logo

Acceptance effect on the NtNp/Nd2 ratio of light nuclei coalescence yields as a probe of nucleon density fluctuations

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Acceptance effect on the NtNp/Nd2 ratio of light nuclei coalescence yields as a probe of nucleon density fluctuations

An Gu
Michael X. Zhang
Nuclear Science and TechniquesVol.36, No.3Article number 53Published in print Mar 2025Available online 13 Feb 2025
5000

A coalescence model was employed to form deuterons (d), tritons (t), and helium-3 (3He) nuclei from a uniformly-distributed volume of protons (p) and neutrons (n). We studied the ratio NtNp/Nd2 of light nuclei yields as a function of the neutron density fluctuations. We investigated the effect of finite transverse momentum (pT) acceptance on the ratio, in particular, the “extrapolation factor” (f) for the ratio as a function of the pT spectral shape and the magnitude of neutron density fluctuations. The nature of f was found to be monotonic in pT spectra “temperature” parameter and neutron density fluctuation magnitude; variations in the latter are relatively small. We also examined f in realistic simulations using the kinematic distributions of protons measured from the heavy-ion collision data. The nature of f was found to be smooth and monotonic as a function of the beam energy. Therefore, we conclude that extrapolation from limited pT ranges does not create, enhance, or reduce the local peak of the NtNp/Nd2 ratio in the beam energy. Our study provides a necessary benchmark for light nuclei ratios as a probe for nucleon density fluctuations, an important observation in the search for the critical point of nuclear matter.

Heavy-ion collisionCritical pointLight nuclei coalescenceNucleon density fluctuations
1

Introduction

Matter comprises quarks and gluons, which are the most fundamental constituents of nature, together with leptons and gauge bosons. The interactions between quarks and gluons are governed by quantum chromodynamics (QCD). At low temperatures or matter densities, quarks and gluons are confined in hadrons, whereas at high temperatures or matter densities, they are deconfined in an extended volume called the quark–gluon plasma (QGP). The phase transition at low temperature and high matter density is first-order, and at high temperature and low matter density, it is a smooth crossover, as predicted by lattice QCD [1]. It has been conjectured that a critical point (CP) exists in the nuclear matter phase diagram of temperature versus matter density between the first-order phase transition and smooth crossover [2-6]. The correlation length increases dramatically near the CP, causing large fluctuations in conserved quantities such as the net baryon number [7]. Searching for the subject of active research in heavy-ion collisions [8-17].

Light nuclei production is well-modeled by nucleon coalescence [18-25]. The coalescence model predicts that large baryon number fluctuations affect the production rate of light nuclei [26, 27]. For example, the production of tritons (t) is enhanced relative to that of deuterons (d) when there are extra fluctuations in the neutron density because a triton contains two neutrons (n), whereas a deuteron contains only one. As a result, the compound ratio NtNp/Nd2 involving the multiplicities of protons (Np), deuterons (Nd), and tritons (Nt) was enhanced. Similarly, the production of helium-3 (3He) and the ratio of N3heNn/Nd2 with respect to extra fluctuations in the proton density.

Following Ref. [26-28], the average deuteron multiplicity density in the coalescence model is given by n¯d/(32A)=(n¯p+δnp)(n¯n+δnn)=n¯pn¯n+δnpδnn=n¯pn¯n(1+(αΔnn)d), (1) where the protons and neutrons are assumed to be in thermal equilibrium with an effective temperature Teff and A=(2πmNTeff)3/2 is the shorthand notation (mN is the nucleon mass). Here, n¯ denotes the average density. The neutron density fluctuations are denoted as Δnn=(δnn)2/n¯n2, (2) and α=δnpnnn¯pn¯n/Δnn=n¯nn¯pδnpnn(δnn)2 denotes the correlations between proton and neutron number fluctuations. If the proton and neutron numbers fluctuate independently, α=0; if they fluctuate simultaneously, α=1. It is noteworthy that we have thus far neglected the details of deuteron formation, which is determined by the deuteron wavefunction and is often implemented by the Wigner function formalism (see below). The fluctuations affecting deuteron formation are those within the typical volumes of deuterons. We note this in Eq. (1) by the subscript ‘d’ in (αΔnn)d to indicate that it is the average αΔnn within the typical deuteron size that is relevant.

Similarly, the triton average multiplicity density is given by n¯t/(334A2)=(n¯p+δnp)(n¯n+δnn)2=n¯pn¯n2(1+(Δnn)t+2(αΔnn)t+(βΔnn)t), (3) where Δnn=(δnn)3/n¯n3 and β=δnp(δnn)2n¯pn¯n2/Δnn=n¯nn¯pδnp(δnn)2(δnn)3. Similar to α in Eq. (1), the β parameters denote the three-body correlations. When the proton and neutron numbers fluctuate independently, β=0, and when they fluctuate together, β=1. The subscript ‘t’ in Eq. (3) indicates that only fluctuations averaged within the typical volume of the triton size matter. Note that in Eq. (3), we simply write nn2 for the neutron pair density; however, for an identical particle pair, the multiplicity is N(N-1), so the pair fluctuations of Eq. (2) should be understood as those beyond Poisson fluctuations.

