospin factor and write
(40) |
The factor is due to the fact that in terms of isospin base the pair is an anti-symmetric superposition of isospin triplet () and singlet () with three-component projection , while on the other hand the final depends only on the pion isospin, hence it is a pure triplet state.

As can be seen from Fig. 20 the experimental two-body cross section for breakup into -pair without pion emission in the final state is quite small and of the order of mb at the peak.
In Fig. 21 the measured inclusive total cross section for scattering from the Particle Data Group Tanabashi et al. (2018) are shown, respectively, with orange and red marks with as a function of . We focus on the peak region where the cross section reaches a value of the order of mb. Therefore, the inelastic two-body channel shown in Fig. 20 exhausts less than 5% of the total inclusive cross section. This comparison demonstrates the necessity to implement inelastic processes for breakup by energetic involving more than 2 particles in the final state and consequently the inverse process where a deuteron can be formed by the interaction of two nucleons catalyzed by colliding pions. A theoretical study to understand why the cross section is larger than the typical inelastic hadronic cross section has been conducted in Ref. Ikeno et al. (2021).
II. In this work we consider the reaction within the PHQMD transport approach where two-body and, more importantly, three-body processes are treated using the convariant rate formalism Cassing (2002). In this formalism the only requested input is a parametrization of the total cross section as function of since the transition rate should depend only on the invariant energy which means to assume an isotropic differential cross section. Here we describe the procedure to extract such phenomenological cross section.
The inclusive cross section is strongly peaked at GeV due to excitation of underline isospin channel which, as we have just said, is possible both for and due to the presence of a proton and a neutron in the deuteron target. Referring to Fig. 21, we firstly perform a fit (violet curve) of this total inclusive cross section using the following piecewise expression
(41) |
The center-of-mass energy is in GeV and the is given in mb. For GeV we simply take constant mb. The values of the fit parameters are reported in the upper part of Table (2).
1 | 186.690 | 4.767 | 0.042 |
---|---|---|---|
2 | 16.765 | 7.356 | 0.174 |
3 | 12.907 | 8.808 | 1.282 |
42.586 | 2.009 | ||
15543.600 | 1145.460 | 504.896 | |
4.651 | 0.203 | ||
1 | 143.415 | 4.779 | 0.030 |
2 | 49.652 | 5.587 | 1.603 |
In Fig. 21, subtracting the elastic (green dashed line) and two-body inelastic (blue dash-dotted line) contributions, we can infer the inelastic cross section for deuteron breakup into 2 baryons + mesons which we assume can be only pions with
(42) |
In particular, the two baryons can be regarded as only nucleons plus excitation of resonances which further decay into , hence feeding the number of final pions. Inelastic processes with increasing body add up subsequently with increasing value of where the sum runs over the masses of final produced particles. On the other hand, this causes the closure of the phase-space of few-body production at larger values of . Keeping this in mind, we can estimate the behavior of the leading inelastic process where deuteron breaks up by the incident pion into a pair of nucleons plus a single emitted pion. The cross section for such leading inelastic process is parametrized by the 2-Gaussian expression below
(43) |
The values of the fit parameters are reported in the lower part of Table (2) (inel.) with the cross section in Eq. (43) given in mb. Finally, the resulting inelastic cross section is shown in Fig. 21 by the squared-thick black line.

III. Similarly to the reaction which dominates at relativistic energies due to the large pion abundance, we implement also the three-body inelastic process for which is more important at low energy heavy-ion collisions Kapusta (1980); Siemens and Kapusta (1979). We derive a parametrization of the total inclusive cross section through a fit on experimental data which are shown in Fig. 22 for experimental data (black points) and for some data (grey points) taken from the PDG database Tanabashi et al. (2018) and from other references Bugg et al. (1966). The lower horizontal axis is the range of the proton beam momentum , while the upper one is the corresponding calculated by relativistic kinematics. The total inclusive cross section is composed by an elastic and an inelastic part
(44) |
In this case the elastic and inelastic contributions are distinctly separated in kinematics due to the existing energy threshold for the processes which is always an endothermic process for any final number of pions . In particular, for the reaction with , this threshold corresponds to the difference . Replacing the deuteron mass , this threshold energy corresponds to nothing else than the absolute value of deuteron binding energy MeV.

