Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

Chapter 3

Dynamical Evolution of Heavy-Ion


Collisions

H. Elfner, J. Y. Jia, Z. W. Lin, Y. Nara, L. G. Pang, C. Shen, S. S. Shi,


M. Stephanov, L. Yan, Y. Yin, and P. F. Zhuang

Relativistic viscous hydrodynamics and transport models are the primary tools to
study the real-time dynamics of relativistic nuclear collisions. At high collision ener-
gies, the hybrid approach that combines hydrodynamics with hadronic transport
achieves quantitative descriptions of many experimental observables across differ-
ent systems. Transport models with both parton and hadron phases such as AMPT and
PHSD (Parton–Hadron–String Dynamics) also describe many experimental observ-
ables including anisotropic flows. In Sect. 3.1, we focus on recent developments on
the initial condition in these frameworks for nuclear collisions at energies relevant

Y. Nara
Akita International University, Yuwa, Akita-city 010-1292, Japan
J. Y. Jia
Department of Chemistry, Stony Brook University, Stony Brook, NY 11794, USA
J. Y. Jia · C. Shen
Brookhaven National Laboratory, Upton, NY 11973, USA
C. Shen
Wayne State University, Detroit, Michigan 48201, USA
L. G. Pang · S. S. Shi
Central China Normal University, Wuhan, China
Z. W. Lin
East Carolina University, Greenville, NC 27858, USA
H. Elfner
GSI Helmholtzzentrum für Schwerionenforschung, Planckstr. 1, 64291 Darmstadt, Germany
Frankfurt Institute for Advanced Studies, Ruth-Moufang-Strasse 1, 60438 Frankfurt am Main,
Germany
Goethe University, Max-von-Laue-Strasse 1, 60438 Frankfurt am Main, Germany

© Science Press 2022 135


X. Luo et al. (eds.), Properties of QCD Matter at High Baryon Density,
https://doi.org/10.1007/978-981-19-4441-3_3
136 H. Elfner et al.

for high baryon densities. In Sect. 3.2, we discuss recent theoretical developments to
describe the system evolution in a baryon-rich environment. Fluctuation dynamics
in the hydrodynamics framework are discussed in Sect. 3.3. Experimental flow mea-
surements and related phenomenological studies are discussed in Sect. 3.4. Finally,
Sect. 3.5 gives a summary of the chapter.

3.1 Initial Condition

In the study of QGP properties at high baryon densities, key variables of the initial
condition include the energy–momentum current and net-baryon density produced
in relativistic nuclear collisions. Their initial three-dimensional (3D) distributions
and the subsequent time evolution determine the trajectory of the events on the
temperature–baryon chemical potential phase diagram. For collision energies such
as those in the Beam Energy Scan program at RHIC [1–4], the event trajectories
relative to the location of the possible QCD critical point [5, 6] could significantly
affect the experimental observables and their sensitivities to the critical point [5].
For heavy-ion collisions at O(10) GeV that can create high baryon densities, one
should consider the finite nuclear thickness along the beam directions because the
Lorentz contraction factor is not very large [7–9]. The finite nuclear thickness will
obviously affect the initial densities [9–11]. Furthermore, it will lead to a significant
time duration of the initial particle and energy production; therefore, one cannot use
a fixed proper time to describe the initial condition for hydrodynamics-based models
but needs to use a dynamical initialization scheme [12]. It is also essential for the
initial condition to properly describe baryon stopping after the initial encounter of
the colliding nuclei since this directly affects the net-baryon density [13–15].

Helmholtz Research Academy Hesse for FAIR (HFHF), GSI Helmholtz Center, Campus
Frankfurt, Max-von-Laue-Strasse 12, 60438 Frankfurt am Main, Germany
L. Yan
Fudan University, Shanghai 200433, China
Y. Yin
Institute of Modern Physics, Chinese Academy of Sciences, Gansu 730000, China
P. F. Zhuang (B)
Tsinghua University, Beijing 100084, China
e-mail: zhuangpf@mail.tsinghua.edu.cn
M. Stephanov
University of Illinois at Chicago, Chicago, IL 60607, USA
3 Dynamical Evolution of Heavy-Ion Collisions 137

3.1.1 Finite Nuclear Thickness



For heavy-ion collisions at lower energies, e.g., at sNN ≤ 50 GeV for central Au+Au
collisions, the thickness of the incoming projectile and target nuclei in the center-of-
mass frame becomes significant due to the modest Lorentz contraction. Therefore,
one needs to consider the finite nuclear thickness in dynamical models of heavy-ion
collisions at high baryon densities. For a central collision of two identical nuclei of
mass number A, it takes time

2R A
dt = (3.1)
sinh ycm

for them to completely cross each other in the center-of-mass frame of the two nuclei
in the hard-sphere approximation. Here, R A is the hard-sphere radius of the nucleus
and ycm is the rapidity of the projectile nucleus. For central Au+Au collisions at

sNN = 50 GeV, for example, dt ≈ 0.5 fm/c is comparable to the typical value of
the parton formation time or thermalization time when one takes R A = 1.12 A1/3 fm.
Therefore, one may expect the effect from finite nuclear thickness to be significant

for central Au+Au collisions at sNN ≤ 50 GeV [9].
Semi-analytical methods [9, 11] have been developed to calculate the effect of
finite nuclear thickness on the initial energy density, which is crucial for determining
the initial temperature (and the net-baryon chemical potential at lower energies) of
the produced QGP. Traditionally, the Bjorken formula [16] has been the standard
semi-analytical tool for estimating the initial energy density in the central rapidity
region at the formation time of the QGP:

1 d ET
! B j (τ ) = . (3.2)
A T τ dy

In the above, A T represents the full transverse area of the overlap volume, and
d E T /dy is the initial rapidity density of the transverse energy at mid-rapidity, which
is often approximated with the experimental d E T /dy value in the final state, and τ
is often taken as the proper formation time of the produced quanta (τ F ). However,
a serious limitation of the Bjorken formula results from the fact that it neglects the
finite thickness of the colliding nuclei along the beam direction z, which shall lead
to a finite duration time and a finite longitudinal width in z, for the initial energy
production. One, therefore, expects that the Bjorken formula may break down when
the crossing time is not small compared to the formation time [17].
With the semi-analytical methods that include the finite nuclear thickness, the
initial energy density !(τ ) averaged over the transverse area of the overlap region
as a function of time, including its maximum value !max , has been calculated. These
studies [9, 11] have yielded the following qualitative conclusions: the initial energy
density after considering the finite nuclear thickness approaches the Bjorken formula
at large formation time τ F and/or high collision energies; at low collision energies,
the initial energy density has a lower maximum, evolves longer, and is less sensitive
138 H. Elfner et al.