Thus, the compound ratio is given by NpNtNd2=n¯pn¯tn¯d2=1+(Δnn)t+2(αΔnn)t+(βΔnn)t23(1+(αΔnn)d)2. (4) If one neglects α and β, then Eq. (4) is reduced to NpNtNd2123(1+(Δnn)t), (5) as in Ref. [26, 27]. Note that the factor 1/23 originates from the thermal equilibrium assumption of nucleon abundances. Therefore, the NtNp/Nd2 may be a good measure of neutron density fluctuations, a large value of which can signal the CP.

A unique signature of the CP is the nonmonotonic behavior of the ratio NtNp/Nd2 in the beam energy, where a peak of the ratio in a localized region of beam energy can signal large neutron fluctuations and the CP [26, 27]. The STAR experiment at RHIC recently observed non-monotonic behavior of the NtNp/Nd2 ratio in the top 10% of central Au+Au collisions as a function of the nucleon-nucleon center-of-mass energy (sNN) in the Beam Energy Scan (BES) data [29].

A nonmonotonic bump was observed in the ratio localized in the energy region sNN=2030 GeV. It should be noted that the local bump is prominent in the NtNp/Nd2 ratio of the extrapolated yields to all transverse momenta (pT) but not as prominent in the measured fiducial range of 0.5<pT/A<1.0 GeV/c and 0.4<pT/A<1.2 GeV/c (where A is the mass number corresponding to each light nucleus in the ratio) [29].

This is illustrated in the left panels of Figs. 1 where the STAR-measured NtNp/Nd2 ratios from the total extrapolated yields and two measured pT/A ranges are reproduced. The “extrapolation factor” (f), that is, the NtNp/Nd2 ratio of the total light nuclei yields extrapolated to the entire pT/A range [0–) divided by the ratio of those measured within a fiducial pT/A range, is shown in the right panel of Fig. 1 for the two measured pT/A ranges. It should be noted that the integrated yields of protons, deuterons, and tritons over the full momentum space in the STAR experiment were extrapolated from slightly different ranges of the scaled transverse momentum pT/A. The aforementioned ratio for the measured fiducial range was obtained using particle yields within the same pT/A range as in the STAR experiment. The statistical uncertainties between the fiducial yield of a given pT spectrum and the extrapolated total yield are correlated, as are the systematic uncertainties. Thus, the statistical and systematic uncertainties are considered to be the quadratic difference of the corresponding uncertainties in the NtNp/Nd2 between the fiducial pT/A and total ranges. As expected from the STAR results [29], the f values peaked in the sNN=2030 GeV range and were non-monotonic; the non-monotonic effect was of the order of 10%.

Fig. 1
(Color online) (Left) The NtNp/Nd2 ratio in 0-10% central Au+Au collisions as functions of beam energy sNN measured by STAR [29]. The ratio of the extrapolated total yields and those from two measured pT/A ranges are shown. The ratios from the measured pT/A ranges are shifted in the horizontal axis for clarity. (Right) The extrapolation factor f, i.e., the NtNp/Nd2 ratio from the extrapolated total yields divided by that from the fiducial yields in a given measured pT/A range, is shown for the two measured pT/A ranges as functions of sNN. The statistical and systematic uncertainties on f are the quadratic difference of the corresponding uncertainties of the NtNp/Nd2 ratio between the given measured pT/A range and the total range. The rightmost square is shifted in the horizontal axis for clarity
pic

In this study, we use a toy model to generate nucleons and form d, t, and 3He by using a coalescence model. We study the ratios of the light nuclei yields as functions of the magnitude of neutron multiplicity fluctuations and examine the role of pT acceptance in these ratios. First, we confined ourselves to a simple toy model to gain insight. We then attempted to conduct more realistic simulations mimicking STAR BES data to examine what STAR measurements may entail.

2

Coalescence model

The probability of forming a composite particle from particle 1 at position r1 with momentum p1 and particle 2 at position r2 with momentum p2 is estimated using the Wigner function: W(r1,p1,r2,p2)=g8exp(r122σr2p122σp2), (6) where r12=|r1r2|,p12=μ|p1m1p2m2|, and μ=m1m2m1+m2 denote the reduced mass of a two-body system. The parameter σr is the characteristic coalescence size in the configuration space and σp=1/σr (where =c=1) is that in the momentum space.

Based on the Wigner function expressed in Eq. (6), the root mean square (RMS) radius of the coalesced composite particle is calculated as R=3m1m2/2m1+m2σr. Therefore, the coalescence parameter σr can be determined from the particle size asfollows: σr=m1+m23m1m2/2R. (7) Deuterons coalesce with protons and neutrons. The RMS size of the deuteron is Rd=1.96 fm [30], so for deuteron σr=83Rd=3.20 fm. The triton (helium-3) is formed by the coalescence of a deuteron and neutron (proton), following the same prescription as Eq. (6). The Triton RMS size is Rt=1.59 fm [30]. Therefore, for triton σr=3Rt=2.75 fm. For 3He we also assume σr=2.75 fm to be symmetric.