Our expression for the total inclusive cross section is derived applying similar procedure to what is done for the case of cross section. In the low energy regime we employ a functional expression in terms of the projectile nucleon momentum which is used in Ref. Cugnon et al. (1996) to parametrize the inelastic scattering cross section. Then, our calculated expression is welded to a parametric function of which we us in the high energy regime. In PHQMD this procedure is applied also for parametrizing the scattering cross section. The resulting curve for is depicted in Fig. 22 with an orange line (dashed for the low energy regime, solid for the high one) and it provides a good fit of the experimental data (blue points) taken from PDG database Tanabashi et al. (2018). In Fig. 22 this function is denoted as “ PHSD parametrization”, because it is implemented in PHQMD directly from the original (P)HSD original code Cassing and Bratkovskaya (2008); Cassing (2009); Cassing and Bratkovskaya (2009); Bratkovskaya et al. (2011); Linnyk et al. (2016) where it is used to describe nucleon-nucleon collisions. For the case of scattering the complete expression for the inclusive total cross section is reported below and the resulting curve is shown in Fig. 22 with a solid green line
The low-energy parametrization of the total inclusive cross section (in mb) as a function of is valid within the range of GeV which corresponds to a value of GeV. Instead, the high-energy parametrization which is valid for and it is given by the following formula
(46) | |||||
Finally, in order to pass from the total inclusive to the total inelastic cross section, we perform a smooth cut of the expression Eq. (46). This is motivated by the fact that at high values inelastic channels with final particles multiplicity larger than 3 (i.e. ) start to contribute and the three-body phase-space is suppressed. Therefore, similarly to what we have done for the reaction, we consider a gaussian function with parameters which we attach to the inclusive cross section formula Eq. (46) at GeV. The resulting curve is depicted in Fig. 22 (thick black solid line). At lower values of the cross section for inelastic reaction equals the total inclusive . as it is clearly visible in Fig. 22 where the two lines are superimposed.
Appendix B: Box simulations
Additionally to the pion catalysis process considered in Sec. IV, in this Appendix we present the box study for other reactions for deuteron production:
i) reaction of nucleon catalysis - , which plays a dominant role at low collision energies where the nucleon density is high (see Fig. (4));
ii) reaction - , which are a subdominant channel due to the small cross section.
In Figs. 23 and 24 we show the time evolution for the density of nucleons (red), deuterons (green) and pions (orange) for simulations performed in a static box at initial temperature GeV and initial nuclear density and pion density in comparison with analytic results obtained as solution of the rate equations. These two plots are similar to Fig. 2 of Sec. IV, where we tested the correct numerical implementation of process.
In particular, in Fig. 23 we test the formation/breakup of deuterons by nucleon catalysis by switching off all other reactions and verify that the collisions inside the box performed by stochastic method (circles) are in agreement with analytic expectations (solid lines). Here the result for pions is not shown, because they are not involved in the main reaction channel, but they can scatter elastically with deuterons in order to drive them to faster equilibration.


In Fig. 24 we investigate the correct implementation of the two-body inelastic process. The analytic expectations of densities as function of evolution time are represented with the lines (red solid for , green solid for and orange dash-dotted for ). For nucleons and deuterons we show box results obtained adopting either the geometric criterium (open squares) or the stochastic collision method (full squares) with cross section taken from Ref. Oh et al. (1997). For pions we show only the results employing the latter collision method. In this case the particle densities show a longer equilibration time compared to the box results where either -catalysis or catalysis reactions are switched on. This is expected because the cross section of disintegration is much smaller than those of other reactions. Consequently, the inverse process of fusion into final state, whose cross section is experimentally proved to fulfill detailed balance (see Eq. (36)), slowly proceeds with an equilibration time about 10 times larger than that of catalysis.
As follows from Figs. 23 and 24 the simulated density distributions in the box are in a good agreement with the analytic thermodynamic results.
We remind here that the “box” framework is just a toy-model to test in controlled conditions that the detailed balance as well as the agreement with the analytic solutions of the rate equations (see Sec. IV) is fulfilled for all the implemented “kinetic” deuteron reactions, in particular , and . In realistic HIC simulations with PHQMD we describe the phase-space evolution of the fireball on an expanding 3D-grid where the cells and the time-step parameters are described at the end of Sec. IV.
