Evidence of isospin-symmetry violation in high-energy collisions of atomic nuclei \PreprintIdNumberCERN-EP-2023-283 \ShineAbstract Strong interactions preserve an approximate isospin symmetry between up () and down () quarks, part of the more general flavor symmetry. In the case of meson production, if this isospin symmetry were exact, it would result in equal numbers of charged ( and ) and neutral ( and ) mesons produced in collisions of isospin-symmetric atomic nuclei. Here, we report results on the relative abundance of charged over neutral meson production in argon and scandium nuclei collisions at a center-of-mass energy of 11.9 GeV per nucleon pair. We find that the production of and mesons at mid-rapidity is higher than that of the neutral mesons. Although with large uncertainties, earlier data on nucleus-nucleus collisions in the collision center-of-mass energy range GeV are consistent with the present result. Using well-established models for hadron production, we demonstrate that known isospin-symmetry breaking effects and the initial nuclei containing more neutrons than protons lead only to a small (few percent) deviation of the charged-to-neutral kaon ratio from unity at high energies. Thus, they cannot explain the measurements. The significance of the flavor-symmetry violation beyond the known effects is 4.7 when the compilation of world data with uncertainties quoted by the experiments is used. New systematic, high-precision measurements and theoretical efforts are needed to establish the origin of the observed large isospin-symmetry breaking.
1 Introduction
One of the main aims of basic research is to understand the fundamental constituents of matter and the interactions between them. Within Quantum Chromodynamics (QCD) [1], the theory of strong interactions, the fundamental particles are quarks and gluons carrying color – the charge of strong interactions. Because of confinement, quarks and gluons are hidden in colorless hadrons, particularly protons and neutrons. The strong force binds them, forming atomic nuclei.
Accelerator-based experiments recording collisions of highly energetic hadrons and nuclei allow for systematic studies of the properties of strong interactions. In these collisions, many new particles are produced. They are predominantly mesons containing one valence quark () and one valence anti-quark (). The most copiously produced are the lightest mesons, pions and kaons, built from up (), down () and strange () quarks and the corresponding anti-quarks.
QCD assumes that interactions are independent of quark type (flavor) for equal quark masses and in the absence of other interactions, a feature known as flavor symmetry. When only the light quarks up and down are considered, flavor symmetry reduces to isospin symmetry, historically introduced in the pre-QCD period by Heisenberg to understand the properties of nuclei [2]. The masses of up and down quarks, MeV and MeV [3], are not equal, but they are much smaller than the QCD scale, [4, 5]. (Note that units in this paper follow the Particle Data Group (PDG) [3] convention: masses and energies are expressed in MeV (or GeV), whereas momenta in (or ). The relative differences are given as the ratio of the difference to the mean.) Hence isospin-symmetry breaking effects are small, as confirmed by the mass ratios of pions and kaons, and . Moreover, the elastic cross sections for pion-pion, pion-nucleon, and nucleon-nucleon scattering closely follow the predictions of isospin symmetry [6, 7]. Here, of special interest is a specific isospin transformation, an inversion of the third component of the isospin, called the charge transformation for historical reasons. It is equivalent to swapping quarks. At the hadronic level, the charge transformation implies swapping , , , , etc.
Let us consider nucleus-nucleus (+) collisions and, for simplicity, assume that both nuclei have an equal number of protons and neutrons. Without referring to a detailed mathematical formalism, charge symmetry means that strong interactions are invariant under the charge transformation of every nucleus and hadron of the initial and final states. For an ensemble of initial states being invariant under the charge transformation, the probabilities of having initial states related by this transformation are equal. This is indeed the case of nucleus-nucleus collisions, for which each nucleus has an equal number of protons and neutrons. Then, the invariance under charge transformation also holds for the final state ensemble, implying that the mean multiplicities of charge-transformation related hadrons, such as and as well as and , coincide:
(1) |
Note that since models predict only properties of ensembles of events and not outcomes of single events, we need to consider quantities averaged over the event ensembles. The subject has a vast literature; see, for example Refs. [8, 9, 10, 11, 12, 13]. Consequently, the exact isospin symmetry prediction for the charged-to-neutral kaon ratio in nucleus-nucleus collisions with electric charge to baryon number reads
(2) |
(Note that the and states are produced in strong interactions, but they decay through weak interactions. Consequently, in the final state, one observes linear combinations of the latter known as short () and long (), where short and long refer to their weak decay lifetime [3]. By neglecting the small CP violation, the multiplicities corresponding to weak and strong eigenstates are related by .) The prediction given by Eq. (2) is a reference for experimental testing of the isospin symmetry in hadron production processes. For a more detailed introduction and didactic derivations see Ref. [14].
Here, we report a measurement of the ratio in the 10% most central collisions of argon () and scandium () nuclei at center-of-mass energy per nucleon pair equal to 11.9 GeV. Further on, we compare the NA61/SHINE result with the world data on charged and neutral kaon production in nucleus-nucleus collisions. The results indicate a significant excess of charged over neutral kaon production. This excess cannot be explained by known effects violating the isospin symmetry. This is discussed and demonstrated by comparing experimental results to well-known theoretical approaches, the statistical Hadron Resonance Gas (HRG) [15] and the dynamical Ultrarelativistic Quantum Molecular Dynamics (UrQMD) [16] models. The predictions of models are calculated for reactions corresponding to experimental data, generally with . They consider isospin-breaking effects in strong interactions and, importantly, the production and subsequent decays of the mesons.
Summarizing, the NA61/SHINE Collaboration measures a charged-to-neutral kaon ratio in Ar+Sc collisions at 11.9 GeV per nucleon pair. This value aligns with previous experimental measurements, albeit their uncertainties are larger. The significance of the isospin symmetry violation beyond the known effects amounts to 4.7 when all measurements are considered, and uncertainties quoted by the experiments are used. This is the first evidence of an unexplained isospin symmetry violation in hadron production processes.
2 Results
2.1 Production of mesons in central Ar+Sc collisions at the CERN SPS
The new experimental results presented here have been obtained by the NA61/SHINE fixed-target experiment at the CERN Super Proton Synchrotron [17]. The measurements of and production in the 10% most central Ar+Sc reactions at GeV have been published elsewhere [18]. The analysis procedure and details of systematic uncertainties are given in Refs. [19, 20]. Here, we present the first measurement of production in nucleus-nucleus collisions from NA61/SHINE. Earlier data from this experiment, on production in +C, +C, +C, +Be, and + collisions can be found in Refs. [21, 22, 23, 24, 25, 26]. For more details concerning the experimental procedure, see Methods’ "Experimental procedure" subsection.
The comparison of the rapidity distribution of mesons to the average of rapidity distributions for and mesons is presented in Fig. 1. The rapidity is a relativistic generalization of the particle velocity along the direction of the incoming nuclei. We calculate rapidity in the nucleon-nucleon collision center-of-mass system, and positive corresponds to the direction of the Ar nucleus.
In the entire range of rapidity covered by the measurement, the averaged charged mesons yield prevails significantly over the neutral mesons one. To quantify this effect, Table 1 presents the rapidity densities of , and production measured at mid-rapidity (). Here, the relative excess of charged mesons is %. Integration of the two distributions in Fig. 1 over positive rapidity, , gives and for the production rates per collision of and , respectively (total uncertainties are given; the quantities provided for charged mesons are based on Ref. [18]). The resulting difference of corresponds to a surplus of charged ( and ) over neutral ( and ) states equal to at positive rapidity. Under the assumption that the charged-to-neutral ratio would be similar also at negative rapidity, the total excess would amount to additional or mesons per one central Ar+Sc collision.

statistical | systematic | total | ||||
3.732 | 0.016 | 0.148 | 0.149 | |||
2.029 | 0.012 | 0.069 | 0.070 | |||
2.433 | 0.027 | 0.102 | 0.106 | |||
charged-to-neutral meson ratio: | ||||||
1.184 | 0.014 | 0.060 | 0.061 |
A comparison of distributions of with averaged and mesons as a function of transverse momentum (the momentum component perpendicular to the direction of the incoming nuclei) is shown in Fig. 2. Both distributions are integrated over the rapidity range . The prevalence of charged over neutral mesons is again evident. The insert in the figure shows the -dependence of the ratio . The corresponding excess of mesons containing , over those containing , quarks and anti-quarks remains in the range 6–33% over the considered range of .

2.2 Comparison to the world data and models
Figure 3 compares the present measurement of the ratio at mid-rapidity and the world data compiled by us and detailed in Methods’ "World data" subsection. The experimental results were obtained by CERES [27, 28, 29], STAR BES [30, 31], STAR [32, 33, 34, 35], ALICE [36, 37], NA35 [38, 39], NA49 [40, 41, 42], and HADES [43, 44] experiments. We note that the compilation includes measurements at mid-rapidity and total multiplicities. This may increase the overall spread between the data points. We also note the sizeable uncertainties of the earlier measurements. These probably explain why the aforementioned charged-over-neutral anomaly was never reported as an experimental observation. Despite these uncertainties, a consistent picture emerges in the energy range GeV. The ratio is above one for all experiments except NA35 and ALICE.


Figure 3 also compares the data with the HRG and UrQMD model predictions. The HRG calculations were performed with (solid line) and values corresponding to collisions studied in the experiments (black dots). The UrQMD results were obtained for central Au+Au collisions. See Methods’ "Models" subsection for details on the models. The predictions of models agree well with each other but are systematically lower than the experimental data. At energies larger than 10 GeV, the mass difference between charged and neutral kaons, leading to isospin-symmetry breaking (mostly via -meson decays), increases the ratio by about 0.03. Other isospin-breaking effects can be neglected. The ratio is reduced for collisions with , but this is insignificant for energies larger than 10 GeV. At energies lower than 10 GeV, the ratio is significantly more sensitive to the known isospin-breaking effects as well as to the ratio; see Methods’ "Models" subsection for more details.