The gfactor represents the probability of the proper total spin of the composite particle. g=(2S+1)/Πi=12(2si+1), (8) where S is the spin of the composite particle (1 for d, 1/2 for both t and 3He) and si is the spin of each coalescing particle. For the d formed from p and n, the g-factor was 3/4. For t (3He), formed from d and n (p), it is 1/3.

3

Toy-Model Simulation Details

A system of nucleons was generated using a toy model. The nucleons were assumed to be uniformly distributed in a cylinder of 10 fm length and 10 fm radius. The transverse momentum spectra of the nucleons were assumed to be dN/dpTpTexp(pT/T), (9) where the “temperature” parameter is set to T=150 MeV. Note that Eq. (9) is not a thermal distribution for massive particles; we use the term “temperature” for convenience. The azimuthal angle of the momentum vector was uniformly distributed between 0 and 2π. The pseudorapidity η was assumed to be uniform between -1 and 1. The use of rapidity or pseudorapidity was insignificant.

In this study, we did not use a thermal model to predict the average multiplicity; rather, we set the average multiplicities of protons and neutrons to N¯p=N¯n=20. This is simple, because our goal was to study only the effects of neutron multiplicity fluctuations. Therefore, the baseline of the NtNp/Nd2 ratio is not 1/23 but is rather determined by the ratio of the degeneracy g factors as 1334/(34)2=49 [31].

We assigned Poisson fluctuations to the number of protons and only varied the fluctuation magnitude for the number of neutrons (we focused on the compound ratio NtNp/Nd2). The latter is achieved by using negative binomial distributions P(Nn=k)=Ckk+r1(1p)kpr, where r and p are free parameters.

The mean and variance are N¯n=r(1p)p and σNn2=r(1p)p2, respectively. The fluctuation magnitude is given by σNn2=N¯n/pθN¯n (where θ1/p1), which is always larger than Poisson fluctuations unless the probability p=1 when the negative binomial distribution is reduced to Poisson. We used p to control the magnitude of the fluctuations in Nn and selected the proper r value to obtain the desired average neutron multiplicity, N¯n.

It is worth noting that we utilized fluctuations in the total number of neutrons Nn event-by-event to mimic fluctuations in the local neutron number density. The purpose of our study is to investigate the behavior of NtNp/Nd2 as a function of the neutron density fluctuation magnitude but not to suggest that density fluctuations are caused by fluctuations in the total multiplicity. In our simulation, given a total Nn in an event, the neutrons were randomly distributed in the cylinder volume. The fluctuations in Nn determine the magnitude of the neutron density fluctuations, which is averaged over the entire volume. In other words, Δnn in Eq. (2) can be quantified by fluctuations in Nn beyond Poisson, Δnn=(σNn2/N¯n1)/N¯n(θ1)/N¯n. (10) Note that θ quantifies the fluctuations in Nn in the unit of Poisson fluctuations, and θ1 quantifies those beyond Poisson fluctuations.

In total, 1.6 × 108 events were simulated. Deuterons, tritons, and helium-3 were formed by coalescence. For each event, the deuteron formation via double loops over protons and neutrons was considered. For each deuteron, the formation of t and 3He was implemented by looping over the remaining nucleons. The nucleons were randomly reordered such that the probabilities of forming t and 3He were unbiased. A random number uniformly distributed between 0 and 1 was used to determine whether the two particles coalesced into a light nucleus. If the random number is smaller than the Wigner function value in consideration, the corresponding light nucleus is formed. Once a light nucleus formed, the coalescing particles were removed from further consideration.

4

Toy-Model Simulation Results and Discussions

Figure 2 shows the pT spectra of the generated neutrons and the coalesced d, t, and 3He in the left panel, and the right panel those spectra in the scaled pT/A. No extra fluctuations are included beyond Poisson in this figure (i.e., θ=1); therefore, the proton spectrum is identical to that of neutrons. In the right panel, the product of the proton and neutron spectra and the product of the proton and squared neutron spectra are shown as smooth histograms. These products are the corresponding d and t spectra if the coalescence parameters σp=0 and σr=∞ in Eq. (6). The coalesced d and t spectra are steeper than those of the products because of the finite σp and σr implemented in our coalescence model.