Appendix C: Isospin factors in deuteron reactions
In detail, we derive the isospin coefficients for the -catalysis reactions listed in Tab. (1). We look first at the reaction in the forward direction, i.e. to the deuteron disintegration by incident . Since the deuteron has zero isospin and rhe pion has isospin 1, the 2-particle state is already an eigenstate of total initial isospin and the projection along the quantized axis are related to the pion charge indicated by the index . In short notation we can write
(47) |
To calculate the possible eigenstates of total final isospin from the 3-particle state we use the rules for summation of angular momentum in quantum mechanics Sakurai and Napolitano (2020).
Firstly we perform the summation of the -isospin of the two nucleons, which - as is well known - generates the singlet state
(48) |
and the triplet state
(49) | ||||
where indicates the eigenvalue of total isospin of the intermediate two-nucleons system and its projection along the quantized axis. Then we add the contribution of the isospin quantum numbers of the pion which we indicate with small letters as and .
Therefore, the eigenstates of total isospin are given by the following expansion
(50) |
On the right hand side, the ket of the basis represent the tensor product of the two-nucleons state with defined isospin quantum numbers with the pion state . On the left hand side, the new basis is formed by the ket with defined quantum numbers . The angular momentum rules require that which is encoded already in the Clebsch-Gordan coefficients of the expansion Eq. (Appendix C: Isospin factors in deuteron reactions).
By construction, the new states of total isospin will remain eigenstates of the two-nucleons isospin . Hence, formula Eq. (Appendix C: Isospin factors in deuteron reactions) must be separated into two orthogonal contributions according to the two different values of from the singlet Eq. (48) and from the triplet state Eq. (Appendix C: Isospin factors in deuteron reactions).
-
I)
When the two nucleons are in the singlet case , the condition is similar to the 2-particle state , i.e. the total final isospin has eigenvalue from the pion contribution. In the same notation of Eq. (47) we can write
(51) -
II)
In case the two nucleons couple two a triplet state the addition of the pion isopsin generate independent spaces of total final isospin with eigenvalues given by the triangle rule
(52) Due to the conservation of total isospin in strong interaction, we are interested in the eigenstates of final isospin . Using Eq. (Appendix C: Isospin factors in deuteron reactions) with the correct quantum numbers and taking the corresponding Clebsch-Gordan coefficients from available tables, we obtain
(53) (54) (55)
Eq. (I)) and (53)-(55) express the eigenstates in terms of the 3-particles states . Combining them the associated isospin factors for the transition are nothing else than the Fourier coefficients of the following expansion
(56) |
(57) |
(58) |
An overall factor is introduced in order to guarantee the normalization of the corresponding final state to unity. Differently, in calculating the probability of each transition allowed by isospin conservation one should divide by the sum of the squares of all the Fourier coefficients. Finally, the associated isospin probability for each channel are given by the square of such coefficients. Moreover, since in PHQMD the isospin number is not stored during the dynamical evolution of particles, we can make a further simplification by adding the probabilities for those 3-particle states on the right hand side of Eq. (Appendix C: Isospin factors in deuteron reactions)-(Appendix C: Isospin factors in deuteron reactions) which differ only by the order of nucleons. Therefore, we obtain
which, as discussed in Sec. III, correspond to the factors employed to select the final state of each collision where a deuteron disintegrates by inelastic pion reaction according either to the geometric criterium or to the stochastic method, both depending on the total cross section described in Appendix A. It is important that the sum of all probabilities over the possible final state , to which the initial state , can decay according to total isospin conservation equals to one. This condition is clearly fulfilled by summing the terms in each row of Eq. (Appendix C: Isospin factors in deuteron reactions).
In the backward direction, i.e. when an incident catalyzes the fusion of two nucleons to form a deuteron plus an emitted pion, the transition from the 3-particle state into the 2-particle state can happen only if total isospin is conserved. This means, that only when the initial state finds itself in an eigenstate of total isospin 1, it can make the transition to the final state . Formally, we need to invert the basis expansion Eq. (Appendix C: Isospin factors in deuteron reactions) and obtain the following
(60) |
Physically, each 3-particle state is written as a superposition over the eigenstates of total isospin with eigenvalues again provided by the same triangle rule Eq. (52) and eigenvalue of the isospin 3-component . In Eq. (60) the same Clebsch-Gordan coefficients of Eq. (Appendix C: Isospin factors in deuteron reactions) appear because
Once we calculate all the eigenstates of total isospin, we can look at the Fourier coefficient of each state associated to the eigenavalue and obtain the right probability. We point out, that the use of Eq. (60) requires also that the intermediate total isospin of the two-nucleons system is a quantum number of the basis. Therefore, the contributions which come from the singlet and the triplet state must be squared and independently summed without mixing term.