Fig. 3.1 The maximum initial energy density for central Au+Au collisions as a function of collision
energy at τ F = 0.1, 0.3, and 0.9 fm/c from the improved semi-analytical method that includes the
finite nuclear thickness and the Bjorken formula. Dot-dashed line represents the finite thickness
results at τ F = 0

to τ F than the Bjorken formula. Figure 3.1 shows the calculated maximum initial
energy density for central Au+Au collisions as a function of the collision energy,
where one can observe certain features of the above conclusions. The three solid
curves from top to bottom represent the !max values for τ F = 0.1, 0.3, and 0.9 fm/c,
respectively; while the three dashed curves represent the corresponding results from
the Bjorken energy density formula. Numerically, it is found that the Bjorken energy
density formula breaks down (i.e., is different by 20% or more from the results that
include the finite nuclear thickness) when dt > τ F , as one may expect. One also sees
that at low energies the maximum energy density increases with the collision energy
much faster than the Bjorken case. In addition, it is found that the maximum energy
density is finite at τ F = 0 at any energy (dot-dashed curve) [11], where in contrast
the Bjorken energy density formula is divergent as τ F → 0.
The solution of the initial energy density [11] is also found to possess an approx-
imate scaling property, which leads to scaling relations such as the following:
√ ! √ "
!max
AA (for τ F , sNN ) = !max Au
AuAu for τ F = (197/A)
1/3
τ F , sNN . (3.3)

This scaling means that the τ F -dependence of !max also gives the A-dependence of
!max . As a result, the weaker τ F -dependence of !max at low energies after considering
the finite nuclear thickness means a slower increase of !max with A. In addition, the

peak initial energy density at τ F = 0 is independent of A but only depends on sNN .
3 Dynamical Evolution of Heavy-Ion Collisions 139

3.1.2 Dynamical Initialization Schemes

For heavy-ion collisions at O(10) GeV collision energy, the projectile’s participant
nucleons have 2–3 fm/c to pass through the target nucleus as illustrated in Fig. 3.2.
Along their trajectories, they collide with the target’s nucleons either coherently or
incoherently at different space–time positions. Therefore, the energy and momentum
from individual nucleon–nucleon collisions are expected to be thermalized at differ-
ent times. The local equilibration has been investigated [18], and at lower energies,
there is a significant part of the system already close enough to equilibrium before
the passing time of the nuclei is reached. In addition, there is a centrality dependence
of the amount of thermalized cells as a function of time. These energy–momentum
currents need to be treated as source terms to continuously feed the macroscopic
hydrodynamic fields during the nuclei overlap. Generally, such an approach is called
the dynamical initialization scheme [8, 19]. It interweaves the initial collision stage
of heavy-ion collisions with relativistic hydrodynamics on a local basis as shown in
Fig. 3.2.
The initial collision geometry can be modeled by the Monte Carlo Glauber model
[8] as well as more sophisticated transport frameworks such as AMPT [20], JAM
(Jet Aa Microscopic transport model) [21], SMASH [22], UrQMD [23, 24], PHSD
[25], and GiBUU (Giessen Boltzmann–Uehling–Uhlenbeck) [26]. For hydrodynamic
models, these initial energy–momentum currents including their spatial and temporal
dependencies are essential inputs for the subsequent hydrodynamical evolution of


Fig. 3.2 A sketch of the space–time structure for relativistic heavy-ion collisions at sNN ∼
O(10) GeV. The dynamical initialization connects the three-dimensional collision geometry with
hydrodynamics. The figure is taken from Ref. [12]
140 H. Elfner et al.

the QGP matter [19, 27–30]. The hydrodynamic equation of motion needs to be
solved together with external source terms,

∂µ T µν = Jsource
ν
, (3.4)
µ
∂µ J = ρsource . (3.5)

For collision energies sNN ≤ 5 GeV, the nuclei overlapping time is comparable to
the lifetime of the fireball. The initial state, hydrodynamics, and hadronic scatterings
cannot be separated. One needs to interface all these phases together in a concurrent
way [30].

3.1.3 The Three-Dimensional Monte Carlo Glauber with


Classical String Deceleration Model

The space–time distribution of energy–momentum currents can be modeled by the


Monte Carlo (MC) Glauber model in 3D, which takes into account event-by-event
fluctuations of nucleons’ positions in the nucleus with the proper Lorentz contraction
at different collision energies [8, 31]. Assuming all the participant nucleons fly along
eikonal trajectories, we can determine individual nucleon–nucleon (NN) collisions’
spatial positions and their collision time. Valence quark energy loss is considered
in every NN collision. The dynamics of how they lose energy are modeled by the
classical string deceleration model, which provides non-trivial correlations between
the particles’ momenta and their space–time rapidities [8, 32]. Energy–momentum
conservation is imposed locally for the produced strings as source terms for hydro-
dynamics [33]. The baryon charges of the incoming nucleons are distributed either
at the string ends or fluctuate to the string junctions [34].
Figure 3.3 shows the space–time distributions of strings produced by the three-
dimensional MC-Glauber model in central Au+Au collisions at 200 and 19.6 GeV

[8]. The left panels show those distributions in the Cartesian coordinates. At sNN =
200 GeV, all the strings lie on an approximate constant proper time surface τ =
0.5 fm/c. At 19.6 GeV, however, there is a significant spread of the string distribution
in proper time because a Au+Au collision has a non-negligible overlapping time dt ∼
1.5 fm/c. The right panels show the same string distributions in the Milne coordinates
in which more detailed space–time structures at large space–time rapidity ηs can be
shown. For 200 GeV collisions, most of the strings are located at a constant τ near the
mid-rapidity. A large spread of the strings in proper time is observed in the forward
and backward rapidity directions. This spread arises because as we move along the
forward rapidity, the Lorentz contraction of the projectile nucleus becomes weaker,
which leads to a finite overlapping time for the target nucleus to sweep through it.
Therefore, dynamical initialization schemes are important to address the early-time
dynamics of nuclear collisions at large rapidity regions and at low collision energies.
3 Dynamical Evolution of Heavy-Ion Collisions 141

5 4.0

3.5 0-5% Au+Au@200GeV


4
3.0

3 2.5

τ (fm/c)
t (fm/c)

2.0
2 1.5

nucleon 1.0
1 0-5% Au+Au decelerated
@200 GeV ∆τ = 0.5 fm/c 0.5

0 0.0
−5.0 −2.5 0.0 2.5 5.0 −6 −4 −2 0 2 4 6
z (fm) ηs
5 4.0
0-5% Au+Au@19.6GeV
3.5
4
3.0

3 2.5

τ (fm/c)
t (fm/c)

2.0
2 1.5

nucleon 1.0
1 0-5% Au+Au decelerated
@ 19.6 GeV ∆τ = 0.5 fm/c 0.5

0 0.0
−5.0 −2.5 0.0 2.5 5.0 −6 −4 −2 0 2 4 6
z (fm) ηs

Fig. 3.3 The energy–momentum strings’ space–time distributions in t–z and τ –η coordinates in
0–5% central Au+Au collisions at 200 and 19.6 GeV. The figure is taken from Ref. [8]

The three-dimensional MC-Glauber model is efficient for parameterizing the


energy loss of the participant nucleons as they pass through each other during the
nuclei impact [35]. The flexible parameterization provides an effective framework
to constrain the early-stage physics, such as baryon stopping, with experimental
measurements using the Bayesian Inference method.

