Bayesian analysis of nuclear equation of state at high Baryon density

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Bayesian analysis of nuclear equation of state at high Baryon density

Long-Gang Pang
Xin-Nian Wang
Nuclear Science and TechniquesVol.34, No.12Article number 194Published in print Dec 2023Available online 09 Dec 2023
258014

Video Abstract

Bayesian analysis was employed to constrain the Equation of State (EoS) of nuclear matter with a baryon density of up to six times the nuclear saturation density, using data from heavy-ion collisions at beam energies sNN=210 GeV. The resulting EoS excellently agrees with that constrained by astrophysical observations.

The nuclear EoS describes the relationship between different thermodynamic variables in the bulk nuclear matter. The relationship between the pressure P and energy density ϵ determines the square of the speed of sound in the nuclear matter cs2=dP/dϵ. The value of cs2 is approximately 1/3 in the ideal gas limit of the quark-gluon plasma phase (QGP) (or any system with conformal symmetry) and about 1/6 in the hadron resonance gas (HRG) phase of nuclear matter with zero net baryon density. This corresponds to the state of matter existing in the early universe and produced in heavy-ion collisions at the top RHIC and LHC energies, where nearly equal amounts of matter and antimatter are created. Lattice QCD calculations predict that the transition between the QGP and HRG of nuclear matter is a smooth crossover. The nuclear EoS also depends on the net baryon density, such as that produced in low-energy heavy-ion collisions with strong baryon stopping or dense nuclear matter inside neutron stars. Nuclear matter at a finite baryon density can have a different type of phase transition, possibly first order, during which the speed of sound approaches cs20.

Nuclear EoS is essential for describing the dynamic evolution of dense and hot nuclear matter created in heavy-ion collisions because the primary driving force for fluid expansion is the pressure gradient. It is also a vital input for determining the mass-radius relation of neutron stars, as a soft EoS cannot support immense gravitational forces. Nuclear EoS encodes QCD phase transitions, which can be used to identify neutrons and quark stars. Furthermore, the nuclear EoS reveals the emergent properties of numerous nucleons and provides constraints on the nucleon-nucleon interaction, which is crucial for few-nucleon systems such as atomic nuclei.

In a recent publication in Physical Review Letters [1], Manjunath, Steinheimer, Zhou and Stöcker from Goethe University Frankfurt and GSI reported their study on Bayesian analysis to constrain the nuclear EoS in dense regions using data from intermediate-energy heavy-ion collisions. They constructed a density-dependent mean-field potential in UrQMD to simulate heavy-ion collisions at 2-10 GeV. Using published data from heavy-ion collisions, Bayesian analysis was then used to constrain the interaction potential and nuclear EoS. The extracted EoS for nucleon densities ranging from 2n0 to 4n0 agrees well with that given by astrophysical observations; the EoS beyond 3n0 requires more accurate data with higher statistics from the beam energy scan program of STAR-FXT at RHIC, the upcoming CBM experiment at FAIR, and future experiments at HIAF and NICA.

No first-principles calculations are available for the nuclear Equation of State (EoS) at high baryon densities. Although lattice QCD calculations provide the QCD EoS at the vanishing baryon chemical potentials, they cannot accurately predict the EoS at high net baryon densities because of the notorious fermionic sign problem [2]. To investigate the nuclear EoS at high baryon densities and possible critical endpoint between the two phases [3, 4], physicists have turned to a data-driven approach.

The beam energy scan (BES) program at RHIC was designed to accumulate data on heavy-ion collisions at various beam energies with different baryon stopping powers and net baryon chemical potentials in the central rapidity region of the final state. Theoretical approaches, such as the QCD chiral effective field theory (EFT), transport model simulations, and relativistic hydrodynamics, have been employed to extract the nuclear EoS from experimental data on heavy-ion collisions, neutron stars, gravitational waves, and nuclear incompressibility. Fast-moving objects traveling through a medium at speeds greater than the speed of sound create Mach cones. Recent studies also explored the potential use of jet-induced 3D Mach cones to determine the nuclear equation of state [5]. For comparisons between data and theoretical calculations, advanced machine learning methods such as Bayesian analyses [6-10] and deep learning [4, 11-17] are used. For further details, please refer to the following reviews [18-23].