Then, for example, the probability that the state has total isospin is given by
(61) |
where the first term comes from the singlet and the second one from the triplet state. Analogously we can calculate for the other 3-particle states and obtain the probabilities
which named shortly as in the formula Eq. (III) for the full covariant probability of for the reaction of deuteron formation by -catalysis .
The isospin coefficients for the -catalysis reactions can be calculated in similar way. We emphasize only two differences. On the one hand, the final state of the forward reaction, i.e. the deuteron disintegration, contains only nucleons. As mentioned previously, in PHQMD the particle isospins are not propagated dynamically during the system evolution. This means, that we cannot distinguish between those states of the reactions which differ by the order of protons and neutrons. Therefore, the transition from to reduces to one channel and consequently
(63) |
On the other hand, the isospin factors for the backward reaction, i.e. the formation of deuteron by two-nucleons plus third nucleon as catalyzator, demands that the initial state finds itself in an eigenstate of total isospin . We apply the addition rules of angular momentum and construct the intermediate singlet and triplet states by summing isospins of two nucleons and then add the -isospin of the third one. The transition probabilities can be derived from the square of the Fourier coefficients taking the contribution of the singlet and the triplet without mixing term. We obtain
(64) |
which corresponds to the coefficient in Eq. (III) for the deuteron production by -catalysis.
Appendix D: Phase-Space integrals
In Eq. (III) the Lorentz invariant two-body and three-body phase-spaces appear. For we simply adopt its analytic expression
(65) |
where on the right hand side the kinematical function is and is the center-of-mass energy. For we use the well known recursion relations Kajantie and Karimaeki (1971); Byckling and Kajantie (1971) to write it in terms of a factorized product of two-body phase-spaces
The integration is run over the invariant mass variable . It has been shown in Ref. Seifert (2018) that such expression can be fitted by the following function
(67) |
where and the parameter values depend on the physical masses involved. For the and the three-body phase-spaces these parameters are reported in Table (3).
0.000249 | 1.847779 | 0.152221 | 0.071509 | 9.973413 | |
0.000350 | 1.781741 | 0.218259 | 0.052836 | 4.221995 |
References
- Busza et al. (2018) W. Busza, K. Rajagopal, and W. van der Schee, Ann. Rev. Nucl. Part. Sci. 68, 339 (2018), eprint 1802.04801.
- Jacak (2021) B. V. Jacak, Nucl. Phys. A 1005, 122052 (2021).
- Borsanyi et al. (2014) S. Borsanyi, Z. Fodor, C. Hoelbling, S. D. Katz, S. Krieg, and K. K. Szabo, Phys. Lett. B 730, 99 (2014), eprint 1309.5258.
- Bazavov et al. (2014) A. Bazavov et al. (HotQCD), Phys. Rev. D 90, 094503 (2014), eprint 1407.6387.
- Asakawa and Yazaki (1989) M. Asakawa and K. Yazaki, Nucl. Phys. A 504, 668 (1989).
- Stephanov (2004) M. A. Stephanov, Prog. Theor. Phys. Suppl. 153, 139 (2004), eprint hep-ph/0402115.
- Westfall et al. (1976) G. D. Westfall, J. Gosset, P. J. Johansen, A. M. Poskanzer, W. G. Meyer, H. H. Gutbrod, A. Sandoval, and R. Stock, Phys. Rev. Lett. 37, 1202 (1976).
- Gutbrod et al. (1976) H. H. Gutbrod, A. Sandoval, P. J. Johansen, A. M. Poskanzer, J. Gosset, W. G. Meyer, G. D. Westfall, and R. Stock, Phys. Rev. Lett. 37, 667 (1976).
- Nagamiya et al. (1981) S. Nagamiya, M. C. Lemaire, E. Moller, S. Schnetzer, G. Shapiro, H. Steiner, and I. Tanihata, Phys. Rev. C 24, 971 (1981).
- Grunder et al. (1971) H. A. Grunder, W. D. Hartsough, and E. J. Lofgren, Science 174, 1128 (1971).
- Ahle et al. (2000a) L. Ahle et al. (E866, E917), Phys. Lett. B 476, 1 (2000a), eprint nucl-ex/9910008.
- Ahle et al. (2000b) L. Ahle et al. (E866, E917), Phys. Lett. B 490, 53 (2000b), eprint nucl-ex/0008010.