Fig. 2
(Color online) (Left) Transverse momentum pT spectra of deuteron, 3He and triton calculated by the coalescence model from a system of average N¯p=20 protons and N¯n=20 neutrons at “temperature” T=150 MeV (Eq. (9)) randomly distributed within a cylinder of 10 fm radius and 10 fm length. No extra fluctuations are included beyond Poisson, i.e., θ=1. (Right) The same spectra plotted as functions of pT/A (A is the corresponding mass number). Superimposed in curves are the products of (dNp/dpT)×(dNn/dpT) and (dNp/dpT)×(dNn/dpT)2 at the same pT/A value, arbitrarily scaled to compare to the shapes of the deuteron and triton spectra, respectively
pic

Figure 3 shows in the left panel the yield ratios of Nd/Np and Nt/Nd as functions of the neutron fluctuation magnitude Δnn in Eq. (10). The Nt/Nd ratio increases with Δnn, which is consistent with the expected stronger effect of the neutron density fluctuations on t than on d production. The Nd/Np ratio decreases slightly, but is statistically significant. This is counterintuitive, as one would expect no dependence because N¯n is the same for all values of Δnn; events with more neutrons would be balanced out by those with fewer neutrons in terms of deuteron production. However, N¯p=20 is fixed and the fluctuations in Np are treated as uncorrelated with those in Nn in our study. The production of deuterons will be “saturated” in events with large Nn–those neutrons would not get “equal” share of protons to form deuterons. Consequently, Nd/Np is smaller for a larger Δnn.

Fig. 3
(Color online) Yield ratios of Nd/Np and Nt/Nd (left) and NtNp/Nd2 (right) as functions of Δnn of Eq. (10), the magnitude of neutron multiplicity fluctuations beyond Poisson. Numbers inside the parentheses indicate the uncertainty to the corresponding last digit. The light nuclei are formed by coalescence from a system of average N¯p=20 protons and N¯n=20 neutrons at T=150 MeV (Eq. (9)) randomly distributed within a cylinder of 10 fm radius and 10 fm length
pic

The right panel of Fig. 3 shows the NtNp/Nd2 ratio as a function of Δnn. The NtNp/Nd2 clearly increases with Δnn. The slope was approximately 0.477. were significantly nonzero, and the intercept was approximately 0.439. These values are approximately equal to the expected degeneracy factor of 4/9. The slight deviations may be due to the different size parameters of the deuterons and triton as the Wigner function parameters differ.

As previously mentioned, the measured nonmonotonic feature of the NtNp/Nd2 ratio as a function of the beam energy sNN by STAR appears to depend on the considered pT/A range [29], which is stronger for the extrapolated yield ratio than for those with limited pT/A acceptance. Therefore, it is important to investigate whether a nonmonotonic extrapolation factor can result from trivial physics, such as nucleon spectral shape changes as a function of the beam energy. To this end, we first examined the NtNp/Nd2 as a function of pT/A in our toy model simulation. This is shown in Fig. 4 left panel shows the various θ values. In the extreme case of coalescence with an identical momentum, the NtNp/Nd2 was independent of pT. The falling and rising characteristics of the shape are determined by the pT-distribution in Eq. (9), and the coalescence parameters of the light nuclei. The shapes are similar for various fluctuation magnitudes, and the overall ratio increases with θ as expected. The right panel of Fig. 4 shows extrapolation factor f as a function of Δnn. (Note that the total yields in simulation are known of course, not from extrapolation, but we keep use of the term “extrapolation factor”.) Two fiducial ranges are depicted: pT/A=0.1–0.2 and 0.3–0.4 GeV/c. The extrapolation factor f varies with Δnn; however, the variation is relatively small, less than 1% for the pT/A=0.3–0.4 GeV/c range for θ=2, which is the fluctuation magnitude of a factor of 2 of that of Poisson. This implies that even when there are enhanced fluctuations in a local region in the beam energy sNN, the extrapolation factor remains approximately the same and does not cause a change in the shape of the NtNp/Nd2 ratio vs. sNN from the limited pT range to the extrapolated full pT range. However, we examined only relatively narrow pT/A ranges at small pT/A values because of statistical considerations. The fiducial pT/A ranges in the STAR analysis are relatively large and wide. Therefore, no direct comparisons can be made between our study and the STAR results thus far.

Fig. 4
(Color online)(Left) The NtNp/Nd2 ratio as functions of pT/A. (Right) The NtNp/Nd2 extrapolation factor f from pT/A=0.1–0.2 GeV/c and 0.3–0.4 GeV/c as a function of Δnn. The light nuclei are formed by coalescence from a system of average N¯p=20 protons and N¯n=20 neutrons at T=150 MeV (Eq. (9)) randomly distributed within a cylinder of 10 fm radius and 10 fm length
pic
5

More Realistic Simulations

In heavy-ion collisions, pT distributions can be significantly altered by the collective radial flow. The shape of the pT distribution affects the NtNp/Nd2 ratio as a function of pT/A. To investigate the effect of the nucleon pT spectral shape on the extrapolation factor of the NtNp/Nd2 ratio, we first vary the temperature parameter T in Eq. (9). Figure 5 shows the T-dependence of the extrapolation factotr f for pT/A=0.3–0.4 GeV/c, with a given neutron fluctuation of Δnn=0.1 (i.e., θ=3). As expected, the f value varies with T as expected, and the variation is monotonic. Thus, the T-dependent pT spectral shape does not cause an artificial non-monotonic NtNp/Nd2 enhancement from limited to full pT acceptance in any given localized T range, and correspondingly, the beam energy range.