To quantify the measured isospin-symmetry breaking beyond the known effects, the ratio of the measured to the corresponding HRG baseline is shown in Fig. 4. We do not consider the lowest energy data point (HADES) because the known isospin-breaking and effects are significant at the low collision energies; thus, predictions may be model-dependent. We also do not consider the NA35 point because, unlike other measurements, charged kaons were identified by reconstructing their decays, leading to large statistical uncertainties and possible biases. Thus, the number of selected measurements at different collision energies (from SPS to LHC) is . Out of them, only one is below unity. No significant dependence of the double ratio on collision energy and nuclear mass number of colliding nuclei is visible.
The weighted average of all double ratios shown in Fig. 4 is , where the uncertainty was calculated using kaon uncertainties reported by experiments. The HRG uncertainties are small and were neglected. The significance of the isospin violation is . The may indicate either a correlation between results or an overestimation of the uncertainties.
3 Discussion
In the following, we discuss possible effects that may potentially contribute to the violation of isospin symmetry in kaon production.
First, we consider symmetry-breaking effects due to the non-equal bare and masses in strong interactions. They are included in the HRG and UrQMD models. Then, we discuss the possible influence related to electromagnetic and weak processes.
-
(A)
Mass effects within strong interactions. Within QCD, the isospin symmetry is not exact because and quark masses are different, and MeV, respectively. The different quark masses lead to different masses of hadrons within the isospin multiplets, particularly different masses of charged and neutral kaons. This modifies the ratio . We list below the considered effects and quantify their influence on using the statistical Hadron Resonance Gas model (HRG) [15]. The results are cross-checked with the microscopic transport model, UrQMD [45, 16, 46]. For details, see Methods’ "Models" subsection.
-
(i)
Smaller masses of charged kaons than neutral ones, MeV and MeV, lead to an increase of resulting from direct kaon production by about 0.02. This was estimated by removing resonances from the particle list of HRG. We have numerically verified that HRG for and with exact isospin symmetry gives , as expected.
-
(ii)
A significant fraction of kaons (up to several dozen percent) results from resonance decays [47]. Different kaon masses affect the branching ratios of resonances. The most striking example is meson, which decays about more frequently into charged kaons than neutral ones. This large difference is because the mass is just above the kaon-kaon thresholds. Including the kaon production from resonance decays increases by about 0.03.
-
(iii)
In connection to the previous point, other potentially relevant decays refer to the resonances and , whose masses are close to the decay threshold. In the PDG review [3], for both resonances decays are reported as seen. The HRG model used in this work initially assumes equal branching ratios () of and decay channels. Yet, just as for the meson, isospin breaking may be relevant. (Note that isospin breaking for and is experimentally confirmed by the - mixing [48, 49] leading to the otherwise forbidden decays and ; see the experimental results for these small but non-zero transitions in Ref. [50].) Using the estimate (see Methods’ "Models" subsection) / one gets an increase of by about . The yield of charged kaons may also be affected by other effects, such as the electromagnetic interaction between and (see also (B) (ii) below). To calculate the upper limit due to this effect, we assumed (no decays to neutral kaons) for both and . This leads to the increase of by at most at the highest collision energy.
-
(iv)
Mass differences of hadrons from other isospin multiplets also break the flavor symmetry and affect . The largest effect comes from the mass difference between proton and neutron, which reduces at the lowest collision energies. At energies larger than 10 GeV, this effect is negligible.
-
(i)
-
(B)
Electromagnetic processes. Electromagnetic interactions do not obey isospin symmetry because electric charges differ for the quark flavors and . The electromagnetic interaction slightly affects the masses of hadrons. For instance, the neutral pion is lighter than the charged ones. The HRG and UrQMD include such effects since the physical masses are used. However, the effects mentioned below are not included in the models.
-
(i)
Electromagnetic decays of hadrons are typically suppressed by a factor compared to strong ones. Consequently, decays that involve the production of virtual photons and their subsequent decay into charged kaons are suppressed by a factor and thus negligible. Taking into account the charge of the nuclei and , one would expect an effect of the type , which is not observed in the experimental data – isospin-symmetry breaking for collisions of light and heavy nuclei is similar, see Fig. 4.
-
(ii)
There may also be non-perturbative electromagnetic effects at the hadronic level that might affect the kaon multiplicities. One notable example is the case of pairs with low momenta of kaons , see e.g. Ref. [51]. This is, in particular, the case for the decays of the resonances and , whose masses are close to the two-kaon decay threshold. It is then possible that this effect will increase the number of produced pairs. Experimental search for interaction close to the threshold was conducted at COSY in the study of the reaction , see e.g. Ref. [52] and the review [53]. An increase in the cross-section close to the threshold has been measured, where different effects appear to be relevant, with an important role played by the Coulomb attraction. We have quantified the impact of and decays in point (A) (iii) above and in Methods’ "Models" subsection, showing that they cannot explain the measured value of .
-
(iii)
The and pair creation in strong processes may be affected by electromagnetic interactions. They are different for and pairs due to different electric charges of up and down quarks. This leads to a different phase space for their production, favoring pairs and thus charged kaons. In particular, the quark-gluon effective coupling is enhanced by QED effects due to the attraction among quarks, leading to a larger coupling of gluons to -quarks than -quarks [54]. A model of the space-time evolution of the pair creation will be needed to quantify the effect. In addition, the isospin breaking due to the Coulomb potential of highly-charged fireballs formed in heavy-ion collisions is discussed in Ref. [55] within the statistical QGP model. We recall that electromagnetic interactions are expected to modify fusion rates in the Big Bang nucleosynthesis epoch; see, for example, Refs. [56, 57].
-
(i)
-
(C)
Uncertainties in weak decays. The weak interaction does not obey the isospin symmetry. The mean lifetimes [3] of charged and neutral kaons are s and s, s. The charged kaons are typically measured by reconstructing their trajectories in a detector. Due to the large mean lifetime, the corrections for their losses caused by weak decays are small. In contrast, the neutral kaons are measured by reconstructing their decays into two charged pions. Typically, the corrections for the losses caused by weak decays are large. This is because the decay should be far enough from the interaction point to separate the decay point from the background in high-multiplicity + collisions. Assuming that the meson is measurable when the lifetime of a particle in its rest frame is larger than the mean lifetime (typical for NA61/SHINE), one estimates the maximum relative bias of the mean multiplicity, , as:
(3) where is the uncertainty of the mean lifetime. Thus, the maximum deviation of from unity due to the uncertainty of the mean lifetime is 0.13%.
Finally, we discuss the consequences of having collisions with , which corresponds to many experimental results presented in Fig. 3. The third component of isospin equals and therefore the total isospin is limited as . The compiled experimental results in nucleus-nucleus collisions correspond to the ranging from about 0.4 (Pb+Pb and Au+Au collisions) to 0.5 (S+S collisions); see Fig. 3. These limits correspond to the normalized per baryon third component and total isospin , and , respectively. The non-zero and for heavier nuclei can affect the charged-to-neutral kaon ratio in two ways:
-
(a)
The larger fraction of neutrons than protons for heavy nuclei having enhances neutral kaon production compared to charged ones and thus reduces . This fact is taken into account by the employed theoretical models that use the physical value for . The reduction of is significant at low collision energies and is small compared to other effects at energies larger than 10 GeV; see Fig. 3 and Methods’ "Models" subsection.
-
(b)
For , the total isospin is limited by . Generally, nuclei in the ground state have the lowest possible value of the total isospin [58]. This rule extends to a state of two identical nuclei in the ground state, which, for the considered case, implies . Thus, a possible dependence of on and reduces to the dependence on . The latter is discussed above in point (a). The experimental data also allow a rough estimate of the influence of the small but non-zero value of the normalized isospin on the charged-to-neutral kaon ratio. Results for heavy nuclei, 208Pb+208Pb and 197Au+197Au ( for ), are similar to those for intermediate nuclei, 40Ar+45Sc (), and 32S+32S (), see Fig. 3. This suggests that the sensitivity of to the close to zero () is low. Note, however, that there are large uncertainties in the experimental results.
-
(c)
In the case of central collisions of heavy nuclei, the charge-to-baryon ratio of the interacting nucleon system (participant nucleons) can be higher than the total proton-to-nucleon ratio in colliding nuclei. This is because protons tend to be distributed closer to the center of a nucleus [59]. However, this effect is expected to be small because the ratio seems independent of the colliding nuclei size. Low-mass nuclei can be described as alpha clusters [60] (clusters of two protons and two neutrons). Thus, if significant, the effect should disappear for collisions of low-mass nuclei, particularly at the lowest collision energy ( GeV), where is sensitive to the small changes of the electric charge to baryon number ratio, . The data do not support this.
Closing comments on future perspectives of experimental and theoretical efforts are in order here.
-
(I)
Concerning measurements, reviewing the validity of the past results and confirming them with new high-precision data is important. Systematic results on the collision energy and nuclear mass dependence of the isospin-breaking effect should help us understand its nature. Measurements of the charged-to-neutral kaon ratio in collisions of an equal number of protons and neutrons would reduce uncertainty in its interpretation. The NA61/SHINE experiment plans to perform such measurements for O+O and Mg+Mg collisions [61]. If needed, the measurements of deuteron-deuteron interactions may be possible in the long future. They will require the production of primary deuteron beams in the CERN accelerator complex and the construction of a liquid deuterium target for NA61/SHINE.
-
(II)
In connection with the previous point, an interesting experimental test is also possible by considering +C and +C interactions [62]. While the ensemble of only one of them is not invariant under charge transformation, the ensemble having an equal number of +C and +C interactions is invariant. Thus, for the joint ensemble, the exact charge symmetry predicts . NA61/SHINE recorded data on +C and +C interactions at 158 in October 2024 for the test.