- Bennett et al. (1998) M. J. Bennett et al. (E878), Phys. Rev. C 58, 1155 (1998).
- Saito et al. (1994) N. Saito et al. (E886), Phys. Rev. C 49, 3211 (1994).
- Armstrong et al. (2000) T. A. Armstrong et al. (E864), Phys. Rev. C 61, 064908 (2000), eprint nucl-ex/0003009.
- Reisdorf et al. (2010) W. Reisdorf et al. (FOPI), Nucl. Phys. A 848, 366 (2010), eprint 1005.3418.
- Anticic et al. (2004) T. Anticic et al. (NA49), Phys. Rev. C 69, 024902 (2004).
- Anticic et al. (2016) T. Anticic et al. (NA49), Phys. Rev. C 94, 044906 (2016), eprint 1606.04234.
- Adam et al. (2019) J. Adam et al. (STAR), Phys. Rev. C 99, 064905 (2019), eprint 1903.11778.
- Adam et al. (2016) J. Adam et al. (ALICE), Phys. Rev. C 93, 024917 (2016), eprint 1506.08951.
- Acharya et al. (2017) S. Acharya et al. (ALICE), Eur. Phys. J. C 77, 658 (2017), eprint 1707.07304.
- Dönigus et al. (2022) B. Dönigus, G. Röpke, and D. Blaschke, Phys. Rev. C 106, 044908 (2022), eprint 2206.10376.
- Andronic et al. (2011) A. Andronic, P. Braun-Munzinger, J. Stachel, and H. Stocker, Phys. Lett. B 697, 203 (2011), eprint 1010.2995.
- Bratkovskaya et al. (2022) E. Bratkovskaya, S. Glässel, V. Kireyeu, J. Aichelin, M. Bleicher, C. Blume, G. Coci, V. Kolesnikov, J. Steinheimer, and V. Voronyuk, in 20th International Conference on Strangeness in Quark Matter 2022 (2022), eprint 2208.11802.
- Oliinychenko et al. (2019) D. Oliinychenko, L.-G. Pang, H. Elfner, and V. Koch, Phys. Rev. C 99, 044907 (2019), eprint 1809.03071.
- Cleymans and Satz (1993) J. Cleymans and H. Satz, Z. Phys. C 57, 135 (1993), eprint hep-ph/9207204.
- Andronic et al. (2013) A. Andronic, P. Braun-Munzinger, K. Redlich, and J. Stachel, Nucl. Phys. A 904-905, 535c (2013), eprint 1210.7724.
- Andronic et al. (2018) A. Andronic, P. Braun-Munzinger, K. Redlich, and J. Stachel, Nature 561, 321 (2018), eprint 1710.09425.
- Sombun et al. (2019) S. Sombun, K. Tomuang, A. Limphirat, P. Hillmann, C. Herold, J. Steinheimer, Y. Yan, and M. Bleicher, Phys. Rev. C 99, 014901 (2019), eprint 1805.11509.
- Butler and Pearson (1963) S. T. Butler and C. A. Pearson, Phys. Rev. 129, 836 (1963).
- Scheibl and Heinz (1999) R. Scheibl and U. W. Heinz, Phys. Rev. C 59, 1585 (1999), eprint nucl-th/9809092.
- Zhu et al. (2015) L. Zhu, C. M. Ko, and X. Yin, Phys. Rev. C 92, 064911 (2015), eprint 1510.03568.
- Sun et al. (2018) K.-J. Sun, L.-W. Chen, C. M. Ko, J. Pu, and Z. Xu, Phys. Lett. B 781, 499 (2018), eprint 1801.09382.
- Zhao et al. (2020) W. Zhao, C. Shen, C. M. Ko, Q. Liu, and H. Song, Phys. Rev. C 102, 044912 (2020), eprint 2009.06959.
- Kittiratpattana et al. (2022) A. Kittiratpattana, T. Reichert, J. Steinheimer, C. Herold, A. Limphirat, Y. Yan, and M. Bleicher, Phys. Rev. C 106, 044905 (2022), eprint 2206.01625.
- Aichelin (1991) J. Aichelin, Phys. Rept. 202, 233 (1991).
- Aichelin et al. (2020) J. Aichelin, E. Bratkovskaya, A. Le Fèvre, V. Kireyeu, V. Kolesnikov, Y. Leifels, V. Voronyuk, and G. Coci, Phys. Rev. C 101, 044905 (2020), eprint 1907.03860.