Fig. 5
(Color online) The NtNp/Nd2 “extrapolation factor” for the pT/A=0.3–0.4 GeV/c acceptance as a function of T, with θ=3 or Δnn=0.1. The light nuclei are formed by coalescence from a system of average N¯p=20 protons and N¯n=20 neutrons randomly distributed within a cylinder of 10 fm radius and 10 fm length
pic

The collective radial flow in heavy ion collisions creates a correlation between the momentum of a particle and its freeze-out radial position. This correlation is not present in the above simulations, but can be important for coalescence. To fully comprehend the implications of STAR data [29], we performed simulations using more realistic kinematic distributions. We obtained freeze-out information from the measured data [32, 33] as described below. The other simulation details are the same as those described in Sect. 3.

The charged hadron mulitplicity Nch, the inclusive proton multiplicity density dNp/dy, the chemical freeze-out temperature Tchem and volume (Vchem4π3Rsphere3), the kinematic freeze-out temperature Tkin and the average collective radial flow velocity β in the Blast-Wave parameterization have been reported for the STAR experiment using various beam energies [32, 33]. We concentrate only on the central 0-5% collisions and parameterize the freeze-out quantities as functions of Nch. The parameters have fit uncertainties; however, because we were only interested in the input values in our simulation to study light nuclei coalescence, we simply used the parameterized values.

• The measured dNp/dy is shown in Fig. 6 (upper left) as a function of Nch. The measured protons originate from two sources: transport protons whose abundance decreases with energy, and produced protons whose abundance increases with energy. We thus parameterize dNp/dy by combination of two functions, one decreasing and the other increasing with Nch. The parameterization is given by dNp/dy=318eNch/148+68Nch1.4, as superimposed.

Fig. 6
(Color online) The measured proton multiplicity density dNp/dy (upper left), the chemical freeze-out temperature Tchem (upper right), the kinetic freeze-out temperature Tkin (lower left), and the average radial velocity β (lower right) as functions of the charged hadron multiplicity density dNch/dy [32, 33]. The parameterizations to the data are superimposed and used as inputs to our simulations
pic

• The chemical freeze-out temperature Tchem is shown in Fig. 6 (upper right) as a function of Nch. The Tchem increases with Nch and then saturates, so we parameterize it by Tchem=167(1eNch/144) MeV.

• The effective sphere radius for the chemical freeze-out volume appears linear in Nch1/3, so we parameterize it as Rsphere=3.1+0.42Nch1/3. We allow the intercept to be a free parameter because the fit is otherwise not as good.

• The kinetic freeze-out temperature Tkin is shown in Fig. 6 (lower left) as a function of Nch. It is found to decrease with Nch, and parameterized by Tkin=142eNch/1827 MeV.

• The average radial flow velocity is shown in Fig. 6 (lower right) as a function of Nch. It increases with Nch and appears to saturate at large Nch, so we parameterize it as β=0.6(1eNch/241).

The information on the nucleons to be input into the coalescence model is that of kinetic freeze-out. The kinetic freeze-out volume Vkin is obtained by assuming the system expands adiabatically, so Vkin/Vchem=(Tchem/Tkin)3/2. In our simulation, we assumed that the collision zone was a cylinder with a radius R=7 fm, and the length of the cylinder was determined by Vkin/(πR2). Protons and neutrons were positioned uniformly and randomly inside the cylinder. The mean numbers of protons and neutrons were assumed to be N¯p=N¯n=2×dNp/dy over two units of speed. Protons and neutrons were first generated according to the thermal distribution at Tkin. dN/dpTpTmTemT/T. (11) The generated protons and neutrons are boosted radially with a boost velocity dependent on the radial position of the nucleon (r) within the cylinder: β=βs(r/R)n, (12) where the surface velocity is βs=(1+n/2)β. In this study, the parameter n for all the collision energies was set to 1.

Figure 7 shows the calculated pT/A spectra of triton as an example for selected beam energies for θ=1 and θ=10. The pT/A spectra had the same shape for θ=1 and θ=10 for a given beam energy. The right panel shows the yield fractions of the protons (red), deuterons (blue), triton (green) in the limited pT range of 0.5<pT/A<1.0 GeV/c. It is interesting to note that while the proton fiducial yield fraction steadily decreases with the beam energy, as expected from the flattening of the proton pT spectra, the deuteron and triton fiducial yield fractions first increase with the energy and then decrease. This is because the yields in the low pT/A<0.5 GeV/c region are disproportionately more dominant at lower energies for the coalesced light nuclei, and more so for triton. However, no differences were observed in the yield fractions for θ=1 and θ=10. This implies that the yield extrapolations are the same for different fluctuation magnitudes. This, in turn, suggests that the effect of extrapolation on the NtNp/Nd2 ratio is the same, independent of the fluctuation magnitude, as illustrated below.