-
(III)
During the current CERN accelerator complex operation period (Run 3), NA61/SHINE records high-statistics data on inelastic Pb+Pb collisions at 150. This will allow precision measurements of the ratio as a function of collision centrality. The larger number of neutrons than protons in Pb nuclei may reduce the ratio . This should be taken into account when interpreting the results.
-
(IV)
The relationship between the production rates of charged and neutral kaons in hadronic collisions was used to estimate the flux of secondary neutral kaons produced in +Be collisions at beam momenta of several 100 in neutral kaon and neutrino beams [63]. Recently, the charged-to-neutral kaon ratio was studied in + interactions at SPS energies, and no excess in comparison to a simple quark counting model emerged [64].
-
(V)
Kaons play a special role due to their simple isospin structure and easy measurement. This explains why the first results on a large isospin-symmetry breaking in multi-particle production are reported for kaons. Yet, it is important to perform a similar study for other isospin multiplets in the future. For example, using the same methods, we have checked that the (anti-)proton to (anti-)neutron ratio is even less sensitive to the known isospin-symmetry breaking effects and, thus, is predicted to be almost exactly one in nucleus-nucleus collisions with in a broad range of collision energies including the low energies.
-
(VI)
One can extend the current models by introducing new isospin-breaking processes and fitting their parameters to the data. This can be done either for the quark-gluon processes or the hadron-resonance processes. For example, within the statistical hadronization models, one can introduce the quark fugacity factors for and quarks separately [65, 66, 67]. This could allow us to make predictions for other hadron ratios but will not explain the origin of the violation.
-
(VII)
The possibility of having a phase of strongly interacting matter with a significant isospin violation was suggested by Pisarski and Wilczek within a linear model of QCD [68]. They expect masses of and mesons to decrease if the symmetry is effectively restored at temperatures lower than the one of the chiral phase transition. The chiral anomaly [69, 70] could break isospin (especially, it affects the pion isotriplet) but does not affect the charged-to-neutral kaon ratio.
-
(VIII)
Creating Disoriented-Chiral-Condensate (DCC) domains in heavy-ion collisions has been considered for many years [71, 72, 73]. They may be signaled by large fluctuations of the charged-to-neutral pion [74] and kaon ratios [75, 76]. A puzzling result on kaon fluctuations was recently reported by ALICE at LHC [77]. Its possible interpretation by the DCC or disoriented-isospin-condensates formation is discussed in Refs. [78, 79]. The considered models for the charged-symmetric ensemble of collisions predict [80]. The inclusion of isospin-breaking effects in the extended Linear Sigma model [81] was recently discussed in Ref. [82], where through a fit to available experimental masses and decays of light mesons, it is shown that the relative difference between the and chiral condensates amounts to , implying that only very small deviations from are expected from this effect.
Thus, the presented results on the charged-to-neutral kaon ratio are the first evidence of an unexplained isospin symmetry violation in hadron production processes. Further studies are needed to understand the underlying physics, particularly reducing the experimental uncertainties and quantifying the role of electromagnetic effects. If these steps do not solve the issue, more speculative explanations shall be investigated.
Methods
A. Experimental procedure
Experimental setup. The SPS Heavy Ion and Neutrino Experiment (SHINE) is a fixed-target detector operating at the CERN Super Proton Synchrotron (SPS). It is a multi-purpose spectrometer optimized to study hadron production in various collisions (hadron-proton, hadron-nucleus, and nucleus-nucleus). The detection setup used for the measurements reported here is described below. Its details and a description of the detector performance can be found in Ref. [17].
The beamline is equipped with an array of beam detectors upstream and downstream of the target, used to identify and measure the trajectory of the beam particles and trigger the spectrometer data acquisition. The tracking devices of the NA61/SHINE spectrometer are Time Projection Chambers (TPCs). Two Vertex TPCs are placed inside a magnetic field. Two large-volume Main TPCs measure the charged particle trajectories downstream of the 4.5 Tm magnetic field. The latter provides the bending power for a precise determination of particle momenta. The information about the energy losses () of the charged particles in the TPCs, together with Time-of-Flight (ToF) measurements, allows for particle identification in a wide momentum range. The most downstream detector on the beamline is the Projectile Spectator Detector (PSD). It measures the energy of the spectator remnant of the projectile nucleus, closely related to the collision centrality in nucleus-nucleus reactions.
Physics objects. This article compares the production of charged and neutral mesons in Ar+Sc collisions at a center-of-mass energy per nucleon pair of 11.9 GeV. The beam had a momentum of 75. The stationary target consisted of six plates, with a total thickness of 6 mm.
A detailed account on the extraction of charged (, ) yields can be found in Ref. [18]. Only the neutral mesons are considered in the present analysis. They can be detected via their weak decay into two charged pions . The mean lifetime for this decay is 2.7 cm. A detailed presentation of the analysis procedure and systematic uncertainties can be found in Ref. [83].
Analysis. Before analyzing mesons, the recorded Ar+Sc collision data undergo event and track selection procedures. Event selection uses information from the beam detectors to ensure the quality of the measured beam trajectory. It rejects events with more than one beam-target interaction during the trigger-time window. It also reduces the background from off-target interactions based on information about the quality of the main interaction vertex. Finally, it selects the 10% most central collisions using the information from the PSD. This is realized by selecting the 10% lowest energy deposits from the spectator remnant of the Ar nucleus. The total number of recorded collisions (events) was , from which (37%) remained after all cuts. For more details, especially the centrality selection, see Ref. [18].
The next step is reconstructing the charged particle tracks in the TPCs. Pattern recognition algorithms combine space points recorded in the TPCs into tracks. Their curvature and the magnetic field are used to compute the momenta of the corresponding particles. The minimum number of reconstructed space points in the VTPCs must be more than 10, and the computed momenta must be larger than 400 (in the laboratory frame). The latter selection excludes a large fraction of low-momentum electrons from the analysis. The known positions of the target and the most probable intersection point of measured tracks define the position of the primary vertex.
reconstruction. Unlike the charged particles, the neutral mesons do not leave a measurable track in the detectors. They are measured by reconstructing their oppositely charged decay products (daughter particles). The two-body decays of create characteristic -shaped particle pairs originating at the decay vertex. This topology is called . It is searched with a dedicated -finder algorithm that looks for track pairs of particles with opposite charges. These track pairs are extrapolated backwards until their mutual distance of the closest approach is reached. If this distance is smaller than a given limit value, the track pair becomes a candidate with its origin at the decay vertex.
Two further cuts are placed on the track pairs. The first cut imposes a minimum value on the angle between the direction of the line joining the primary and decay vertices and the direction given by the vector sum of the momenta of the decay daughters. The second condition requires a minimum distance between the primary and decay vertex (a minimum length of the decaying particle). The corresponding cut requirements depend on rapidity and are listed in Methods’ "Extended data" subsection. Starting with the approximate decay point, a -fitter program optimizes the decay point position and the momenta of the decay daughters. Assuming that the daughter particles are pions, it is straightforward to reconstruct the invariant mass of the decaying particle (the invariant mass is defined as , where are the energies of the decay products, are their momenta, and is assumed).
The invariant mass distribution of candidates is populated by and decays, photon conversion in some detector material, and spurious particle crossings. A signal will appear as a peak on a slowly varying background. For a double-differential analysis, the momentum space was divided into seven rapidity bins ranging from to 2 and nine transverse-momentum bins ranging from 0 to 2.7 . The raw number of in a given kinematic bin is obtained from fits of appropriate signal and background functions to the invariant mass distribution of the corresponding candidates. The fitted signal function is taken as a Lorentzian, and the background function is a third-order Chebychev polynomial. The integral of the signal function divided by the bin width is equal to the raw (uncorrected) number of the reconstructed in a given kinematic bin. Two typical invariant mass distributions with signal and background fits are shown in Methods’ "Extended data" subsection Fig. 5.
Corrections. To correct the results for losses due to detection and data processing inefficiencies, detailed Monte Carlo simulations were performed. These simulations comprised Ar+Sc collisions generated by the EPOS model [84], and particles propagated in the NA61/SHINE detector using the GEANT framework [85]. The charged particle tracks were reconstructed and analyzed using the same software as used for the experimental data. The branching ratio of decays was considered in the GEANT framework. The final output of the simulation consisted of reconstructed multiplicities. The ratio of the simulated and reconstructed numbers of was used as a correction factor in each – bin.
Systematic uncertainties of the measured data points were estimated by comparing the results of the entire analysis (including Monte Carlo simulations and corrections) obtained with varying cut values. The reliability of the reconstruction and fitting procedures can be scrutinized by studying the lifetime. Methods’ "Extended data" subsection Fig. 6 shows the computed mean lifetime of in seven rapidity bins. Good agreement with the average value provided by the PDG [3] is observed.
Transverse momentum distributions. The distributions shown in Methods’ "Extended data" subsection Fig. 7 represent the final results of the analysis. The yields are shown as a function of transverse momentum in seven bins of rapidity. The data points are fitted with the function:
(4) |
in which is a normalisation factor, is the inverse slope parameter, and is the mass taken from Ref. [3]. The formula assumes for simplicity. The fit functions are plotted as red curves, and the inverse slope parameters obtained from the fits are reported in the figure legends.
The transverse momentum distributions of charged and neutral mesons drawn in Fig. 2 (of the main text) are also fitted with the function defined by Eq. (4). The bottom panel of the figure presents the ratio of the two fitted curves, with its uncertainty band obtained by the propagation of the uncertainties of the fitted parameters.