- Gläßel et al. (2022) S. Gläßel, V. Kireyeu, V. Voronyuk, J. Aichelin, C. Blume, E. Bratkovskaya, G. Coci, V. Kolesnikov, and M. Winn, Phys. Rev. C 105, 014908 (2022), eprint 2106.14839.
- Oliinychenko et al. (2021) D. Oliinychenko, C. Shen, and V. Koch, Phys. Rev. C 103, 034913 (2021), eprint 2009.01915.
- Staudenmaier et al. (2021) J. Staudenmaier, D. Oliinychenko, J. M. Torres-Rincon, and H. Elfner, Phys. Rev. C 104, 034908 (2021), eprint 2106.14287.
- Danielewicz and Bertsch (1991) P. Danielewicz and G. F. Bertsch, Nucl. Phys. A 533, 712 (1991).
- Kapusta (1980) J. I. Kapusta, Phys. Rev. C 21, 1301 (1980).
- Siemens and Kapusta (1979) P. J. Siemens and J. I. Kapusta, Phys. Rev. Lett. 43, 1486 (1979).
- Weil et al. (2016) J. Weil et al., Phys. Rev. C 94, 054905 (2016), eprint 1606.06642.
- Cassing (2002) W. Cassing, Nucl. Phys. A 700, 618 (2002), eprint nucl-th/0105069.
- Sun et al. (2021) K.-J. Sun, R. Wang, C. M. Ko, Y.-G. Ma, and C. Shen (2021), eprint 2106.12742.
- Aichelin et al. (1987) J. Aichelin, A. Rosenhauer, G. Peilert, H. Stoecker, and W. Greiner, Phys. Rev. Lett. 58, 1926 (1987).
- Aichelin et al. (1988) J. Aichelin, A. Bohnet, G. Peilert, H. Stoecker, W. Greiner, and A. Rosenhauer, Phys. Rev. C 37, 2451 (1988).
- Hartnack et al. (1998) C. Hartnack, R. K. Puri, J. Aichelin, J. Konopka, S. A. Bass, H. Stoecker, and W. Greiner, Eur. Phys. J. A 1, 151 (1998), eprint nucl-th/9811015.
- Cassing and Bratkovskaya (2008) W. Cassing and E. L. Bratkovskaya, Phys. Rev. C 78, 034919 (2008), eprint 0808.0022.
- Cassing (2009) W. Cassing, Eur. Phys. J. ST 168, 3 (2009), eprint 0808.0715.
- Cassing and Bratkovskaya (2009) W. Cassing and E. L. Bratkovskaya, Nucl. Phys. A 831, 215 (2009), eprint 0907.5331.
- Bratkovskaya et al. (2011) E. L. Bratkovskaya, W. Cassing, V. P. Konchakovski, and O. Linnyk, Nucl. Phys. A 856, 162 (2011), eprint 1101.5793.
- Linnyk et al. (2016) O. Linnyk, E. L. Bratkovskaya, and W. Cassing, Prog. Part. Nucl. Phys. 87, 50 (2016), eprint 1512.08126.
- Feldmeier (1990) H. Feldmeier, Nucl. Phys. A 515, 147 (1990).
- Reichert et al. (2022) T. Reichert, A. Elz, T. Song, G. Coci, M. Winn, E. Bratkovskaya, J. Aichelin, J. Steinheimer, and M. Bleicher, J. Phys. G 49, 055108 (2022), eprint 2111.07652.
- Sorge et al. (1989) H. Sorge, H. Stoecker, and W. Greiner, Annals Phys. 192, 266 (1989).
- Marty and Aichelin (2013) R. Marty and J. Aichelin, Phys. Rev. C 87, 034912 (2013), eprint 1210.3476.
- Kadanoff and Baym (1962) L. P. Kadanoff and G. Baym, Quantum Statistical mechanics (W. A. Benjamin, Inc., New York, 1962).
- Cassing and Juchem (2000a) W. Cassing and S. Juchem, Nucl. Phys. A 665, 377 (2000a), eprint nucl-th/9903070.
- Cassing and Juchem (2000b) W. Cassing and S. Juchem, Nucl. Phys. A 672, 417 (2000b), eprint nucl-th/9910052.
- Cassing (2021) W. Cassing, Lect. Notes Phys. 989, pp. (2021).