Fig. 7
(Color Online) (Left panel) The pT/A spectra of triton for different beam energies. Two values of θ are shown: θ=1 for Poisson (hollow markers) and θ=10 for enhanced neutron fluctuations (filled markers). (Right panel) The yield fractions of proton (red), deuteron (blue), triton (green) within 0.5<pT/A<1.0 GeV/c as functions of sNN with θ=1 (open marker) and θ=10 (filled marker)
pic

Figure 8 left panel shows the NtNp/Nd2 calculated using our coalescence model as a function of sNN. The ratios are shown for the entire pT range and for the two limited pT/A ranges measured by STAR: 0.4<pT/A<1.2 GeV/c and 0.5<pT/A<1.0 GeV/c. The results for two values of θ are depicted: θ=1 corresponds to Poisson fluctuations in the neutron multiplicity, and θ=10 corresponds to enhanced fluctuations with a magnitude corresponding to ~10% increase in the measured NtNp/Nd2 ratio by STAR (cf Eq. (10), where a typical N¯n60100, as shown in Fig. 6). Similar to the STAR measurements [29], the calculated NtNp/Nd2 ratios show weak energy dependence. The calculated NtNp/Nd2 values were lower than the measured values.

Fig. 8
(Color online) (Left panel) The NtNp/Nd2 ratio as functions of sNN for the entire pT range (red) and for two limited pT/A ranges (blue and green). Two values of θ are shown: θ=1 for Poisson (hollow markers) and θ=10 for enhanced neutron fluctuations (filled markers). (Right panel) The NtNp/Nd2 extrapolation factor for the pT/A=0.4–1.2 GeV/c (blue) and 0.5–1.0 GeV/c (green) acceptance as functions of sNN, with θ=1 (hollow markers) and 10 (solid markers)
pic

The right panel of Fig. 8 show the extrapolation factors f of the NtNp/Nd2 from pT/A=0.4–1.2 GeV/c and 0.5–1.0 GeV/c, respectively. The extrapolation increases monotonically with an increase sNN and is smooth. For a given pT/A range, the f values were almost identical for the two θ values; thus, we is so for all θ values we have simulated. The chosen θ=10 corresponds to Δnn = 0.1 for N¯n=90 only, which corresponds to different values of Δnn at various beam energies. We checked our results for f using a fixed Δnn = 0.1 value at all beam energies studied. No qualitative differences were observed between groups. Our results suggest that pT-extrapolation alone does not create, enhance, or reduce a local peak in the NtNp/Nd2 ratio as a function of sNN. In other words, the fluctuation increase in the NtNp/Nd2 ratio as a function of sNN~20–30 GeV region implied by the STAR-extrapolated NtNp/Nd2 ratio [29] (such as a jump from the hollow red points to the solid red points at sNN=2030 GeV in our simulation) should also be present in the ratios in the limited fiducial pT/A regions (a jump from the hollow blue/green points to the solid blue/green points). Therefore, we postulate that the local peak in the extrapolated NtNp/Nd2 in the STAR measurements [29] is unlikely to be caused by the physics of coalescence.

6

Summary

We simulated light nuclei production from a system of nucleons with varying magnitudes of neutron multiplicity fluctuations, Δnn. It was found that the light nuclei yield ratio NtNp/Nd2 increased linearly with Δnn, confirming the findings of Ref. [26, 27]. We have further investigated the effect of finite acceptance by studying the “extrapolation factor” f for the NtNp/Nd2 ratio as a function of Δnn and the pT spectra parameter T. The f value is found to be monotonic as a function of T and Δnn; the variations in the latter were relatively small.

We also conducted a coalescence model study with realistic kinematic distributions of nucleons using measured freeze-out parameters, kinetic freeze-out temperature, and collective radial flow velocity. The NtNp/Nd2 ratio was calculated using the coalescence model as a function of beam energy, and a weak beam energy dependence was found, similar to the experimental data. The extrapolation factor f was found to be smooth and monotonic in beam energy and independent of the magnitude of the neutron density fluctuations. We conclude that the extrapolation of the NtNp/Nd2 ratio in pT does not create, enhance, or reduce a local peak in the ratio of the beam energy. Therefore, our study suggests that the measured enhancement of the pT-extrapolated NtNp/Nd2 [29] is unlikely to be caused by coalescence.

Our study provides a necessary benchmark for light nuclei ratios as a probe for nucleon fluctuations, an important observation in the search for the critical point of nuclear matter. In the present study, we assumed that the fluctuations in the number of protons and neutrons were independent. One may implement varying degrees of correlation between these fluctuations and study their effects on NtNp/Nd2. In addition, the effects of clustering or clumping [34-36], out-of-equilibrium effects [34] and feedback from excited states [37, 38] have not been considered. We leave these studies for future work.