Rapidity distribution. The final yields in each bin of rapidity were obtained as the integrals of the curves fitted to the respective transverse momentum spectra, Eq. (4), including extrapolations to unmeasured regions. A comparison to the alternative method of replacing integrals in the measured regions by sums of data points only brought a negligible contribution to the systematic uncertainty. Methods’ "Extended data" subsection Fig. 7 showed that extrapolations were needed only in the first and last rapidity bin. They amount to 88% and 6.2%, respectively. The large extrapolation in the first bin of rapidity increases the total uncertainty of the corresponding data point shown in Fig. 1 (of the main text). In this figure, the obtained rapidity distribution of the has been fitted with a function consisting of two Gaussians with centers displaced by a value of with respect to . These Gaussians have the same widths but may have different amplitudes. The resulting small asymmetry of the fitted rapidity distribution originates from a combined effect of the mass asymmetry of the colliding target and projectile nuclei ( and ) and the selection of central collisions by the energy measured in the kinematic region of the projectile spectator remnants. The former favors backward and the latter forward rapidities. The yields of charged mesons at mid-rapidity listed in Table 1 (of the main text) were taken from Ref. [18]. They were determined in the interval as discussed therein. The yield of neutral mesons at mid-rapidity was determined at from the aforementioned fit. Its systematic uncertainty was estimated the same way as for the data points (see above), and its statistical uncertainty was obtained by propagation of the statistical uncertainties of the fit. Both statistical and systematic uncertainties of charged and neutral yields were propagated into the ratio . The additional uncertainty of resulting from the difference in the mid-rapidity definition for charged and neutral mesons was estimated to be 0.5%, about 10% of the total systematic uncertainty.
B. World data
This section presents the yields of charged and neutral kaons measured by various experiments across different collision systems and energies and within specified centrality and rapidity regions. The results, presented in Table LABEL:expConditions, are sourced directly from the original experimental publications without any modifications to ensure consistency of the quantities reported. The exceptions are HADES and yields, where two sources of systematic uncertainties were reported [43]. In our analysis, they were added in quadrature, and the square root of such a sum is shown in Table LABEL:expConditions as the final systematic uncertainty (although in further calculations we used more precise values than 0.0014 and 0.000032 displayed in the table). In the NA49 experiment, the yield in Pb+Pb at 7.6 GeV [40] was reported with asymmetric systematic uncertainty; in this case, the upper limit was taken as . For and yields in NA35 S+S collisions at 19.4 GeV only statistical uncertainties were reported in the form of numerical values [38]. We took the NA35 estimate of systematic uncertainty as 3% [38], and the resulting numerical values are presented in the table. Finally, for the STAR experiment at 130 GeV [32], two types of uncertainties were reported: uncorrelated errors (first) and correlated systematic errors (second); see Ref. [32] for details.
For all kaon yields reported with statistical and systematic uncertainties separately, we calculated the total uncertainties as (for STAR at 130 GeV was taken as ). Their rounded (to two significant digits) values are displayed in the third column of Table LABEL:expConditions however, more precise values were used when propagating them to of presented in Table LABEL:Rkvalues.
The notation "Yield (4)" refers to particle mean multiplicity in full phase space. The "Yield ()" corresponds to mid-rapidity production, in most cases expressed as rapidity density measured in the region specified in Table LABEL:expConditions as " range" (for CERES results and NA61/SHINE mesons the fits at mid-rapidity were used). In some cases, different intervals were used for charged and neutral kaons. When calculating the charged-to-neutral kaon ratio, we used the originally published results.
The HADES data for Au+Au collisions at 2.4 GeV [86, 87], the FOPI data for Al+Al collisions at 2.7 GeV [88, 89], and the NA49 data for Pb+Pb collisions at 17.3 GeV [41, 42] are excluded from this paper, as the charged and neutral kaons were measured in significantly different centrality intervals. Normalizing these results by the number of participants would introduce model dependence of the ratio. Moreover, we also omit kaon yields evaluated by the Authors of Ref. [90] based on rapidity spectra measured by AGS experiments in Si+Al/Si collisions at 5.4 GeV. The spectra of charged and neutral kaons were measured for different centralities [90], and the type of presented uncertainties is not clear. Finally, we also exclude NA35 kaon yields from S+Ag collisions at 19.4 GeV [90, 39]. The type of uncertainties for charged kaon yields [90] is not specified, and the charged and neutral kaons might have been measured for different centralities [90, 91, 39].
NA61/SHINE experiment | |||||
Ar+Sc collisions at 11.9 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
3.732 0.016 0.148 | 0.15 | 0–10% | 0.0 < < 0.2 | [18] | |
2.029 0.012 0.069 | 0.070 | 0–10% | 0.0 < < 0.2 | [18] | |
2.433 0.027 0.102 | 0.11 | 0–10% | 0 | this analysis | |
HADES experiment | |||||
Ar+KCl collisions at 2.6 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
0.028 0.002 0.0014 (∗) | 0.0024 | 0–35% | extrapolated to | [43] | |
0.00071 0.00015 0.000032 (∗) | 0.00015 | 0–35% | extrapolated to | [43] | |
0.0115 0.0005 0.0009 | 0.0010 | 0–35% | extrapolated to | [44] | |
STAR (BES I) experiment | |||||
Au+Au collisions at 7.7 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
20.8 | 1.7 | 0–5% | 0.1 < < 0.1 | [30] | |
7.7 | 0.6 | 0–5% | 0.1 < < 0.1 | [30] | |
12.67 0.12 0.44 | 0.46 | 0–5% | 0.5 < < 0.5 | [31] | |
STAR (BES I) experiment | |||||
Au+Au collisions at 11.5 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
25.0 | 2.5 | 0–5% | 0.1 < < 0.1 | [30] | |
12.3 | 1.2 | 0–5% | 0.1 < < 0.1 | [30] | |
15.93 0.12 0.58 | 0.59 | 0–5% | 0.5 < < 0.5 | [31] | |
STAR (BES I) experiment | |||||
Au+Au collisions at 19.6 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
29.6 | 2.9 | 0–5% | 0.1 < < 0.1 | [30] | |
18.8 | 1.9 | 0–5% | 0.1 < < 0.1 | [30] | |
20.89 0.08 0.67 | 0.67 | 0–5% | 0.5 < < 0.5 | [31] | |
STAR (BES I) experiment | |||||
Au+Au collisions at 27 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
31.1 | 2.8 | 0–5% | 0.1 < < 0.1 | [30] | |
22.6 | 2.0 | 0–5% | 0.1 < < 0.1 | [30] | |
23.24 0.09 0.70 | 0.71 | 0–5% | 0.5 < < 0.5 | [31] | |
STAR (BES I) experiment | |||||
Au+Au collisions at 39 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
32.0 | 2.9 | 0–5% | 0.1 < < 0.1 | [30] | |
25.0 | 2.3 | 0–5% | 0.1 < < 0.1 | [30] | |
24.9 0.1 1.7 | 1.7 | 0–5% | 0.5 < < 0.5 | [31] | |
NA49 experiment | |||||
Pb+Pb collisions at 7.6 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
52.9 0.9 3.5 (∗) | 3.6 | 0–7.2% | extrapolated to | [40] | |
16.0 0.2 0.4 | 0.45 | 0–7.2% | extrapolated to | [40] | |
29.3 0.3 2.9 | 2.9 | 0–7.2% | extrapolated to | [42] | |
NA49 experiment | |||||
Pb+Pb collisions at 8.7 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
59.1 1.9 3 | 3.6 | 0–7.2% | extrapolated to | [41] | |
19.2 0.5 1.0 | 1.1 | 0–7.2% | extrapolated to | [41] | |
34.2 0.2 3.4 | 3.4 | 0–7.2% | extrapolated to | [42] | |
CERES experiment | |||||
Pb+Au collisions at 17.3 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
31.8 0.6 2.5 | 2.6 | 0–7% | 0 | [27] | |
19.3 0.4 2.0 | 2.0 | 0–7% | 0 | [27] | |
21.2 0.9 1.7 | 1.9 | 0–7% | 0 | [28, 29] | |
NA35 experiment | |||||
S+S collisions at 19.4 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
12.5 0.4 0.375 (∗) | 0.55 | 0–2% | extrapolated to | [38] | |
6.9 0.4 0.207 (∗) | 0.45 | 0–2% | extrapolated to | [38] | |
10.5 | 1.7 | 0–2% | extrapolated to | [39] | |
STAR experiment | |||||
Au+Au collisions at 62.4 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
37.6 | 2.7 | 0–5% | 0.1 < < 0.1 | [33] | |
32.4 | 2.3 | 0–5% | 0.1 < < 0.1 | [33] | |
27.4 0.6 2.9 | 3.0 | 0–5% | 1 < < 1 | [34] | |
STAR experiment | |||||
Au+Au collisions at 130 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
46.2 0.6 6.0 | 6.0 | 0–6% | < < | [32] | |
41.9 0.6 5.4 | 5.4 | 0–6% | < < | [32] | |
33.9 1.1 5.1 | 5.2 | 0–6% | 0.5 < < 0.5 | [32] | |
STAR experiment | |||||
Au+Au collisions at 200 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
51.3 | 6.5 | 0–5% | 0.1 < < 0.1 | [33] | |
49.5 | 6.2 | 0–5% | 0.1 < < 0.1 | [33] | |
43.5 | 2.4 | 0–5% | 0.5 < < 0.5 | [35] | |
ALICE experiment | |||||
Pb+Pb collisions at 2760 GeV | |||||
hadron | Yields () | Centrality | ranges | Ref. | |
109 | 9 | 0–5% | 0.5 < < 0.5 | [36] | |
109 | 9 | 0–5% | 0.5 < < 0.5 | [36] | |
110 | 10 | 0–5% | 0.5 < < 0.5 | [37] |
Table LABEL:Rkvalues presents the ratios of charged-to-neutral kaons from various experiments, with estimated statistical and total uncertainties where available. Taking into account the possibility of using our compilation in future analyses, the numerical values in Table LABEL:Rkvalues are presented with unusually high precision.