To describe the EoS of dense nuclear matter above two times of the nuclear saturation density 2n0, the authors of Ref. [1] use a 7-th order polynomial function to parameterize the density dependence of baryon potential energy V(nB(r)). The seven coefficients in the polynomial function form the parameter space θ=[θ1,θ2,,θ7] to be extracted using Bayesian analysis. The dataset on elliptic flow v2 and the mean transverse kinetic energy mTm0 of protons at the mid-rapidity of Au+Au collisions at s=210 GeV, referred to as D=[v2,mTm0], were used to constrain the EoS. The goal of their Bayesian analysis was to determine the posterior distributions of the parameters in the polynomial function, given the dataset P(θ|D)P(θ)P(D|θ) The posterior distribution was not normalized. The Markov Chain Monte Carlo (MCMC) method is typically employed to obtain samples from this unnormalized distribution by performing random walks in the parameter space. For instance, the Metropolis-Hastings (MH) algorithm begins at a randomly chosen location θstep0 in the parameter space and proposes the next step θstep1=θstep0+δθ. The algorithm then determines whether to accept this new position based on posterior distribution P(θ | D). If the posterior distribution exhibits a higher probability density at the proposed position than at the current position, θstep1 is accepted as a new sample. Otherwise, it is accepted with a probability ratio between the new and old positions. If θstep1 is rejected, the old position θstep0 is duplicated as a new sample. This process was implemented in PyMC and EMCEE. This study uses PyMC4.0 to sample θ from the posterior distribution P(θ | D). Density estimation provides the distribution of θ, from which one can identify the mean values of θ or parameter combinations that maximize P(θ | D), known as the Maximum A Posteriori (MAP).

In this study, the key equation that relates the potential V(nB(r)) to the equation of state (pressure) is given by [24] P(nB)=Pid(nB)+0nBn'U(n')n'dn' where Pid is the pressure of the ideal fermionic gas and U(nB)=(nBV(nB))nB. Once the pressure is known, the square of the speed of sound cs2 in dense nuclear matter can be computed using its derivatives.

As shown in Fig. 1, the extracted cs2 values using 13 and 15 data points from heavy-ion collisions are compared with those obtained from constraints by Binary Neutron Star Mergers (grey band) [25] and the Chiral Mean Field (CMF) model calculations (purple curve) [24]. The cs2 values extracted using the Bayesian analysis agreed with the Binary Neutron Star Merger constraints when 15 data points were used. However, reducing the number of data points to 13 leads to a rapid change in cs2, indicating that the EoS is sensitive to the observables selected in the Bayesian analysis. Therefore, higher statistical data from experimental measurements are required to constrain nuclear EoS in dense regions better.

Fig. 1
The square of the speed of sound cs2 at the density of 2n0 - 6n0, where n0 is the nuclear saturation density, is extracted from comparisons between experimental data and UrQMD simulations of heavy-ion collisions at the beam energies of 2-10 GeV in Ref. [1]. The Bayesian analysis was employed to constrain the EoS, and the results agree with neutron star merger observations. Notably, the results are sensitive to the number of data points used in the constraining process.
pic

The performance of the Bayesian analysis was verified using closure tests. In these tests, a random potential function was incorporated into UrQMD to compute the physical observables and Bayesian analysis was employed to obtain the posterior distribution of θ in the potential function. The ground-truth potential function is centered around the resulting V(nB(r)). Closure tests are crucial in Bayesian analysis because many scientific problems do not have well-defined inverse solutions. For example, if two or more potential functions yield the same v2 and mTm0, then it is impossible to retrieve the used V(nB(r)). This does not imply that there is no way to distinguish between different EoS but suggests that additional physical observables may be necessary. In this study, the closure tests indicated that the physical observables selected by the authors were sensitive to changes in the EoS.