References
1. Y. Aoki, G. Endrodi, Z. Fodor, et al.,

The Order of the quantum chromodynamics transition predicted by the standard model of particle physics

. Nature 443, 675678 (2006). https://doi.org/10.1038/nature05120
Baidu ScholarGoogle Scholar
2. A.M. Halasz, A.D. Jackson, R.E. Shrock, et al.,

On the phase diagram of QCD

. Phys. Rev. D 58, 096007 (1998). https://doi.org/10.1103/PhysRevD.58.096007
Baidu ScholarGoogle Scholar
3. M.A. Stephanov, K. Rajagopal, E.V. Shuryak,

Signatures of the tricritical point in QCD

. Phys. Rev. Lett. 81, 48164819 (1998). https://doi.org/10.1103/PhysRevLett.81.4816
Baidu ScholarGoogle Scholar
4. M.A. Stephanov,

QCD phase diagram and the critical point

. Prog. Theor. Phys. Suppl. 153, 139156 (2004). https://doi.org/10.1142/S0217751X05027965
Baidu ScholarGoogle Scholar
5. K. Fukushima, T. Hatsuda,

The phase diagram of dense QCD

. Rept. Prog. Phys. 74, 014001 (2011). https://doi.org/10.1088/0034-4885/74/1/014001
Baidu ScholarGoogle Scholar
6. Y. WU, X. LI, L. CHEN, et al.,

Several problems in determining the qcd phase boundary by relativistic heavy ion collisions

. NUCLEAR TECHNIQUES 46, 040006 (2023).
Baidu ScholarGoogle Scholar
7. M.A. Stephanov,

On the sign of kurtosis near the QCD critical point

. Phys. Rev. Lett. 107, 052301 (2011). https://doi.org/10.1103/PhysRevLett.107.052301
Baidu ScholarGoogle Scholar
8. M.M. Aggarwal, et al., An Experimental Exploration of the QCD Phase Diagram: The Search for the Critical Point and the Onset of De-confinement.
9. L. Adamczyk, et al.,

Energy Dependence of Moments of Net-proton Multiplicity Distributions at RHIC

. Phys. Rev. Lett. 112, 032302 (2014). https://doi.org/10.1103/PhysRevLett.112.032302
Baidu ScholarGoogle Scholar
10. X. Luo,

Exploring the QCD Phase Structure with Beam Energy Scan in Heavy-ion Collisions

. Nucl. Phys. A 956, 7582 (2016). https://doi.org/10.1016/j.nuclphysa.2016.03.025
Baidu ScholarGoogle Scholar
11. A. Bzdak, S. Esumi, V. Koch, et al.,

Mapping the Phases of Quantum Chromodynamics with Beam Energy Scan

. Phys. Rept. 853, 187 (2020). https://doi.org/10.1016/j.physrep.2020.01.005
Baidu ScholarGoogle Scholar
12. S. Acharya, et al.,

Global baryon number conservation encoded in net-proton fluctuations measured in Pb-Pb collisions at sNN = 2.76 TeV

. Phys. Lett. B 807, 135564 (2020). https://doi.org/10.1016/j.physletb.2020.135564
Baidu ScholarGoogle Scholar
13. J. Adamczewski-Musch, et al.,

Proton-number fluctuations in sNN =2.4 GeV Au + Au collisions studied with the High-Acceptance DiElectron Spectrometer (HADES)

. Phys. Rev. C 102, 024914 (2020). https://doi.org/10.1103/PhysRevC.102.024914
Baidu ScholarGoogle Scholar
14. J. Adam, et al.,

Nonmonotonic Energy Dependence of Net-Proton Number Fluctuations

. Phys. Rev. Lett. 126, 092301 (2021). https://doi.org/10.1103/PhysRevLett.126.092301
Baidu ScholarGoogle Scholar
15. M. Abdallah, et al.,

Cumulants and correlation functions of net-proton, proton, and antiproton multiplicity distributions in Au+Au collisions at energies available at the BNL Relativistic Heavy Ion Collider

. Phys. Rev. C 104, 024902 (2021). https://doi.org/10.1103/PhysRevC.104.024902
Baidu ScholarGoogle Scholar
16. M.S. Abdallah, et al.,

Measurements of Proton High Order Cumulants in sNN = 3 GeV Au+Au Collisions and Implications for the QCD Critical Point

. Phys. Rev. Lett. 128, 202303 (2022). https://doi.org/10.1103/PhysRevLett.128.202303
Baidu ScholarGoogle Scholar
17. Y. ZHANG, D. ZHANG, X. LUO,

Experimental study of the qcd phase diagram in relativistic heavy-ion collisions

. NUCLEAR TECHNIQUES 46, 040001 (2023).
Baidu ScholarGoogle Scholar
18. S.T. Butler, C.A. Pearson,

Deuterons from High-Energy Proton Bombardment of Matter

. Phys. Rev. 129, 836 (1963). https://doi.org/10.1103/PhysRev.129.836
Baidu ScholarGoogle Scholar
19. H. Sato, K. Yazaki,

On the coalescence model for high-energy nuclear reactions

. Phys. Lett. B 98, 153157 (1981). https://doi.org/10.1016/0370-2693(81)90976-X
Baidu ScholarGoogle Scholar
20. L.P. Csernai, J.I. Kapusta,

Entropy and Cluster Production in Nuclear Collisions

. Phys. Rept. 131, 223318 (1986). https://doi.org/10.1016/0370-1573(86)90031-1
Baidu ScholarGoogle Scholar
21. C.B. Dover, U.W. Heinz, E. Schnedermann, et al.,

Relativistic coalescence model for high-energy nuclear collisions

. Phys. Rev. C 44, 1636 (1991). https://doi.org/10.1103/PhysRevC.44.1636
Baidu ScholarGoogle Scholar
22. R. Scheibl, U.W. Heinz,