Experiment | Collision system | (GeV) | |||
NA61/SHINE | Ar+Sc | 11.9 | 1.1839 | 0.0138 | 0.0615 |
HADES | Ar+KCl | 2.6 | 1.2483 | 0.1027 | 0.1545 |
STAR (BES I) | Au+Au | 7.7 | 1.1247 | - | 0.0819 |
STAR (BES I) | Au+Au | 11.5 | 1.1707 | - | 0.0973 |
STAR (BES I) | Au+Au | 19.6 | 1.1584 | - | 0.0910 |
STAR (BES I) | Au+Au | 27 | 1.1553 | - | 0.0819 |
STAR (BES I) | Au+Au | 39 | 1.1446 | - | 0.1079 |
NA49 | Pb+Pb | 7.6 | 1.1758 | 0.0198 | 0.1325 |
NA49 | Pb+Pb | 8.7 | 1.1447 | 0.0295 | 0.1263 |
CERES | Pb+Au | 17.3 | 1.2052 | 0.0539 | 0.1340 |
NA35 | S+S | 19.4 | 0.9238 | - | 0.1533 |
STAR | Au+Au | 62.4 | 1.2774 | - | 0.1525 |
STAR | Au+Au | 130 | 1.2994 | - | 0.2331 |
STAR | Au+Au | 200 | 1.1586 | - | 0.1214 |
ALICE | Pb+Pb | 2760 | 0.9909 | - | 0.1071 |
C. Models
Hadron Resonance Gas model. We use the Hadron Resonance Gas model implementation from Ref. [15] to quantify the isospin-breaking effects and their interplay. HRG includes all hadrons and resonances with confirmed status in the PDG tables [92]. The PDG-listed masses, charges, lifetimes, and decay modes are used. Thus, HRG includes the isospin-symmetry violation due to masses and branching ratios of hadrons and resonances.
In HRG calculations, the exact net strangeness conservation is enforced, i.e., the calculations are done within the strangeness canonical ensemble (SCE) [93, 94]. The model parameters are baryo-chemical potential, , temperature, , volume of the system , and the strangeness under-saturation parameter (see Ref. [95]). We adopt the simple parametrization of and as a function of collision energy introduced in Ref. [96]. We have checked that is weakly sensitive to the strangeness suppression effect introduced by having the parameter . Therefore, calculations are done for for simplicity. The energy-dependent Breit-Wigner spectra [97] model the resonance widths.
Figure 3 (black line) shows the HRG predictions for as a function of collision energy for . At low collision energies, due to the enhancement of neutral kaon production caused by a larger number of neutrons than protons, . On the other hand, at high collision energies, HRG predicts due to the mass difference between charged and neutral kaons produced directly. Finally, when kaon production via resonance decays is included. This is mostly due to decays, which strongly prefer the decay into charged kaons over the one into neutral kaons.
We have checked that the system electric to baryon charge ratio significantly affects up to GeV. At higher energies, pions dominate, and total electric and baryon charges are significantly larger than the corresponding net charges. Thus, becomes increasingly less sensitive to with an increase in .
We have checked that the strangeness grand-canonical ensemble and other popular parametrizations [98, 47] of the model parameters as a function of collision energy lead to quantitatively similar results for , for GeV. The uncertainties of estimated by changing the parametrization of model parameters [96] are less than % for GeV. The effect of including light nuclei in the particle list is negligible.
A dedicated discussion for resonances and is needed. For three strong decay channels are reported as seen in PDG [3]: , , and , where the latter is phase-space suppressed because . The PDG average ratio of decay widths [3] implies that the branching ratio amounts to :
(5) |
For the strong decay channels and have been seen [3]. A list of measurements for the branching ratios is reported, each of them larger than , but no PDG average is provided. Building an average for the branching ratio leads to , but to enhance reliability we vary the range between 10–40%. As a starting point, the HRG approach used in this work assumes for both resonances. Taking into account the difference of the charged and neutral kaon masses within Flatté-like distributions [99, 100, 101], the ratio of charged and neutral decay rates is about and remains smaller than when the distribution parameters are varied within reasonable ranges (see, e.g. the compilation in Ref. [49]). We therefore recalculated the HRG predictions assuming more charged than neutral kaons produced by decays of and . The ratio increases by less than . Under the extreme assumption (no decays to neutral kaons) for both resonances, and , and taking the upper limit for the branching ratio of equal to , increases by up 3% at high collision energies.
UrQMD model. The UrQMD transport model [45, 16, 46] describes + collisions by explicitly propagating hadrons in phase space. During the propagation, rescattering among hadrons takes place. The particle production in this model happens via resonance decay or string excitation and fragmentation following the LUND model [102].
The gray squares in Fig. 3 indicate the UrQMD model predictions. Here, we have considered central Au+Au collisions ( 197, 79, ). The predictions are shown within the range of 2.4 to 20 GeV. At each energy, events are used for the analysis.
One sees that for GeV, the predictions of UrQMD and HRG are similar. At higher energies, the ratio in HRG is systematically higher than the one predicted by UrQMD. This is likely caused by UrQMD assuming -meson decays to be exactly isospin symmetric instead of taking the branching ratios from PDG. This is the reason for showing the UrQMD predictions only up to 20 GeV.
D. Extended data
In this part, we present values of track pair cuts used in the analysis (Table 4), examples of fitted invariant mass distributions (Fig. 5), mean lifetime (Fig. 6) and transverse momentum distributions (Fig. 7) of in rapidity bins.
rapidity bin | ||||||||
---|---|---|---|---|---|---|---|---|
cut value | cosine of angle | 0.999 | 0.9995 | 0.9995 | 0.9995 | 0.9995 | 0.9999 | 0.9999 |
distance | 5 cm | 5 cm | 7.5 cm | 12.5 cm | 12.5 cm | 15 cm | 12.5 cm |



Data availability
All data shown in plots are publicly available on the HEPdata repository (https://hepdata.net).
Code availability
The authors can provide the NA61/SHINE source code used upon reasonable request.
References
- [1] H. Fritzsch, M. Gell-Mann, and H. Leutwyler, “Advantages of the color octet gluon picture,” Phys. Lett. B 47 (1973) 365–368.
- [2] W. Heisenberg, “On the structure of atomic nuclei,” Z. Phys. 77 (1932) 1–11.
- [3] S. Navas et al., [Particle Data Group Collab.], “Review of Particle Physics,” Phys. Rev. D 110 (2024) 030001.
- [4] J. L. Kneur and A. Neveu, “ from renormalization group optimized perturbation,” Phys. Rev. D 85 (2012) 014005, arXiv:1108.3501 [hep-ph].
- [5] A. Deur, S. J. Brodsky, and G. F. de Téramond, “The QCD running coupling,” Nucl. Phys. 90 (2016) 1, arXiv:1604.08082 [hep-ph].
- [6] M. R. Pennington, “Swimming with Quarks,” J. Phys. Conf. Ser. 18 (2005) 1–73, arXiv:hep-ph/0504262.
- [7] J. M. Alarcón, J. Martin Camalich, and J. A. Oller, “The chiral representation of the scattering amplitude and the pion-nucleon sigma term,” Phys. Rev. D 85 (2012) 051503, arXiv:1110.3797 [hep-ph].
- [8] I. Shmushkevich , Dokl. Akad. Nauk SSSR 103 (1955) 235.
- [9] N. Dushin and I. Shmushkevich , Dokl. Akad. Nauk SSSR 106 (1956) 801.
- [10] G. Pinski, A. J. MacFarlane, and G. Sudarshan, “Shmushkevich’s Method for a Charge-Independent Theory,” Phys. Rev. 140 (1965) B1045.
- [11] C. G. Wohl, “Isospin relations by counting,” Am. J. Phys. 50 (1982) 748–753.
- [12] P. B. Pal, “An Introductory Course of Particle Physics,” CRC Press, Taylor & Francis Group (2015).
-
[13]
M. Gazdzicki and O. Hansen, “Hadron production in nucleon-nucleon collisions
at 200 GeV/ –
A Compilation,” Nucl. Phys. A 528 (1991) 754–770. - [14] W. Brylinski, M. Gazdzicki, F. Giacosa, M. Gorenstein, R. Poberezhnyuk, S. Samanta, and H. Stroebele, “Large isospin symmetry breaking in kaon production at high energies,” arXiv:2312.07176 [nucl-th].
- [15] V. Vovchenko and H. Stoecker, “THERMAL-FIST: A package for heavy-ion collisions and hadronic equation of state,” Comput. Phys. Commun. 244 (2019) 295–310, arXiv:1901.05249 [nucl-th].
- [16] M. Bleicher et al., “Relativistic hadron-hadron collisions in the ultra-relativistic quantum molecular dynamics model,” J. Phys. G 25 (1999) 1859–1896, arXiv:hep-ph/9909407.
- [17] N. Abgrall et al., [NA61/SHINE Collab.], “NA61/SHINE facility at the CERN SPS: beams and detector system,” J. Inst. 9 (2014) P06005, arXiv:1401.4699.
- [18] H. Adhikary et al., [NA61/SHINE Collab.], “Measurements of , , and spectra in 40Ar+45Sc collisions at 13 to 150 GeV/,” Eur. Phys. J. C 84 no. 4, (2024) 416, arXiv:2308.16683 [nucl-ex].
- [19] M. Lewicki, ”Charged hadron production in central Ar+Sc collisions at the CERN SPS,” PhD thesis CERN-THESIS-2020-349, University of Wrocław, 2020. https://cds.cern.ch/record/2772291.