3.1.4 Transport-Based Initial Conditions

The initial condition of the AMPT model [20] is based on the HIJING two-component
model [36]. The primary nucleon–nucleon interactions between the two incoming
nuclei are divided into a soft component described by the Lund string fragmentation
model [37] and a hard component of mini-jets described by perturbative QCD. This
generates a fluctuating initial condition that includes, for example, non-zero initial-
state momentum correlations as well as different harmonic components of the initial
spatial geometry. In the default version of the AMPT model [20, 38, 39], only the
mini-jet partons enter the parton cascade [40]. The rest of the excited strings are
considered to be idle during the parton phase, and they combine with the mini-jet
142 H. Elfner et al.

partons after the parton phase to fragment into initial hadrons that go through further
hadronic interactions. On the other hand, at high energies, the initial condition from
the string melting version of the AMPT model [20, 41, 42] should be applicable
since one expects the initial matter created in the overlap volume to have a high-
enough energy density for the creation of the quark–gluon plasma. In the string
melting version, the initial state essentially consists of the overlap region in parton
degrees of freedom in addition to spectator nucleons. The initial partons come from
the decomposition of hadrons from string fragmentation of the HIJING model into
valence quarks and antiquarks and thus do not include currently any gluons. Almost
all the energy in the overlap volume is converted via string melting to partons and
thus enters the parton cascade [40] for possible scatterings. Because of this and
the quark coalescence mechanism, the string melting AMPT model [20, 41, 42] is
able to reproduce the large anisotropic flows observed in ultra-relativistic heavy-ion
collisions with a rather small parton cross section of a few mb.
To include the finite nuclear thickness, the initial condition of the string melting
AMPT model has been extended [9] to specify the longitudinal coordinate z 0 and
time t0 of each excited string in a heavy-ion event, which then serve as the origin of
the partons produced from the melting of the excited string. Note that in the normal
string melting AMPT model [20, 41, 42], the longitudinal coordinate z 0 and time
t0 of each excited string in the initial state are both set to zero, which would be
correct only at very high energies. After including the finite nuclear thickness in the
model, one obtains the energy densities (solid curves) in Fig. 3.4; at low energies,
they are very different from the AMPT results that neglect the finite nuclear thickness
(dashed curves). One also sees that the increase in the maximum initial energy density
with the collision energy is much faster than the case that neglects the finite nuclear

Fig. 3.4 AMPT results of the average energy density at central space-time rapidity for central
Au+Au collisions at a 4.84 GeV, b 11.5 GeV, and c 27 GeV when including (solid curves) or
excluding (dashed curves) the finite nuclear thickness
3 Dynamical Evolution of Heavy-Ion Collisions 143

thickness, which is consistent with the analytical finding that the Bjorken formula
overestimates the maximum energy density more at lower energies [9, 11].
Strings from the AMPT initial condition have varying lengths along the space–
time rapidity. The strings are mainly formed by the stretching between quark–
antiquark or quark–diquark pairs in the dual parton model. Although nucleons in the
projectile or target have the same longitudinal momentum and rapidity, their com-
posite quarks or diquarks sampled from parton distribution functions have different
longitudinal momenta. As a result, the two ends of each string fluctuate strongly
from one string to another. After string melting, the extent of the space–time rapidity
of each string also fluctuates and is named as “string length fluctuations”. These
fluctuations originated from the parton distribution function provide one source of
fluctuations along the longitudinal direction. The mean “string length” is different
for different beam energies, and the fluctuations around the projectile and the target
become very strong for low beam energies as shown in Fig. 3.5.
Hadronic transport models, such as UrQMD or SMASH, provide the space–time
information of produced hadrons after string fragmentation with the Lund model
[44]. Those initial distributions of hadrons include event-by-event fluctuations from
collision geometry as well as from string fragmentation. They provide distributions
of energy–momentum currents and conserved charge densities. One can evolve the
collision system with hadronic transport for a short amount of time to simulate the
heat-up phase of the QGP and match the coarse-grained energy–momentum tensor
T µν to hydrodynamics at a given proper time after the two colliding nuclei pass
each other [7, 45]. In this approach, non-trivial pre-equilibrium flow profiles will be
developed from the hadronic transport model.
The separation of the high-density part (core) and the low-density part (corona)
is significant for the description of the centrality dependence of nuclear collisions at
SPS and RHIC energies [46]. Core-corona separation has also been implemented into
the UrQMD hybrid approach [47], and it is shown that such a core-corona separation
improves the description of experimental data, especially strange particle ratios and
flow.
Alternatively, a different approach treats the produced hadrons as local energy–
momentum sources and dynamically feeds them to the hydrodynamic fields [29]. This
approach considers only primary collisions between the nucleons in the incoming
nuclei. The produced hadrons from the nucleon–nucleon collisions first free-stream
by a short amount of time before depositing their energy, momentum, and charges
into the hydrodynamic fields. A similar dynamical initialization framework is also
applied for the parton system at LHC energies [48, 49].
A dynamical coupling of the microscopic transport model and macroscopic hydro-
dynamics is implemented in the JAM model [30]; it solves the time evolution of a
particle–fluid system. The fluidization happens by the same method as a dynamical
initialization, and fluid elements are converted into particles by using the Cooper–
Frye formula when local energy density becomes smaller than a particlization energy
density. Fluid–particle conversation occurs locally at each time step; it thus incorpo-
rates a dynamical core-corona separation.
144 H. Elfner et al.