- Cassing and Bratkovskaya (1999) W. Cassing and E. L. Bratkovskaya, Phys. Rept. 308, 65 (1999).
- Nilsson-Almqvist and Stenlund (1987) B. Nilsson-Almqvist and E. Stenlund, Comput. Phys. Commun. 43, 387 (1987).
- Andersson et al. (1993) B. Andersson, G. Gustafson, and H. Pi, Z. Phys. C 57, 485 (1993).
- Sjostrand et al. (2006) T. Sjostrand, S. Mrenna, and P. Z. Skands, JHEP 05, 026 (2006), eprint hep-ph/0603175.
- Kireyeu et al. (2020) V. Kireyeu, I. Grishmanovskii, V. Kolesnikov, V. Voronyuk, and E. Bratkovskaya, Eur. Phys. J. A 56, 223 (2020), eprint 2006.14739.
- Cassing (2007a) W. Cassing, Nucl. Phys. A 791, 365 (2007a), eprint 0704.1410.
- Cassing (2007b) W. Cassing, Nucl. Phys. A 795, 70 (2007b), eprint 0707.3033.
- Peshier and Cassing (2005) A. Peshier and W. Cassing, Phys. Rev. Lett. 94, 172301 (2005), eprint hep-ph/0502138.
- Soloveva et al. (2020) O. Soloveva, P. Moreau, and E. Bratkovskaya, Phys. Rev. C 101, 045203 (2020), eprint 1911.08547.
- Aoki et al. (2009) Y. Aoki, S. Borsanyi, S. Durr, Z. Fodor, S. D. Katz, S. Krieg, and K. K. Szabo, JHEP 06, 088 (2009), eprint 0903.4155.
- Cheng et al. (2008) M. Cheng et al., Phys. Rev. D 77, 014511 (2008), eprint 0710.0354.
- Borsanyi et al. (2015) S. Borsanyi, S. Durr, Z. Fodor, C. Holbling, S. D. Katz, S. Krieg, D. Nogradi, K. K. Szabo, B. C. Toth, and N. Trombitas, Phys. Rev. D 92, 014505 (2015), eprint 1504.03676.
- Bleicher and Bratkovskaya (2022) M. Bleicher and E. Bratkovskaya, Prog. Part. Nucl. Phys. 122, 103920 (2022).
- Bratkovskaya and Cassing (2008) E. L. Bratkovskaya and W. Cassing, Nucl. Phys. A 807, 214 (2008), eprint 0712.0635.
- Cassing et al. (2003) W. Cassing, L. Tolos, E. L. Bratkovskaya, and A. Ramos, Nucl. Phys. A 727, 59 (2003), eprint nucl-th/0304006.
- Song et al. (2021) T. Song, L. Tolos, J. Wirth, J. Aichelin, and E. Bratkovskaya, Phys. Rev. C 103, 044901 (2021), eprint 2012.05589.
- Cassing et al. (2016) W. Cassing, A. Palmese, P. Moreau, and E. L. Bratkovskaya, Phys. Rev. C 93, 014902 (2016), eprint 1510.04120.
- Palmese et al. (2016) A. Palmese, W. Cassing, E. Seifert, T. Steinert, P. Moreau, and E. L. Bratkovskaya, Phys. Rev. C 94, 044912 (2016), eprint 1607.04073.
- Seifert and Cassing (2018) E. Seifert and W. Cassing, Phys. Rev. C 97, 024913 (2018), eprint 1710.00665.
- Kireyeu (2021) V. Kireyeu, Phys. Rev. C 103, 054905 (2021), eprint 2103.10542.
- Seifert (2018) E. Seifert, Ph.D. thesis, Giessen U. (2018).
- Kajantie and Karimaeki (1971) K. Kajantie and V. Karimaeki, Comput. Phys. Commun. 2, 207 (1971).
- Byckling and Kajantie (1971) E. Byckling and K. Kajantie, Particle Kinematics: (Chapters I-VI, X) (University of Jyvaskyla, Jyvaskyla, Finland, 1971).
- Pan and Pratt (2014) Y. Pan and S. Pratt, Phys. Rev. C 89, 044911 (2014).
- Neidig et al. (2022) T. Neidig, K. Gallmeister, C. Greiner, M. Bleicher, and V. Vovchenko, Phys. Lett. B 827, 136891 (2022), eprint 2108.13151.