- [20] P. Podlaski, ”Study of charged hadron production with tof-dE/dx identification method in central Ar+Sc collisions in NA61/SHINE experiment at CERN,” PhD thesis CERN-THESIS-2021-250, University of Warsaw, 2021. https://cds.cern.ch/record/2799198.
- [21] N. Abgrall et al., [NA61/SHINE Collab.], “Measurements of , , , and proton production in proton–carbon interactions at 31 GeV/ with the NA61/SHINE spectrometer at the CERN SPS,” Eur. Phys. J. C 76 no. 2, (2016) 84, arXiv:1510.02703 [hep-ex].
- [22] H. Adhikary et al., [NA61/SHINE Collab.], “Measurements of , , and production in 120 GeV/ p+C interactions,” Phys. Rev. D 107 no. 7, (2023) 072004, arXiv:2211.00183 [hep-ex].
- [23] H. Adhikary et al., [NA61/SHINE Collab.], “Measurement of hadron production in –C interactions at 158 and 350 GeV/ with NA61/SHINE at the CERN SPS,” Phys. Rev. D 107 no. 6, (2023) 062004, arXiv:2209.10561 [nucl-ex].
- [24] A. Aduszkiewicz et al., [NA61/SHINE Collab.], “Measurements of hadron production in +C and +Be interactions at 60 GeV/,” Phys. Rev. D 100 no. 11, (2019) 112004, arXiv:1909.06294 [hep-ex].
- [25] A. Acharya et al., [NA61/SHINE Collab.], “ meson production in inelastic + interactions at 158 beam momentum measured by NA61/SHINE at the CERN SPS,” Eur. Phys. J. C 82 no. 1, (2022) 96, arXiv:2106.07535 [hep-ex].
- [26] N. Abgrall et al., [NA61/SHINE Collab.], “ meson production in inelastic + interactions at 31, 40 and 80 GeV/ beam momentum measured by NA61/SHINE at the CERN SPS,” Eur. Phys. J. C 84 no. 8, (2024) 820, arXiv:2402.17025 [hep-ex].
- [27] M. Kaliský, ”Reconstruction of charged kaons in the three pion decay channel in Pb+Au 158 AGeV collisions by the CERES experiment,” PhD thesis CERN-THESIS-2007-132, Technical University of Darmstadt, 2007. https://cds.cern.ch/record/1497739.
- [28] S. Radomski, [CERES Collab.], “CERES measurement of strangeness production at top SPS energy,” J. Phys. G 35 (2008) 044003.
- [29] S. Radomski, ”Neutral strange particle production at top SPS energy measured by the CERES experiment,” PhD thesis CERN-THESIS-2006-117, Technical University of Darmstadt, 2006. http://cds.cern.ch/record/1497741.
- [30] L. Adamczyk et al., [STAR Collab.], “Bulk properties of the medium produced in relativistic heavy-ion collisions from the beam energy scan program,” Phys. Rev. C 96 no. 4, (2017) 044904, arXiv:1701.07065 [nucl-ex].
- [31] J. Adam et al., [STAR Collab.], “Strange hadron production in Au+Au collisions at 7.7, 11.5, 19.6, 27, and 39 GeV,” Phys. Rev. C 102 no. 3, (2020) 034909, arXiv:1906.03732 [nucl-ex].
- [32] C. Adler et al., [STAR Collab.], “Kaon production and kaon to pion ratio in Au+Au collisions at 130 GeV,” Phys. Lett. B 595 (2004) 143–150, arXiv:nucl-ex/0206008.
- [33] B. I. Abelev et al., [STAR Collab.], “Systematic measurements of identified particle spectra in , +Au and Au+Au collisions at the STAR detector,” Phys. Rev. C 79 (2009) 034909, arXiv:0808.2041 [nucl-ex].
- [34] M. M. Aggarwal et al., [STAR Collab.], “Strange and multistrange particle production in Au+Au collisions at 62.4 GeV,” Phys. Rev. C 83 (2011) 024901, arXiv:1010.0142 [nucl-ex]. [Erratum: Phys.Rev.C 107, 049903 (2023)].
- [35] G. Agakishiev et al., [STAR Collab.], “Strangeness Enhancement in Cu-Cu and Au-Au Collisions at GeV,” Phys. Rev. Lett. 108 (2012) 072301, arXiv:1107.2955 [nucl-ex].
- [36] B. Abelev et al., [ALICE Collab.], “Centrality dependence of , , production in Pb-Pb collisions at 2.76 TeV,” Phys. Rev. C 88 (2013) 044910, arXiv:1303.0737 [hep-ex].
- [37] B. Abelev et al., [ALICE Collab.], “ and Production in Pb-Pb Collisions at 2.76 TeV,” Phys. Rev. Lett. 111 (2013) 222301, arXiv:1307.5530 [nucl-ex].
- [38] J. Baechler et al., [NA35 Collab.], “Production of charged kaons in proton-nucleus and nucleus-nucleus collisions at 200 GeV/nucleon,” Z. Phys. C 58 (1993) 367–374.
- [39] T. Alber et al., [NA35 Collab.], “Strange particle production in nuclear collisions at 200 GeV per nucleon,” Z. Phys. C 64 (1994) 195–207.
- [40] C. Alt et al., [NA49 Collab.], “Pion and kaon production in central Pb+Pb collisions at 20 and 30 GeV: Evidence for the onset of deconfinement,” Phys. Rev. C 77 (2008) 024903, arXiv:0710.0118 [nucl-ex].
- [41] S. V. Afanasiev et al., [NA49 Collab.], “Energy dependence of pion and kaon production in central Pb+Pb collisions,” Phys. Rev. C 66 (2002) 054902, arXiv:nucl-ex/0205002.
- [42] C. Strabel, ”Energieabhangigkeit der K-Produktion in zentralen Pb+Pb Reaktionen,” MSc thesis NA49-PUBLIC, Johann Wolfgang Goethe-Universitat, 2006. https://edms.cern.ch/document/2958534/1.
- [43] G. Agakishiev et al., [HADES Collab.], “ decay: A relevant source for K- production at energies available at the GSI Schwerionen-Synchrotron (SIS)?,” Phys. Rev. C 80 (2009) 025209, arXiv:0902.3487 [nucl-ex].
- [44] G. Agakishiev et al., [HADES Collab.], “In-medium effects on mesons in relativistic heavy-ion collisions,” Phys. Rev. C 82 (2010) 044907, arXiv:1004.3881 [nucl-ex].
- [45] S. A. Bass et al., “Microscopic models for ultrarelativistic heavy ion collisions,” Prog. Part. Nucl. Phys. 41 (1998) 255–369, arXiv:nucl-th/9803035.
- [46] M. Bleicher and E. Bratkovskaya, “Modelling relativistic heavy-ion collisions with dynamical transport approaches,” Prog. Part. Nucl. Phys. 122 (2022) 103920.
- [47] F. Becattini, J. Manninen, and M. Gaździcki, “Energy and system size dependence of chemical freeze-out in relativistic nuclear collisions,” Phys. Rev. C73 (2006) 044905, arXiv:hep-ph/0511092 [hep-ph].
- [48] N. N. Achasov, S. A. Devyanin, and G. N. Shestakov, “S∗– mixing as a threshold phenomenon,” Phys. Lett. B 88 (1979) 367–371.
- [49] J.-J. Wu and B. S. Zou, “Study of (980) - (980) mixing from (980) (980) transition,” Phys. Rev. D 78 (2008) 074017, arXiv:0808.2683 [hep-ph].
- [50] M. Ablikim et al., [BESIII Collab.], “Observation of - Mixing,” Phys. Rev. Lett. 121 no. 2, (2018) 022001, arXiv:1802.00583 [hep-ex].
- [51] S. Krewald, R. H. Lemmer, and F. P. Sassen, “Lifetime of kaonium,” Phys. Rev. D 69 (2004) 016003, arXiv:hep-ph/0307288.
- [52] Q. J. Ye et al., “Production of pairs in proton-proton collisions at 2.83 GeV,” Phys. Rev. C 85 (2012) 035211, arXiv:1202.5451 [nucl-ex].
- [53] C. Wilkin, “The legacy of the experimental hadron physics programme at COSY,” Eur. Phys. J. A 53 no. 6, (2017) 114, arXiv:1611.07250 [nucl-ex].
- [54] T. Goldman, K. R. Maltman, and G. J. Stephenson, “The finite QED correction to the quark-gluon vertex,” Phys. Lett. B 228 (1989) 396–400.
- [55] J. Letessier and J. Rafelski, “Chemical nonequilibrium in high-energy nuclear collisions,” J. Phys. G 25 (1999) 295–309, arXiv:hep-ph/9810332.
- [56] E. E. Salpeter, “Electron Screening and Thermonuclear Reactions,” Austral. J. Phys. 7 (1954) 373–388.
- [57] C. Grayson, C. T. Yang, M. Formanek, and J. Rafelski, “Self-consistent Strong Screening Applied to Thermonuclear Reactions,” Astrophys. J. 976 no. 1, (2024) 31, arXiv:2406.13055 [nucl-th].
- [58] S. M. Lenzi and M. A. Bentley, “Test of Isospin Symmetry Along the Line,” Lect. Notes Phys. 764 (2009) 57–98.
- [59] M. Thiel, C. Sfienti, J. Piekarewicz, C. J. Horowitz, and M. Vanderhaeghen, “Neutron skins of atomic nuclei: per aspera ad astra,” J. Phys. G 46 no. 9, (2019) 093003, arXiv:1904.12269 [nucl-ex].