Fig. 3.5 The fluctuations along the space–time rapidity from the AMPT initial condition for
√ √
s N N = 200 GeV Au+Au collisions at RHIC and s N N = 2.76 TeV Pb+Pb collisions at LHC.
The ends of the strings fluctuate more strongly at low beam energies than high beam energies. The
figure is taken from Ref. [43]

In the left panel of Fig. 3.6, the time evolution of the energy density from JAM

hybrid simulations of central Pb+Pb collisions at sNN = 6.4 GeV is plotted. The
dynamical fluid–particle conversion shows that hydrodynamical evolution already
starts before two nuclei pass through each other. The right panel of Fig. 3.6 shows the
time evolution of the fluid fraction in the central region |z| < 1 fm for central Au+Au
√ √
collisions at s N N = 2.7, 3.3, and 4.9 GeV and Pb+Pb collisions at s N N = 12.4
and 17.3 GeV. In the calculations, fluidization energy density e f = 0.5 GeV/fm3
is assumed. The fluid fraction increases slowly with time at lower beam energies.
As beam energy becomes higher, the fluid fraction increases rapidly, and the fluid–
particle conversion process becomes closer to a single thermalization time.
3 Dynamical Evolution of Heavy-Ion Collisions 145

Fig. 3.6 Time evolution of the energy density at the coordinate origin (x, y, z) = (0, 0, 0) in Pb+Pb

collisions at sNN = 6.4 GeV (left panel) and the fractions of the fluid energy at the central region

|z| < 1.0 fm (right panel) from JAM hybrid simulations of central Au+Au at s N N = 2.7, 3.3, and

4.9 GeV and Pb+Pb collisions at s N N = 12.4 and 17.3 GeV. The figure is taken from Ref. [30]

3.2 Transport and Hydrodynamics

In this section, we will discuss recent theoretical developments to describe the colli-
sion system’s evolution in a baryon-rich environment. We start with highlighting pro-
gresses in studying different equations of state (EOS) in microscopic transport models
and an improved parton–hadron interface in the AMPT model. For the strongly cou-
pled QGP phase, the system is more effectively described by macroscopic relativistic
viscous hydrodynamics. We will discuss the essential ingredients to generalize the
fluid-dynamic framework in the high baryon density region and how it interfaces with
the microscopic transport models in the hadronic phase. Lastly, recent advancements
to constrain QCD EOS from machine learning techniques are highlighted.

3.2.1 Equation of State in Microscopic Transport Models

Transport model simulates the Boltzmann-type collision term in a Monte Carlo fash-
ion; thus, the equation of state (EOS) in the cascade model is expected to be close to a
free hadron or parton gas. It is possible to implement a different EOS in a microscopic
transport model by including mean field. There are two approaches; the Boltzmann–
Uehling–Uhlenbeck (BUU) or Vlasov–Uehling–Uhlenbeck (VUU) model [50–52],
which were developed to simulate a space–time evolution of one particle distribution
function f (x, p):
∂f
+ v · ∇r f − ∇r U · ∇ p f = Icoll (3.6)
∂t
where Icoll is a collision integral including the Pauli blocking. If Icoll = 0, one obtains
the Vlasov equation. U is a mean-field potential. The Skyrme parametrization is often
used [53, 54]:
146 H. Elfner et al.
#
f (x, p ( )
d 3 p(
γ
U (ρ B , p) = αρ B + βρ B +C (3.7)
1 + ( p − p( )2 /!2

where the baryon density ρ B -dependent potential has two terms: α is attractive and β
is repulsive, and a momentum p-dependent part. Another approach is called quantum
molecular dynamics (QMD)[55, 56], an N -body theory, and can be used for event-by-
event simulations. The molecular dynamics is a simulation that solves Hamilton’s
equations of motion for N -body system. QMD is an extension of the molecular
dynamics approach by including a collision term. QMD also uses the same Skyrme-
type potential. Later relativistic QMD (RQMD) was formulated, in which the Skyrme
potential is implemented as a Lorentz scalar [57–60]. A relativistic transport approach
based on the relativistic mean field (RMF) has been developed, called RVUU [61–
64], RBUU [26, 65, 66], or RLV [67]. In these covariant approaches, relativistic
meson mean fields are used instead of non-relativistic Skyrme potential:
$ ∗ %
p ∗µ ∂µ + ( pν∗ F µν + m ∗ ∂ µ m ∗ )∂µp f (x, p ∗ ) = Icoll (3.8)

where F µν = ∂ µ ω ν − ∂ ν ω µ is the field strength tensor, m ∗ = m − gs σ is an effec-


tive mass due to scalar meson σ, and p ∗ = p − gv ω is a kinetic momentum. A
relativistic version of QMD with the relativistic mean field has also been developed
(RQMD.RMF) [68–70].
A first-order phase transition at high baryon density has been studied in a micro-
scopic transport model [71–73]. The effects of the spinodal region of the phase
transition and the critical point were studied in the hadronic transport model [74].
The parity doublet model is implemented for the first time in a RBUU approach [75],
called DJBUU, and applied to low-energy heavy-ion collisions.
It is also known that the EOS can be controlled by changing the scattering style
in the two-body collision term [76–78]. For example, choosing a repulsive orbit
in two-body collisions can simulate the effect of a repulsive potential. The effects
of a phase transition have been investigated by using attractive orbits in two-body
collisions [79]. It is possible to control the pressure of the system by changing the
scattering style to simulate a given EOS [80]. Recently, this idea was employed to
investigate the sensitivity of the anisotropic flows (v1 , v2 , v4 ) to the properties of
phase transitions [81–83].

3.2.2 Parton–Hadron Interface in AMPT

Since the AMPT model [20] has both parton and hadron phases, a hadronization
model is needed to convert the parton matter into a hadron matter at the end of the
parton phase or the QGP lifetime of each event. In the default version of the AMPT
model [38, 39], after parton scatterings are over, the mini-jet partons recombine with
their parent strings and then fragment to initial hadrons via the Lund string fragmen-
tation. On the other hand, in the string melting version of the AMPT model [41], a
3 Dynamical Evolution of Heavy-Ion Collisions 147

quark coalescence model is used to describe the hadronization process. It is a spa-


tial coalescence model [20, 41] where a quark combines with the closest antiquark
to form a meson or with the two closest quarks to form a baryon. The species of
the formed hadron is determined according to the flavors and invariant mass of the
coalescing partons. The original quark coalescence model in AMPT [20, 41, 42]
searches for a meson partner before searching for baryon or antibaryon partners.
In addition, the original coalescence algorithm depends on whether a quark comes
from the melting of a meson or a baryon; as a result, for each event it conserves
the numbers of baryons, antibaryons, and mesons separately during the coalescence
process.
Recently, the quark coalescence model is improved [84], where the artificial con-
straint that forced the separate conservation of the numbers of baryons, antibaryons,
and mesons for each event is removed. Note that the new coalescence model still
satisfies the conservation of the number of net baryons for each event, as well as the
conservation of the number of net-strangeness and the number of net-charges. When
both the meson partner and baryon partners are available, a quark will form a meson
or a baryon according to the following criteria:

d B < d M r B M : form a baryon;


otherwise : form a meson. (3.9)

In the above, d M and d B represent, respectively, the relative distance among the two
or three coalescing partons in their rest frame, and r B M is the coalescence parameter
that controls the relative probability of a quark forming a baryon versus that forming
a meson [84]. This way the new quark coalescence allows a quark the freedom to
form a meson or a baryon depending on the distance from the coalescence partner(s);
this algorithm is thus more physical. The improved quark coalescence model better
describes baryon observables in general, especially the (anti)baryon pT spectra and
antibaryon-to-baryon ratios for multi-strange baryons [84]. It is also found to be
important for the description of two-particle angular correlations in pp and p-Pb
collisions at LHC energies [85], particularly the near side anti-correlation of baryon
pairs as observed in the experiment.