- Cannoni (2014) M. Cannoni, Phys. Rev. D 89, 103533 (2014), eprint 1311.4494.
- Lang et al. (1993) A. Lang, H. Babovsky, W. Cassing, U. Mosel, H.-G. Reusch, and K. Weber, Journal of Computational Physics 106, 391 (1993), ISSN 0021-9991.
- Xu and Greiner (2005) Z. Xu and C. Greiner, Phys. Rev. C 71, 064901 (2005), eprint hep-ph/0406278.
- Kodama et al. (1984) T. Kodama, S. B. Duarte, K. C. Chung, R. Donangelo, and R. A. M. S. Nazareth, Phys. Rev. C 29, 2146 (1984).
- Hirano et al. (2013) T. Hirano, P. Huovinen, K. Murase, and Y. Nara, Prog. Part. Nucl. Phys. 70, 108 (2013), eprint 1204.5814.
- Röpke et al. (1982) G. Röpke, L. Münchow, and H. Schulz, Nucl. Phys. A 379, 536 (1982).
- Röpke and Schulz (1988) G. Röpke and H. Schulz, Nucl. Phys. A 477, 472 (1988).
- McGee (1967) I. J. McGee, Phys. Rev. 158, 1500 (1967).
- Lacombe et al. (1980) M. Lacombe, B. Loiseau, J. M. Richard, R. Vinh Mau, J. Cote, P. Pires, and R. De Tourreil, Phys. Rev. C 21, 861 (1980).
- Lacombe et al. (1981) M. Lacombe, B. Loiseau, R. Vinh Mau, J. Cote, P. Pires, and R. de Tourreil, Phys. Lett. B 101, 139 (1981).
- Le Fèvre et al. (2019) A. Le Fèvre, J. Aichelin, C. Hartnack, and Y. Leifels, Phys. Rev. C 100, 034904 (2019), eprint 1906.06162.
- Kireyeu et al. (2022) V. Kireyeu, J. Steinheimer, J. Aichelin, M. Bleicher, and E. Bratkovskaya, Phys. Rev. C 105, 044909 (2022), eprint 2201.13374.
- Norem (1971) J. H. Norem, Nucl. Phys. B 33, 512 (1971).
- Oh et al. (2009) Y. Oh, Z.-W. Lin, and C. M. Ko, Phys. Rev. C 80, 064902 (2009), eprint 0910.1977.
- VerWest and Arndt (1982) B. J. VerWest and R. A. Arndt, Phys. Rev. C 25, 1979 (1982).
- Arndt et al. (1997) R. A. Arndt, C. H. Oh, I. I. Strakovsky, R. L. Workman, and F. Dohrmann, Phys. Rev. C 56, 3005 (1997), eprint nucl-th/9706003.
- Oh et al. (1997) C.-H. Oh, R. A. Arndt, I. I. Strakovsky, and R. L. Workman, Phys. Rev. C 56, 635 (1997), eprint nucl-th/9702006.
- Anderson et al. (1974) H. L. Anderson, D. A. Larson, L. C. Myrianthopoulos, L. Dubal, C. K. Hargrove, E. P. Hincks, R. J. Mckee, H. Mes, D. Kessler, and A. C. Thompson, Phys. Rev. D 9, 580 (1974).
- Shimizu et al. (1982) F. Shimizu, Y. Kubota, H. Koiso, F. Sai, S. Sakamoto, and S. S. Yamamoto, Nucl. Phys. A 386, 571 (1982).
- Tanabashi et al. (2018) M. Tanabashi et al. (Particle Data Group), Phys. Rev. D 98, 030001 (2018).
- Ikeno et al. (2021) N. Ikeno, R. Molina, and E. Oset, Phys. Rev. C 104, 014614 (2021), eprint 2103.01712.
- Bugg et al. (1966) D. V. Bugg, D. C. Salter, G. H. Stafford, R. F. George, K. F. Riley, and R. J. Tapper, Phys. Rev. 146, 980 (1966).
- Cugnon et al. (1996) J. Cugnon, J. Vandermeulen, and D. L’Hote, Nucl. Instrum. Meth. B 111, 215 (1996).
- Sakurai and Napolitano (2020) J. J. Sakurai and J. Napolitano, Modern Quantum Mechanics, Quantum physics, quantum information and quantum computation (Cambridge University Press, 2020), ISBN 978-0-8053-8291-4, 978-1-108-52742-2, 978-1-108-58728-0.