- [60] T. Otsuka, T. Abe, T. Yoshida, Y. Tsunoda, N. Shimizu, N. Itagaki, Y. Utsuno, J. Vary, P. Maris, and H. Ueno, “-Clustering in atomic nuclei from first principles with statistical learning and the Hoyle state character,” Nature Commun. 13 no. 1, (2022) 2234.
- [61] H. Adhikary et al., [NA61/SHINE Collab.], “Addendum to the NA61/SHINE Proposal: Request for light ions beams in Run 4,” Tech. Rep. CERN-SPSC-2023-022, SPSC-P-330-ADD-14, CERN, Geneva, 2023. https://cds.cern.ch/record/2867952.
- [62] H. Adhikary et al., [NA61/SHINE Collab.], “Memorandum requesting use of the allocated test beam for data-taking on +C and +C interactions at 158 GeV/,” Tech. Rep. CERN-SPSC-2024-022, SPSC-M-797, CERN, Geneva, 2024. https://cds.cern.ch/record/2907307.
- [63] M. Bonesini, A. Marchionni, F. Pietropaolo, and T. Tabarelli de Fatis, “On particle production for high energy neutrino beams,” Eur. Phys. J. C 20 (2001) 13–27, arXiv:hep-ph/0101163.
- [64] J. Stepaniak and D. Pszczel, “On the relation between and charged kaon yields in proton–proton collisions,” Eur. Phys. J. C 83 no. 10, (2023) 928, arXiv:2305.03872 [hep-ph].
- [65] M. Petráň, J. Letessier, V. Petráček, and J. Rafelski, “Hadron production and quark-gluon plasma hadronization in Pb-Pb collisions at TeV,” Phys. Rev. C 88 no. 3, (2013) 034907, arXiv:1303.2098 [hep-ph].
- [66] J. Rafelski and M. Petráň, “Universal QGP Hadronization Conditions at RHIC and LHC,” EPJ Web Conf. 78 (2014) 06004, arXiv:1406.1871 [nucl-th].
- [67] J. Rafelski and M. Petran, “QCD Phase Transition Studied by Means of Hadron Production,” Phys. Part. Nucl. 46 no. 5, (2015) 748–755, arXiv:2212.13302 [hep-ph].
- [68] R. D. Pisarski and F. Wilczek, “Remarks on the chiral phase transition in chromodynamics,” Phys. Rev. D 29 (1984) 338–341.
- [69] F. Giacosa, A. Koenigstein, and R. D. Pisarski, “How the axial anomaly controls flavor mixing among mesons,” Phys. Rev. D 97 no. 9, (2018) 091901, arXiv:1709.07454 [hep-ph].
- [70] F. Giacosa, S. Jafarzade, and R. D. Pisarski, “Anomalous interactions between mesons with nonzero spin and glueballs,” Phys. Rev. D 109 no. 7, (2024) L071502, arXiv:2309.00086 [hep-ph].
- [71] A. A. Anselm and M. G. Ryskin, “Production of classical pion field in heavy ion high energy collisions,” Phys. Lett. B 266 (1991) 482–484.
- [72] J.-P. Blaizot and A. Krzywicki, “Soft pion emission in high-energy heavy-ion collisions,” Phys. Rev. D 46 (1992) 246–251.
- [73] K. Rajagopal and F. Wilczek, “Emergence of coherent long wavelength oscillations after a quench: application to QCD,” Nucl. Phys. B 404 (1993) 577–589, arXiv:hep-ph/9303281.
- [74] J. D. Bjorken, K. L. Kowalski, and C. C. Taylor, “Observing Disoriented Chiral Condensates,” in Workshop on Physics at Current Accelerators and the Supercollider. 9, 1993. arXiv:hep-ph/9309235.
- [75] J. Schaffner-Bielich and J. Randrup, “Disoriented chiral condensate dynamics with the SU(3) linear sigma model,” Phys. Rev. C 59 (1999) 3329–3342, arXiv:nucl-th/9812032.
- [76] S. Gavin and J. I. Kapusta, “Kaon and pion fluctuations from small disoriented chiral condensates,” Phys. Rev. C 65 (2002) 054910, arXiv:nucl-th/0112083.
- [77] S. Acharya et al., [ALICE Collab.], “Neutral to charged kaon yield fluctuations in Pb – Pb collisions at 2.76 TeV,” Phys. Lett. B 832 (2022) 137242, arXiv:2112.09482 [nucl-ex].
- [78] J. I. Kapusta, S. Pratt, and M. Singh, “Confronting anomalous kaon correlations measured in Pb-Pb collisions at 2.76 TeV,” Phys. Rev. C 107 no. 1, (2023) 014913, arXiv:2210.03257 [hep-ph].
- [79] J. I. Kapusta, S. Pratt, and M. Singh, “Disoriented isospin condensates may be the source of anomalous kaon correlations measured in Pb-Pb collisions at 2.76 TeV,” Phys. Rev. C 109 no. 3, (2024) L031902, arXiv:2306.13280 [hep-ph].
- [80] M. Singh et al., “Disoriented Isospin Condensates as source of anomalous kaon correlations at LHC,” in 21st International Conference on Strangeness in Quark Matter (SQM 2024). 2024. {https://indico.in2p3.fr/event/29792/contributions/137151}.
- [81] D. Parganlija, P. Kovács, G. Wolf, F. Giacosa, and D. H. Rischke, “Meson vacuum phenomenology in a three-flavor linear sigma model with (axial-)vector mesons,” Phys. Rev. D 87 no. 1, (2013) 014011, arXiv:1208.0585 [hep-ph].
- [82] P. Kovács, G. Wolf, N. Weickgenannt, and D. H. Rischke, “Phenomenology of isospin-symmetry breaking with vector mesons,” Phys. Rev. D 109 no. 9, (2024) 096007, arXiv:2401.04527 [hep-ph].
- [83] W. Bryliński, ”Study of meson production in central Ar+Sc collisions at SPS energies,” PhD thesis CERN-THESIS-2023-404, Warsaw University of Technology, 2023. https://cds.cern.ch/record/2907562.
- [84] K. Werner, “The hadronic interaction model EPOS,” Nucl. Phys. Proc. Suppl. 175-176 (2008) 81–87.
- [85] R. Brun, R. Hagelberg, M. Hansroul, and J. C. Lassalle, “Simulation Program for Particle Physics Experiments, GEANT: User Guide and Reference Manual,” Tech. Rep. CERN-DD-78-2-REV, CERN-DD-78-2, CERN, Geneva, 1978. https://cds.cern.ch/record/118715.
- [86] J. Adamczewski-Musch et al., [HADES Collab.], “Deep sub-threshold production in Au+Au collisions,” Phys. Lett. B 778 (2018) 403–407, arXiv:1703.08418 [nucl-ex].
- [87] J. Adamczewski-Musch et al., [HADES Collab.], “Sub-threshold production of K mesons and hyperons in Au+Au collisions at 2.4 GeV,” Phys. Lett. B 793 (2019) 457–463, arXiv:1812.07304 [nucl-ex].
- [88] P. Gasik et al., [FOPI Collab.], “Strange meson production in Al+Al collisions at 1.9 A GeV,” Eur. Phys. J. A 52 no. 6, (2016) 177, arXiv:1512.06988 [nucl-ex].
- [89] X. Lopez et al., [FOPI Collab.], “Measurement of and mesons in Al+Al collisions at 1.9 GeV,” Phys. Rev. C 81 (2010) 061902, arXiv:1006.1905 [nucl-ex].
- [90] M. Gaździcki and D. Röhrich, “Strangeness in nuclear collisions,” Z. Phys. C 71 (1996) 55–64, arXiv:hep-ex/9607004.
- [91] D. Röhrich et al., [NA35 Collab.], “Hadron Production in S+Ag and S+Au Collisions at 200 GeV/Nucleon,” Nucl. Phys. A 566 (1994) 35C–44C.
- [92] C. Patrignani et al., [Particle Data Group Collab.], “Review of Particle Physics,” Chin. Phys. C 40 no. 10, (2016) 100001.
- [93] P. Braun-Munzinger, J. Cleymans, H. Oeschler, and K. Redlich, “Maximum relative strangeness content in heavy-ion collisions around 30 GeV,” Nucl. Phys. A 697 (2002) 902–912, arXiv:hep-ph/0106066.
- [94] H. Oeschler, J. Cleymans, B. Hippolyte, K. Redlich, and N. Sharma, “Thermal Model Description of Collisions of Small Nuclei,” arXiv:1603.09553 [hep-ph].
- [95] J. Rafelski, “Melting hadrons, boiling quarks,” Eur. Phys. J. A 51 no. 9, (2015) 114, arXiv:1508.03260 [nucl-th].
- [96] R. V. Poberezhnyuk, V. Vovchenko, A. Motornenko, M. I. Gorenstein, and H. Stöcker, “Chemical freeze-out conditions and fluctuations of conserved charges in heavy-ion collisions within quantum van der Waals model,” Phys. Rev. C 100 no. 5, (2019) 054904, arXiv:1906.01954 [hep-ph].
- [97] V. Vovchenko, M. I. Gorenstein, and H. Stoecker, “Finite resonance widths influence the thermal-model description of hadron yields,” Phys. Rev. C 98 no. 3, (2018) 034906, arXiv:1807.02079 [nucl-th].
- [98] J. Cleymans, H. Oeschler, K. Redlich, and S. Wheaton, “Comparison of chemical freeze-out criteria in heavy-ion collisions,” Phys. Rev. C 73 (2006) 034905, arXiv:hep-ph/0511094.
- [99] S. M. Flatté, “On the nature of mesons,” Phys. Lett. B 63 (1976) 228–230.
- [100] V. Baru, J. Haidenbauer, C. Hanhart, A. E. Kudryavtsev, and U.-G. Meißner, “Flatté-like distributions and the (980) / (980) mesons,” Eur. Phys. J. A 23 (2005) 523–533, arXiv:nucl-th/0410099.