3.2.3 Relativistic Viscous Hydrodynamics



For relativistic nuclear collisions around sNN ∼ 10 GeV, dynamics of conserved
charges need to be considered together with the evolution of energy and momentum
in the system. The relativistic hydrodynamic equation of motion can be written as

∂µ T µν = 0, (3.10)
µ
∂µ Ji = 0. (3.11)
148 H. Elfner et al.

Here, the system energy–momentum tensor is

T µν = !u µ u ν − (P + ")#µν + π µν (3.12)

and the conserved charge currents are


µ µ
Ji = n i u µ + qi with i = (B, S, Q). (3.13)

Here, ! is the local energy density, n i is the conserved charge density, P(!, n i )
is the thermal pressure given by the equation of state (EOS), and u µ is the flow
velocity. The off-equilibrium aspects are characterized by the shear viscous tensor
µ
π µν , the bulk viscous pressure ", and the charge diffusion currents qi . In heavy-ion
collisions, there are multiple conserved charge currents, namely the net baryon B,
net-strangeness S, and net electric charges Q, carried by fluid constituents.
The hydrodynamic equation of motion needs to be solved together with the fluid’s
equation of state P(!, n i ). At zero net-charge density, the QGP’s equation of state
can be computed from lattice QCD [86, 87]. It is smoothly connected with hadron
resonance gas (HRG) EOS at low temperature. The Taylor expansion technique has
been employed to extend the QCD EOS to finite densities [88, 89]. Lattice QCD
calculations show that the phase transition between QGP and HRG remains as a
smooth crossover for the region where µ B /T ≤ 2. Reference [90] gives a detailed
review on the recent progress in constructing QCD EOS at finite densities.
To continue our discussion of the hydrodynamic evolution equations, we present
the second-order causal constitutive relations for the shear and bulk viscous parts
and the diffusion for net-baryon current. We use the expressions derived in [91, 92],
given by
˙ + " = −ζ θ − δ"" " θ + λ"π π µν σµν ,
τ" " (3.14)
*µ *µ
τπ π̇ *µν+ + π µν = 2η σ µν − δππ π µν θ + ϕ7 πα π ν+α − τππ πα σ ν+α + λπ" " σ µν , (3.15)
µB µB
τq q̇ *µ+ + q µ = κ B ∇ µ − δqq q µ θ − λqq qν σ µν + lqπ #µν ∂λ π λ ν − λqπ π µν ∇ν , (3.16)
T T

µν
where A*µ+ = #µν Aν and A*µν+ = #αβ Aαβ , θ = ∇µ u µ is the expansion rate, and
& '
µν 1 µ ν ν µ 2 µν α
σ = ∇ u + ∇ u − # (∇α u ) (3.17)
2 3

is the velocity shear tensor, with the spatial projection operator ∇µ ≡ #µν ∂ ν =
µν
(gµν − u µ u ν )∂ ν . The higher rank spatial projection operator is defined as #αβ ≡
1
2
[#µ α #ν β + #µ β #ν α ] − 13 #µν #αβ . In Eqs. (3.14)–(3.16), the viscous stress ten-
sor and diffusion current are driven by their corresponding thermodynamic forces,
θ, σ µν , and ∇ µ µTB , and ζ, η, and κ B are the medium’s bulk and shear viscosity and
net-baryon diffusion coefficient, respectively. The evolution of the viscous stress ten-
sor and diffusion current is controlled by their relaxation times, τπ , τ" , and τq . The
DNMR theory included additional second-order terms in spatial gradient together
3 Dynamical Evolution of Heavy-Ion Collisions 149

with their transport coefficients [91, 92]. Some of them coupled the π µν , ", and q µ
with each other in their evolution equations. Moreover, because there are multiple
conserved charges in QCD and quarks in their flavor eigenstates carry a mixture of
these quantum charges (B, S, Q), the charge diffusion currents will mix with each
other and the diffusion coefficient κ becomes a 3 × 3 matrix [93, 94]. Theoretical
frameworks that incorporate diffusion effects on multiple conserved charge currents
are being developed [95].
The three-fluid hydrodynamics (3FH) simulation [96, 97] treats the collision of
two fluids, and during the collision, a third fluid was developed by particle production.
Thus, it incorporates the effects of EOS including the compression stages of heavy-
ion collisions. The 3FH has been used to analyze heavy-ion collisions in the high
baryon density region.

3.2.4 Hydrodynamics-Transport Interface

As hydrodynamics evolves the system to dilute regions with energy density below 0.5
GeV/fm3 , the strongly coupled collective description becomes less and less valid. In
these regions, we can switch to a hadronic transport model to describe the dynamics
microscopically. A review of hybrid approach applications to lower beam energies
is available in [98]. This switching requires to map individual fluid cells to particles,
dubbed the particlization procedure. The conversion from fluid cells to particles uses
the Cooper–Frye procedure,

gi d 3 p µ
d N (x, p) = p dσµ ( f 0 + δ f ), (3.18)
(2π)3 E

where f 0 is the Bose/Fermi distribution at thermal equilibrium, and the out-of-


equilibrium corrections from shear and bulk viscosity and net-baryon diffusion are
included in the δ f .
The Cooper–Frye prescription provides averaged value of particle emission from
the individual fluid cell. Conservation laws for energy–momentum and charge cur-
rents are only ensured on an event-averaged basis. One can impose global or local
conservation laws during the particlization stage, which will build in non-trivial
multi-particle correlations among the emitted particles.
One conceptual issue is the negative contributions to the Cooper–Frye hypersur-
face. Due to the fast expansion of the system, only a very small fraction of particles
is emitted inwards from the hypersurface and would either need to be added to the
hydro evolution or subtracted from the final stage of hadronic rescattering. The sec-
ond option has been explored in [99], and the magnitude of the effect, depending on
phase space and particle mass, has been assessed in [100]. In particular, in event-by-
event calculations, the negative contributions can become sizeable, and a concurrent
treatment of hydrodynamics and hadronic transport has to be implemented to quan-
tify the sensitivity of observables.
150 H. Elfner et al.