Constraining the EoS at the high net baryon density is essential for understanding the dynamic evolution of dense nuclear matter during heavy-ion collisions. This knowledge is also crucial for providing key inputs to obtain the mass-radius relationship of neutron stars and interpreting neutron star merger observations. Heavy-ion collisions at various intermediate and low beam energies offer a unique opportunity to study the nuclear EoS under extreme conditions in a laboratory setting [26, 27]. Furthermore, this research can help constrain the fundamental nucleon-nucleon interactions in many-body quantum systems. The incorporation of advanced statistical tools and machine learning methods into data theory comparisons in nuclear physics will accelerate the advancement of this frontier.

References
1. M. Omana Kuttan, J. Steinheimer, K. Zhou et al.,

The QCD EoS of dense nuclear matter from Bayesian analysis of heavy ion collision data

. Phys. Rev. Lett., (in press) (2023); [arXiv:2211.11670 [nucl-th]].
Baidu ScholarGoogle Scholar
2. J.N. Guenther, R. Bellwied, S. Borsanyi et al.,

The QCD equation of state at finite density from analytical continuation

. Nucl. Phys. A 967, 720-723 (2017). https://doi.org/10.1016/j.nuclphysa.2017.05.044
Baidu ScholarGoogle Scholar
3. C.M. Ko,

Searching for QCD critical point with light nuclei

. Nucl. Sci. Tech. 34, 80 (2023). https://doi.org/10.1007/s41365-023-01231-1
Baidu ScholarGoogle Scholar
4. Y. Huang, L.G. Pang, X. Luo et al.,

Probing criticality with deep learning in relativistic heavy-ion collisions

. Phys. Lett. B 827, 137001 (2022). https://doi.org/10.1016/j.physletb.2022.137001
Baidu ScholarGoogle Scholar
5. Z. Yang, T. Luo, W. Chen et al.,

3D structure of jet-induced diffusion wake in an expanding quark-gluon plasma

. Phys. Rev. Lett. 130, 052301 (2023). https://doi.org/10.1103/PhysRevLett.130.052301
Baidu ScholarGoogle Scholar
6. S. Pratt, E. Sangaline, P. Sorensen et al.,

Constraining the equation of state of superhadronic matter from heavy-ion collisions

. Phys. Rev. Lett. 114, 202301 (2015). https://doi.org/10.1103/PhysRevLett.114.202301
Baidu ScholarGoogle Scholar
7. Z. Miao, J.L. Jiang, A. Li et al.,

Bayesian inference of strange star equation of state using the GW170817 and GW190425 data

. Astrophys. J. Lett. 917, L22 (2021). https://doi.org/10.3847/2041-8213/ac194d
Baidu ScholarGoogle Scholar
8. Z. Zhang, L.W. Chen,

Bayesian inference of the symmetry energy and the neutron skin in Ca48 and Pb208 from CREX and PREX-2

. Phys. Rev. C 108, 024317 (2023). https://doi.org/10.1103/PhysRevC.108.024317
Baidu ScholarGoogle Scholar
9. W.F. Li, X.Y. Zhang, Z.M. Niu,

Studies of nuclear β-decay half-lives with Bayesian neural network approach

. Nucl. Tech. (in Chinese) 46, 080013 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080013
Baidu ScholarGoogle Scholar
10. W. Xie, Z. Ma, J. Gao,

Bayesian inference of the crust-core transition density via the neutron-star radius and neutron-skin thickness data

. Nucl. Sci. Tech. 34, 91 (2023). https://doi.org/10.1007/s41365-023-01239-7
Baidu ScholarGoogle Scholar
11. L.G. Pang, K. Zhou, N. Su et al.,

An equation-of-state-meter of quantum chromodynamics transition from deep learning

. Nat. Commun. 9, 210 (2018). https://doi.org/10.1038/s41467-017-02726-3
Baidu ScholarGoogle Scholar
12. Y. Wang, F. Li, Q. Li et al.,

Finding signatures of the nuclear symmetry energy in heavy-ion collisions with deep learning

. Phys. Lett. B 822, 136669 (2021). https://doi.org/10.1016/j.physletb.2021.136669
Baidu ScholarGoogle Scholar
13. X. Li, R. Guo, Y. Zhou et al.,

Machine learning phase transitions of the three-dimensional Ising universality class