- [101] F. Giacosa, A. Okopińska, and V. Shastry, “A simple alternative to the relativistic Breit–Wigner distribution,” Eur. Phys. J. A 57 no. 12, (2021) 336, arXiv:2106.03749 [hep-ph].
- [102] B. Andersson, G. Gustafson, and B. Södeberg, “A General Model for Jet Fragmentation,” Z. Phys. C 20 (1983) 317.
Acknowledgments
We would like to thank the CERN EP, BE, HSE and EN Departments for the strong support of NA61/SHINE. We also gratefully acknowledge discussions with Claudia Ahdida, Wojciech Broniowski, Marco van Leeuwen, Krzysztof Golec-Biernat, Stanisław Mrówczyński, Owe Philippsen, Rob Pisarski, Krishna Rajagopal, Jan Rafelski, Jan Steinheimer, Leonardo Tinti, and Volodymyr Vovchenko for fruitful discussions and comments at various stages of the preparation of this article.
This work was supported by
the Hungarian Scientific Research Fund (grant NKFIH 138136/137812/138152 and TKP2021-NKTA-64),
the Polish Ministry of Science and Higher Education
(DIR/WK/2016/2017/10-1, WUT ID-UB), the National Science Centre Poland (grants
2014/14/E/ST2/00018,
2016/21/D/ST2/01983,
2017/25/N/ST2/02575,
2018/29/N/ST2/02595,
2018/30/A/ST2/00226,
2018/31/G/ST2/03910,
2019/33/B/ST2/00613,
2020/39/O/ST2/00277),
the Norwegian Financial Mechanism 2014–2021 (grant 2019/34/H/ST2/00585),
the Polish Minister of Education and Science (contract No. 2021/WK/10),
the Internationalization of the Jan Kochanowski University Doctoral School through the Polish Academy Agency for Academic Exchange NAWA STER No. BPI/STE/2023/1/00014,
the European Union’s Horizon 2020 research and innovation programme under grant agreement No. 871072,
the Ministry of Education, Culture, Sports,
Science and Technology, Japan, Grant-in-Aid for Scientific
Research (grants 18071005, 19034011, 19740162, 20740160 and 20039012,22H04943),
the German Research Foundation DFG (grants GA 1480/8-1 and project 426579465),
the Bulgarian Ministry of Education and Science within the National
Roadmap for Research Infrastructures 2020–2027, contract No. D01-374/18.12.2020,
Serbian Ministry of Science, Technological Development and Innovation (grant
OI171002), Swiss Nationalfonds Foundation (grant 200020117913/1),
ETH Research Grant TH-01 07-3, National Science Foundation grant
PHY-2013228 and the Fermi National Accelerator Laboratory (Fermilab),
a U.S. Department of Energy, Office of Science, HEP User Facility
managed by Fermi Research Alliance, LLC (FRA), acting under Contract
No. DE-AC02-07CH11359 and the IN2P3-CNRS (France),
the German Research Foundation grants GA1480/8-1, and the Alexander von Humboldt
Foundation.
The data used in this article were collected before February 2022.
Author contributions
The NA61/SHINE Collaboration obtained the new experimental results presented in the paper and prepared a compilation of world data needed to calculate the charged-to-neutral kaon ratios.
The individual authors, F. Giacosa, M. Gorenstein, R. Poberezhniuk, and S. Samanta, contributed to the theoretical aspects of the paper.
Competing interests
The authors declare no competing interests.
The NA61/SHINE Collaboration
H. Adhikary 13,
P. Adrich
15,
K.K. Allison
26,
N. Amin
5,
E.V. Andronov
22,
I.-C. Arsene
12,
M. Bajda
16,
Y. Balkova
18,
D. Battaglia
25,
A. Bazgir
13,
S. Bhosale
14,
M. Bielewicz
15,
A. Blondel
4,
M. Bogomilov
2,
Y. Bondar
13,
A. Brandin 22,
W. Bryliński
21,
J. Brzychczyk
16,
M. Buryakov
22,
A.F. Camino 28,
M. Ćirković
23,
M. Csanád
8,
J. Cybowska
21,
T. Czopowicz
13,
C. Dalmazzone
4,
N. Davis
14,
A. Dmitriev
22,
P. von Doetinchem
27,
W. Dominik
19,
J. Dumarchez
4,
R. Engel
5,
G.A. Feofilov
22,
L. Fields
25,
Z. Fodor
7,20,
M. Friend
9,
M. Gaździcki
13,
O. Golosov
22,
V. Golovatyuk
22,
M. Golubeva
22,
K. Grebieszkow
21,
F. Guber
22,
S.N. Igolkin 22,
S. Ilieva
2,
A. Ivashkin
22,
A. Izvestnyy
22,
N. Kargin 22,
N. Karpushkin
22,
E. Kashirin
22,
M. Kiełbowicz
14,
V.A. Kireyeu
22,
R. Kolesnikov
22,
D. Kolev
2,
Y. Koshio 10,
V.N. Kovalenko
22,
S. Kowalski
18,
B. Kozłowski
21,
A. Krasnoperov
22,
W. Kucewicz
17,
M. Kuchowicz
20,
M. Kuich
19,
A. Kurepin
22,
A. László
7,
M. Lewicki
20,
G. Lykasov
22,
V.V. Lyubushkin
22,
M. Maćkowiak-Pawłowska
21,
Z. Majka
16†,
A. Makhnev
22,
B. Maksiak
15,
A.I. Malakhov
22,
A. Marcinek
14,
A.D. Marino
26,
H.-J. Mathes
5,
T. Matulewicz
19,
V. Matveev
22,
G.L. Melkumov
22,
A. Merzlaya
12,
Ł. Mik
17,
S. Morozov
22,
Y. Nagai
8,
T. Nakadaira
9,
M. Naskret
20,
S. Nishimori
9,
A. Olivier
25,
V. Ozvenchuk
14,
O. Panova
13,
V. Paolone
28,
O. Petukhov
22,
I. Pidhurskyi
13,
R. Płaneta
16,
P. Podlaski
19,
B.A. Popov
22,4,
B. Pórfy
7,8,
D.S. Prokhorova
22,
D. Pszczel
15,
S. Puławski
18,
J. Puzović 23†,
R. Renfordt
18,
L. Ren
26,
V.Z. Reyna Ortiz
13,
D. Röhrich 11,
E. Rondio
15,
M. Roth
5,
Ł. Rozpłochowski
14,
B.T. Rumberger
26,
M. Rumyantsev
22,
A. Rustamov
1,
M. Rybczynski
13,
A. Rybicki
14,
D. Rybka 15,
K. Sakashita
9,
K. Schmidt
18,
A.Yu. Seryakov
22,
P. Seyboth
13,
U.A. Shah
13,
Y. Shiraishi 10,
A. Shukla
27,
M. Słodkowski
21,
P. Staszel
16,
G. Stefanek
13,
J. Stepaniak
15,
M. Strikhanov 22,
H. Ströbele 6,
T. Šuša
3,
Ł. Świderski
15,
J. Szewiński
15,
R. Szukiewicz
20,
A. Taranenko
22,
A. Tefelska
21,
D. Tefelski
21,
V. Tereshchenko 22,
R. Tsenov
2,
L. Turko
20,
T.S. Tveter
12,
M. Unger
5,
M. Urbaniak
18,
F.F. Valiev
22,
D. Veberič
5,
V.V. Vechernin
22,
O. Vitiuk
20,
V. Volkov
22,
A. Wickremasinghe
24,
K. Witek
17,
K. Wójcik
18,
O. Wyszyński
13,
A. Zaitsev
22,
E. Zherebtsova
20,
E.D. Zimmerman
26,
A. Zviagina
22, and
R. Zwaska
24
and individual authors
1 National Nuclear Research Center, Baku, Azerbaijan
2 Faculty of Physics, University of Sofia, Sofia, Bulgaria
3 Ruder Bošković Institute, Zagreb, Croatia
4 LPNHE, Sorbonne University, CNRS/IN2P3, Paris, France
5 Karlsruhe Institute of Technology, Karlsruhe, Germany
6 University of Frankfurt, Frankfurt, Germany
7 HUN-REN Wigner Research Centre for Physics, Budapest, Hungary
8 Eötvös Loránd University, Budapest, Hungary
9 Institute for Particle and Nuclear Studies, Tsukuba, Japan
10 Okayama University, Japan
11 University of Bergen, Bergen, Norway
12 University of Oslo, Oslo, Norway
13 Jan Kochanowski University, Kielce, Poland
14 Institute of Nuclear Physics, Polish Academy of Sciences, Cracow, Poland
15 National Centre for Nuclear Research, Warsaw, Poland
16 Jagiellonian University, Cracow, Poland
17 AGH University of Krakow, Cracow, Poland
18 University of Silesia, Katowice, Poland
19 University of Warsaw, Warsaw, Poland
20 University of Wrocław, Wrocław, Poland
21 Warsaw University of Technology, Warsaw, Poland
22 Affiliated with an institution covered by a cooperation agreement with CERN
23 University of Belgrade, Belgrade, Serbia
24 Fermilab, Batavia, USA
25 University of Notre Dame, Notre Dame, USA
26 University of Colorado, Boulder, USA
27 University of Hawaii at Manoa, Honolulu, USA
28 University of Pittsburgh, Pittsburgh, USA
29 Bogolyubov Institute for Theoretical Physics, Kyiv, Ukraine
30 Frankfurt Institute for Advanced Studies, Giersch Science Center, Frankfurt am Main, Germany
31 University of Houston, Houston, USA
32 School of Applied Sciences, Kalinga Institute of Industrial Technology, Bhubaneswar, Odisha, India