The standard Cooper–Frye particlization sampling algorithms are based on the


assumption of a grand-canonical ensemble. Therefore, the quantum numbers such as
energy, momentum, and other conserved charges are only conserved on the average.
For fluctuation and correlation observables that are sensitive to event-by-event and
local fluctuations, observing conservation laws might have crucial effects. There-
fore, within the UrQMD hybrid approach, global conservation of energy and baryon
number, strangeness as well as electric charge has been implemented [45]. Later, it
has been found that this “mode sampling” introduced some bias that can be cured
by employing a Metropolis algorithm [101]. Within the BEST collaboration, more
sophisticated schemes for global and local conservation of charges as well as energy
and momentum have been developed [102–104] and are publicly available. The
remaining issue is that those algorithms rely on splitting the hypersurface into regions
(“patches”) and it is not obvious how their size can be uniquely determined.
Another crucial aspect in matching the hydrodynamic evolution to hadronic trans-
port for the final stage of decoupling is the equation of state. In principle, the equation
of state employed during the hydrodynamic evolution should match the degrees of
freedom of the particular hadronic transport approach. In practice, the pressure is not
very sensitive to additional hadronic degrees of freedom, and therefore, it is sufficient
to match the thermodynamic quantities on the hypersurface to the proper hadronic
equation of state. At finite net-baryon densities, it is important to use an energy
density as a switching criterion instead of a constant temperature as often done in
ultra-relativistic heavy-ion reactions. The switching has to take place when most
cells have transitioned to the hadronic regime within the hydrodynamic evolution,
but before the chemical and kinetic freeze-out which will be dynamically generated
within the hadronic transport calculation.

3.2.5 QCD Equation of State from Machine Learning

The EOS is an input to the relativistic hydrodynamic model. As a result, different


EOS lead to different space–time evolution histories of QGP and hadronic resonance
gas (HRG). The evolution histories are visually separable for EOS with crossover
transitions or first-order phase transitions between QGP and HRG. Quantitatively,
the differences are largest around the “soft point” of the first-order phase transition
region, where the pressure gradient is zero in the Maxwell construction or negative
in Spinodal construction. The radial flow decreases in this region, and the signal
was expected to survive and exist in the final state of heavy-ion collisions through
pT spectra and dv1 /dy [105] after the particlization. In practice, it is very difficult
to discriminate the QCD phase transition types using these observables as they are
coupled with many other model parameters.
One exploratory direction that may help in the future is machine learning which is
the state-of-the-art pattern recognition method. Different from traditional observables
using feature engineering, the machine learning method can help to design new
observables from old ones or from high-dimensional data. This is the easiest in
3 Dynamical Evolution of Heavy-Ion Collisions 151

supervised learning where the machine is trained to classify labeled events into two
different categories if they are generated with two different EOS. The inputs to the
machine learning algorithms are the EOStypes used as labels and the four-momenta of
final-state particles in each single event from relativistic hydrodynamic simulations.
This procedure is employed to identify crossover from first-order phase transition
[106, 107], to distinguish the Maxwell construction from Spinodal construction in
the first-order phase transition region [108], and to classify hard and soft EOS at
high-density region [109]. At the same time, an unsupervised auto-encoder is used
to classify nuclear liquid–gas transitions [110].
Recently, a dynamical edge-convolution neural network is constructed to look
for critical correlations and self-similarity in the momentum distribution of final-
state hadrons[111]. As shown in Fig. 3.7, the input to the network is the particle
list and the outputs have two branches. One branch has one neuron in the output
layer, whose value represents the probability that critical correlations are encoded in
some of the final-state hadrons. The other branch has the same number of neurons
as input particles, whose values are used to tag whether a particle belongs to a signal
or background. The dynamical edge-convolution block repeats two times. It first
looks for the k-nearest-neighbor of each particle in momentum space. Here, the
Euclidean distance between each pair of particles is computed, and the k-nearest-
neighbor of one particle stands for the k number of particles that are closest to the
chosen particle. The information of each particle and its neighbors is sent to the
edge-convolution block to get a new representation of each particle, which is simply

Particles ( kNN + Edge Convolution ) × 2 Point Cloud Network Tagging

Px , Py 1D CNN latent features 1D CNN latent features 1D CNN Signal

Px , Py 1D CNN latent features 1D CNN latent features 1D CNN Noise

Px , Py 1D CNN latent features 1D CNN latent features 1D CNN Noise

… … … … … 1D CNN …

Px , Py 1D CNN latent features 1D CNN latent features 1D CNN Noise

Global Max Pooling


Classification

Find its k nearest neighbors in feature space. CP / JAM

Fig. 3.7 A dynamical edge-convolution neural network developed to identify events with critical
correlations and to tag signal particles
152 H. Elfner et al.

a list of numbers called features. These features are nonlinear transformations of


particles from their momentum space, which will form a feature space with each
feature carrying important information. The second time, the network looks for the
k-nearest-neighbor of each particle in feature space. It is thus possible to learn multi-
particle short-range correlation in feature space which corresponds to long-range
correlation in momentum space. This is important for many classification and tagging
tasks in the field of heavy-ion collisions.
The supervised learning mentioned above is different from the Bayesian analy-
sis method which is used to obtain a posterior distribution of the QCD EOS that
is consistent with Lattice QCD calculations at zero net-baryon density [112]. The
supervised learning tries to learn features (or observables) that are most relevant to
the EOS while at the same time ignoring the effect of other model parameters. This
is achieved by training machine on diverse data to make consistent predictions on
EOS for various other conditions. In the end, although it is still unclear how the
machine makes its decision, it is clear that the machine has indeed learned some
important features for EOS classification, based on its high prediction accuracy. At
the same time, it is easy to extract the most relevant momentum space region for
EOSclassification using the attention method in interpretable machine learning.

3.3 Fluctuation Dynamics

Exploring and characterizing the properties of hot and dense QCD matter in heavy-ion
collisions has been one of the main motivations for developing fluctuating hydro-
dynamics in the heavy-ion theory community. It was known long ago [113, 114]
that fluctuations of the produced hadron multiplicities, in particular, those of pro-
tons, would be enhanced near the conjectured QCD critical point. It was appreciated
at about the same time that the critical fluctuations would inescapably fall out of
equilibrium in the expanding fireball created in a heavy-ion collision because of
the critical slowing down [115]. This calls for the building-up of the dynamical
framework to describe off-equilibrium critical fluctuations based on the fluctuating
hydrodynamics. Besides studying the critical dynamics, the enhanced thermody-
namic fluctuations near the phase boundary may also contribute significantly to the
behavior of transport coefficients and equation of state [116, 117]. More recently,
the study of hydrodynamic fluctuations has allowed characterizing the evolution of
the QGP properties as a function of frequency and wavelength in certain contexts.
Hydrodynamic fluctuations may also be significant in the far-from-equilibrium QGP
created in the initial stages of both large and small colliding systems.
The development of the theory of hydrodynamic fluctuations can be dated back to
the 1950s, when Landau and Lifshitz formulated their stochastic approach. Remark-
ably, this is still an area under active research. In Sect. 3.3.1, we shall briefly review
three main approaches to fluctuating hydrodynamics, with an emphasis on the two
recent ones. We will also explain the qualitative features of hydrodynamic fluctua-
tions. The subsequent section, Sect. 3.3.2, provides a discussion on the application
3 Dynamical Evolution of Heavy-Ion Collisions 153

of fluctuating hydrodynamic to the search for the QCD critical point, while its other
possible applications will be discussed in Sect. 3.3.3. In Sect. 3.3.4, we list a number
of open questions.