. Chin. Phys. C 47, 034101 (2023). https://doi.org/10.1088/1674-1137/aca5f5
Baidu ScholarGoogle Scholar
14. S. Soma, L. Wang, S. Shi et al.,

Reconstructing the neutron star equation of state from observational data via automatic differentiation

. Phys. Rev. D 107, 083028 (2023). https://doi.org/10.1103/PhysRevD.107.083028
Baidu ScholarGoogle Scholar
15. Z. Gao, Q. Li,

Studies on several problems in nuclear physics by using machine learning

. Nucl. Tech. (in Chinese) 46, 080009 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080009
Baidu ScholarGoogle Scholar
16. Z. Zhang, R. Ma, J. Hu et al.,

Approach the Gell–Mann–Okubo formula with machine learning

. Chin. Phys. Lett. 39, 111201 (2022). https://doi.org/10.1088/0256-307X/39/11/111201
Baidu ScholarGoogle Scholar
17. R. Wang, Y.G. Ma, R. Wada et al.,

Nuclear liquid–gas phase transition with machine learning

. Phys. Rev. Res. 2, 043202 (2020). https://doi.org/10.1103/PhysRevResearch.2.043202
Baidu ScholarGoogle Scholar
18. A. Boehnlein, M. Diefenthaler, N. Sato et al.,

Colloquium: machine learning in nuclear physics

. Rev. Mod. Phys. 94, 031003 (2022). https://doi.org/10.1103/RevModPhys.94.031003
Baidu ScholarGoogle Scholar
19. W.B. He, Q.F. Li, Y.G. Ma et al.,

Machine learning in nuclear physics at low and intermediate energies

. Sci. China Phys. Mech. Astron. 66, 282001 (2023). https://doi.org/10.1007/s11433-023-2116-0
Baidu ScholarGoogle Scholar
20. W.B. He, Y.G. Ma, L.G. Pang et al.,

High-energy nuclear physics meets machine learning

. Nucl. Sci. Tech. 34, 88 (2023). https://doi.org/10.1007/s41365-023-01233-z
Baidu ScholarGoogle Scholar
21. K. Zhou, L. Wang, L.G. Pang et al.,

Exploring QCD matter in extreme conditions with Machine Learning

. arXiv: 2303.15136 [hep-ph]
Baidu ScholarGoogle Scholar
22. Y.J. Wang, Q.F. Li,

Machine learning transforms the inference of the nuclear equation of state

. Front. Phys. 18, 64402 (2023). https://doi.org/10.1007/s11467-023-1313-3
Baidu ScholarGoogle Scholar
23. Y.G. Ma, L.G. Pang, R. Wang et al.,

Phase transition study meets machine learning

. Chin. Phys. Lett. (2023). https://doi.org/10.1088/0256-307X/40/12/122101
Baidu ScholarGoogle Scholar
24. J. Steinheimer, A. Motornenko, A. Sorensen et al.,

The highdensity equation of state in heavy-ion collisions: constraints from proton flow

. Eur. Phys. J. C 82, 911 (2022). https://doi.org/10.1140/epjc/s10052-022-10894-w
Baidu ScholarGoogle Scholar
25. S. Altiparmak, C. Ecker, L. Rezzolla,

On the sound speed in neutron stars

. Astrophys. J. Lett. 939, L34 (2022). https://doi.org/10.3847/2041-8213/ac9b2a
Baidu ScholarGoogle Scholar
26. P.C. Li, Y.J. Wang, Q.F. Li et al.,

Transport model analysis of the pion interferometry in Au + Au collisions at Ebeam = 1.23 GeV/nucleon

. Sci. China Phys. Mech. Astron. 66, 222011 (2023). https://doi.org/10.1007/s11433-022-2026-5
Baidu ScholarGoogle Scholar
27. P.C. Li, J. Steinheimer, T. Reichert et al.,

Effects of a phase transition on two-pion interferometry in heavy ion collisions at sNN = 2.4–7.7 GeV, Science China

. Physics, Mechanics and Astronomy 66, 232011 (2023); https://doi.org/10.1007/s11433-022-2041-8
Baidu ScholarGoogle Scholar