Coalescence and flow in ultrarelativistic heavy ion collisions

. Phys. Rev. C 59, 15851602 (1999). https://doi.org/10.1103/PhysRevC.59.1585
Baidu ScholarGoogle Scholar
23. L.W. Chen, C.M. Ko, B.A. Li,

Light clusters production as a probe to the nuclear symmetry energy

. Phys. Rev. C 68, 017601 (2003). https://doi.org/10.1103/PhysRevC.68.017601
Baidu ScholarGoogle Scholar
24. Y. Oh, Z.W. Lin, C.M. Ko,

Deuteron production and elliptic flow in relativistic heavy ion collisions

. Phys. Rev. C 80, 064902 (2009). https://doi.org/10.1103/PhysRevC.80.064902
Baidu ScholarGoogle Scholar
25. K. SUN, L. CHEN, K.C. Ming, et al.,

Light nuclei production and qcd phase transition in heavy-ion collisions

. NUCLEAR TECHNIQUES 46, 040012 (2023).
Baidu ScholarGoogle Scholar
26. K.J. Sun, L.W. Chen, C.M. Ko, et al.,

Probing QCD critical fluctuations from light nuclei production in relativistic heavy-ion collisions

. Phys. Lett. 774, 103107 (2017). https://doi.org/10.1016/j.physletb.2017.09.056
Baidu ScholarGoogle Scholar
27. K.J. Sun, L.W. Chen, C.M. Ko, et al.,

Light nuclei production as a probe of the QCD phase diagram

. Phys. Lett. B 781, 499504 (2018). https://doi.org/10.1016/j.physletb.2018.04.035
Baidu ScholarGoogle Scholar
28. K.J. Sun, L.W. Chen,

Analytical coalescence formula for particle production in relativistic heavy-ion collisions

. Phys. Rev. C 95, 044905 (2017). https://doi.org/10.1103/PhysRevC.95.044905
Baidu ScholarGoogle Scholar
29. M. Abdulhamid, et al.,

Beam Energy Dependence of Triton Production and Yield Ratio (Nt×Np/Nd2) in Au+Au Collisions at RHIC

. Phys. Rev. Lett. 130, 202301 (2023). https://doi.org/10.1103/PhysRevLett.130.202301
Baidu ScholarGoogle Scholar
30. G. Ropke,

Light nuclei quasiparticle energy shift in hot and dense nuclear matter

. Phys. Rev. C 79, 014002 (2009). https://doi.org/10.1103/PhysRevC.79.014002
Baidu ScholarGoogle Scholar
31. S. Wu, K. Murase, S. Tang, et al.,

Examination of background effects on the light-nuclei yield ratio in relativistic heavy-ion collisions

. Phys. Rev. C 106, 034905 (2022). https://doi.org/10.1103/PhysRevC.106.034905
Baidu ScholarGoogle Scholar
32. L. Adamczyk, et al.,

Bulk Properties of the Medium Produced in Relativistic Heavy-Ion Collisions from the Beam Energy Scan Program

. Phys. Rev. C 96, 044904 (2017). https://doi.org/10.1103/PhysRevC.96.044904
Baidu ScholarGoogle Scholar
33. B.I. Abelev, et al.,

Systematic Measurements of Identified Particle Spectra in pp,d+ Au and Au+Au Collisions from STAR

. Phys. Rev. C 79, 034909 (2009). https://doi.org/10.1103/PhysRevC.79.034909
Baidu ScholarGoogle Scholar
34. J. Steinheimer, J. Randrup,

Spinodal amplification of density fluctuations in fluid-dynamical simulations of relativistic nuclear collisions

. Phys. Rev. Lett. 109, 212301 (2012). https://doi.org/10.1103/PhysRevLett.109.212301
Baidu ScholarGoogle Scholar
35. E. Shuryak, J.M. Torres-Rincon,

Baryon clustering at the critical line and near the hypothetical critical point in heavy-ion collisions

. Phys. Rev. C 100, 024903 (2019). https://doi.org/10.1103/PhysRevC.100.024903
Baidu ScholarGoogle Scholar
36. K.J. Sun, W.H. Zhou, L.W. Chen, et al.,

Spinodal Enhancement of Light Nuclei Yield Ratio in Relativistic Heavy Ion Collisions

. https://doi.org/10.48550/arXiv.2205.11010
Baidu ScholarGoogle Scholar
37. E. Shuryak, J.M. Torres-Rincon,

Baryon preclustering at the freeze-out of heavy-ion collisions and light-nuclei production

. Phys. Rev. C 101, 034914 (2020). https://doi.org/10.1103/PhysRevC.101.034914
Baidu ScholarGoogle Scholar
38. V. Vovchenko, B. Dönigus, B. Kardan, et al.,

Feeddown contributions from unstable nuclei in relativistic heavy-ion collisions

. Phys. Lett. B 809, 135746 (2020). https://doi.org/10.1016/j.physletb.2020.135746
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.