3.3.1 Fluctuating Hydrodynamics

Fluctuating hydrodynamics describe the evolution of (average) hydrodynamic vari-


ables and their fluctuations. Let us consider a fluid. We may view each fluid cell as
a grand-canonical ensemble that exchanges energy–momentum and charge with its
surrounding environment and hence experiences fluctuations. If those fluctuations
were in equilibrium, no special study is needed since the magnitude of those fluc-
tuations is simply given by the appropriate thermodynamic functions. For example,
the equilibrium fluctuation of energy density per unit volume is nothing but the spe-
cific heat cV . Nevertheless, since the conserved densities can only relax through the
diffusive process, the equilibration rate of density fluctuations approaches zero in
the long-wavelength limit. As a result, in many situations of physical interest, such
as the expanding QGP, thermodynamic fluctuations are out of equilibrium. Such
off-equilibrium deviation in turn would back-react on the properties of the medium
and modify the equation of state (EOS)and transport coefficients. Because of that,
the background evolution would also be influenced. In short, there are intertwined
dynamics among the fluctuations, the properties of the medium, and bulk evolution
as illustrated in Fig. 3.8. The goal of fluctuating hydrodynamics is to describe such
interesting and intriguing physics.

Fig. 3.8 An illustration of the key elements and their mutual relations in fluctuation dynamics.
Consider a fluid under expansion (or external disturbance): the real-time fluctuation would lag
behind the equilibrium expectation and back-react on the equation of state (EOS) and transport
coefficients which control the hydrodynamic force that drives the fluid flow. Meanwhile, the flow
will transport off-equilibrium fluctuation in space. The goal of fluctuating hydrodynamics is to
describe such an intertwined dynamical process
154 H. Elfner et al.

One important qualitative feature of the fluctuating hydrodynamics is an asso-


ciated characteristic momentum scale p ∗ . This scales arises from the competition
between diffusive rate and expansion rate ω, which leads to the relation [118]
(
ω
Dp∗2 = ω → p∗ ∼ . (3.19)
D

Furthermore, the relative importance of fluctuation corrections is determined by three


factors; (a) the strength of non-linearity, (b) the ratio between the phase-space volume
of the long-wavelength fluctuating modes, p∗3 , and that of the whole system, i.e., the
entropy density s, and (c) the ratio between the relaxation rate of fluctuation and
microscopic relaxation rate,

#η $fluct p∗3
∼ . (3.20)
η $mic s

Having described the effects of hydrodynamic fluctuations qualitatively, we will


now turn to more quantitative analysis. Here, we will provide a brief review of three
complementary approaches to hydrodynamic fluctuations, which will be referred to
as “stochastic approach”, “deterministic approach”, and “EFT approach”, respec-
tively.
In the traditional stochastic approach, a la Landau–Lifshitz [119], the effects
of fluctuations are accounted for by first adding stochastic noise force F ν to the
conservation equations, i.e.,

∂µ T µν = F ν (3.21)

where T µν denotes part of the stress–energy tensor which can be expressed in terms
of one-point function of hydrodynamic variables such as energy density ! and flow
velocity u µ through the usual constitutive relation. The noise force F µ is assumed
to be local, i.e., *F(t, x) F(t ( , x ( )+ ∝ δ(t − t ( )δ 3 (x − x ( ) (suppressing the Lorentz
index), with the magnitude of noise–noise correlation fixed by the fluctuation–
dissipation theorem. The extension of “stochastic approach” to relativistic hydro-
dynamics can be found in Ref. [120].
More recently, there are rapid developments in studying the same physics using the
“deterministic approach”, see, e.g., Refs. [118, 121–128]. In this approach, the wave-
number-dependent correlation functions of hydrodynamic variables are treated as
slow variables in addition to hydrodynamic ones. The stochastic equations are traded
off by a set of coupled deterministic equations among those correlation functions and
conventional hydrodynamic variables. This approach successfully describes a num-
ber of non-trivial off-equilibrium effects, including the hydrodynamic tail [118],
and has been formulated for both Bjorken-expanding background and general fluid
background [125]. Simulations using the “stochastic approach” is quite straightfor-
ward but numerically demanding. Meanwhile, the “deterministic approach” costs
less numerical resource but relies on a non-trivial formulation.
3 Dynamical Evolution of Heavy-Ion Collisions 155

Although at this moment, “stochastic approach” and “deterministic approach” are


the only two methods that can be used to practically simulate fluctuating dynamics
in heavy-ion collisions, there is an important third approach, “EFT approach”, which
emerges from the recent breakthrough in developing effective field theories (EFTs)
for non-equilibrium dynamics. Consider a many-body system described by a given
microscopic theory. One could imagine formally integrating out fast modes in the
long-wavelength and low-frequency limit. Then, the systems should be described by
an effective action with hydrodynamic modes as slow modes, i.e.,
# #
Z= D ψmic e i Imic[ψmic ]
→ D ψhydro ei Ihydro[ψhydro ] , (3.22)

where Ihydro can be constructed based on the symmetry and physical constraints
of the system, as was shown explicitly in Refs. [129, 130] (see also the lecture
notes [131]). Although the formalism of Refs. [129, 130] has only been published a
few years ago, the studies based on this formalism have already uncovered several
interesting new phenomena in problems that were extensively studied before [132,
133]. To date, this formalism has already helped considerably clarify aspects of the
theoretical structures of hydrodynamic fluctuations. In our view, this new frame-
work has many potential applications in the future, particularly in the perspective of
implementation in quantum computing. Since such formalism does not depend on
any long-wavelength expansion, it may also serve as a basis for studying QGP at a
mesoscopic scale with suitable extension [129].

3.3.2 Fluctuation Dynamics Near the QCD Critical Point

The existence of a critical point manifests itself through large correlations and
enhanced fluctuations. More quantitatively, let us consider (the Wigner transfor-
mation of) the two-point function for the fluctuation of the order parameter field
M(t, x):
#
φ Q (t, x) ≡ d 3 y *δ M (t, x − y/2) δ M (t, x + y/2)+ e−i y· Q . (3.23)

Here, φ Q (t, x) describes the magnitude of the critical fluctuation at wavelength 1/Q.
For a critical system in equilibrium, the equilibrium expectation takes form1

1
φ̄(Q) ∼ , (3.24)
ξ −2 + Q2

1 For illustrative purpose, we only show the approximate expression here. See Ref. [122] and
references therein for more details.
156 H. Elfner et al.

which tells us that long-wavelength fluctuation (LWF) of the order Q ∼ ξ −1 will


grow as ξ 2 as the critical length ξ increases. The critical behavior of LWF leads
to the familiar phenomenon of critical opalescence and is the key to the search for
the QCD critical point. In particular, LWF leads to non-trivial behavior of Gaussian
and non-Gaussian fluctuations of hadron multiplicities and gives rise to non-trivial
scaling behavior of E.o.S near the critical point.
The discussion so far assumes that the critical fluctuation is in equilibrium. How-
ever, as the fireball approaches the critical point, LWF inescapably falls out of equi-
librium due to critical slowing down. In a set of papers [115, 134] (see Ref. [135]
for more references), the real-time fluctuations are found to differ from equilibrium
expectations not only quantitatively but also qualitatively in many phenomenolog-
ically relevant situations. In addition, the order parameter fluctuations also back-
react on the bulk evolution of the fireball. Therefore, a quantitative framework that
describes the coupled dynamics of both the fluctuations near the phase boundary and
the bulk background evolution is crucially needed. As we have explained before, the
basis of this framework should be fluctuating hydrodynamics.
Before turning to the quantitative framework, it is again instructive to first consider
the qualitative features of critical dynamics, which is characterized by the Kibble–
Zurek (KZ) dynamics (see Ref. [136] for a review). Briefly, the evolution of the
critical fluctuations becomes effectively frozen at the point where the time remaining
to reach the critical point is shorter than the relaxation time. This frozen correlation
length lKZ and the frozen time of critical fluctuations τKZ are known as the KZ length
and time, respectively. The KZ length gives the characteristic length scale for the
off-equilibrium fluctuations and is a generalization of “diffusive length scale” as
introduced in Eq. (3.19). Similarly, the KZ time, τKZ gives the time window when
the off-equilibrium effects are important. According to the benchmark estimation
in Ref. [123], τKZ for a heavy-ion collision is around 6 fm. As demonstrated in
Ref. [137], such off-equilibrium scaling leads to a potentially unique signature of the
QCD critical point.
Now we turn to the quantitative approach. The Hydro+ formalism, which fol-
lows the deterministic approach of fluctuating hydrodynamics, has been developed
to describe the intertwined dynamics among critical fluctuations and bulk evolu-
tion [122]. The key new ingredient in Hydro+ is to treat the Wigner function of the
order parameter field, defined in Eq. (3.23), as a dynamical variable. According to
the general formalism of “deterministic approach”, this Wigner function obeys a
relaxation rate equation:
! " ! "
u τ ∂τ + u i ∂i φ Q (t, x) = −$ Q φ Q (t, x) − φ̄ Q (t, x) , (3.25)

where φ̄ is the equilibrium value of φ. The stress–energy tensor T µν and baryon


number current J µ are still conserved. Their conservation equations, ∂µ T µν = 0
and ∂µ J µ = 0, together with Eq. (3.25) form the complete equations of motion for
Hydro+. Importantly, both the transport coefficients and EOShave to be generalized
in Hydro+. Specifically, the constitutive relation for T µν takes the form:
3 Dynamical Evolution of Heavy-Ion Collisions 157

T µν = εu µ u ν + p(+) (g µν + u µ u ν ) + viscous terms , (3.26)

and the expression for J µ is similar; see Ref. [122] for details. Note that the gen-
eralized pressure p(+) depends not only on the hydrodynamic variables ε and n B ,
the energy, and baryon number densities, but also on the additional Hydro+ variable
φ Q (t, x). Since the hydrodynamic (collective) flow is induced by the gradient of the
generalized pressure, the bulk evolution in Hydro+ is intrinsically coupled with that
of φ Q (t, x). Therefore, Hydro+ couples LWF with hydrodynamics self-consistently.
See Refs. [125, 126] for further refinements and improvements on Hydro+ formal-
ism.
There are two recent exploratory simulations, Refs. [138, 139], within the the
Hydro+ framework. These are done in dynamical settings similar to that encountered
in a heavy-ion collision albeit with simplified system geometry and equation of state.
We now discuss the main lessons learned from those simulations.
First, we observe that φ(Q) at small Q stays at its initial value. This clearly
demonstrates the crucial role of the conservation law: φ(Q = 0) corresponds to the
fluctuation of the order parameter averaged over the whole volume. If the order
parameter is associated with conserved densities, the fluctuation at Q = 0 cannot
evolve at all. Second, we see that the radial flow transports fluctuations by advection,
and quantitative studies must consider the advection effects. This is seen in Fig. 3.9
(right) where we show φ versus r at the representative value of Q = 0.4 fm−1 from
[138]. We observe the peak in equilibrium expectation of φ̄(Q) moving inward
from early time to later time. On the other hand, the combination of two out-of-
equilibrium effects determines the spatial dependence of the fully dynamical φ(Q):
(a) the “memory”/“lag behind” effect that is visible in the difference between the
peak in the equilibrium expectation (dashed curve) and the actual φ; (b) the advection
in the outward-flowing fluid that carries the peak in the fluctuations.

Fig. 3.9 The magnitude of the critical fluctuations φ(Q) (with appropriate normalization) versus
the radius r at a representative momentum Q from Ref. [138] (left) and Ref. [139] (right). In all
figures, dashed (solid) lines show the equilibrium (non-equilibrium) values
158 H. Elfner et al.

Finally, we note that those non-equilibrium contributions from the slow modes
to bulk matter properties (e.g., entropy and pressure) are generically small. In both
studies, the off-equilibrium slow-mode contribution #s to the entropy density is
generically of the order:

#s
∼ O(10−5 −10−4 ). (3.27)
s
This can be understood by comparing the phase-space volume of off-equilibrium
critical modes Q 3neq /(2π)3 with the typical entropy density s :

) *) *3
#s 1 T3 Q neq
∼ , (3.28)
s (2π)3 s T

where Q neq ∼ ξ −1 denotes the typical momentum which is not in equilibrium. Using
the benchmark value s = (4π 2 (Nc2 − 1) + 21π 2 N f )T 3 which corresponds to the
entropy density of an ideal QGP at zero baryon chemical potential, we arrived at
(#s/s) ∼ O(10−4 ). Interestingly, the above analysis on back-reaction is consistent
with the study of the bulk viscosity near the QCD critical point in Ref. [140]. From
a practical perspective, the smallness of back-reaction effects suggests that one may
neglect the back-reaction in future phenomenological modelings, which will help
significantly reduce the computational cost.
The “deterministic approach” is not the only possible formalism for the study of
critical fluctuations. The authors of Ref. [141] consider the fluctuations of the net-
baryon density near the critical point based upon simulating stochastic equations. In
this benchmark calculation, both the expected dynamical scaling behavior and the
impact of critical slowing down are observed.

3.3.3 Other Applications

The effects of hydrodynamic fluctuations might be of importance both near and away
from the critical points. Consequently, a theoretical understanding of fluctuating
hydrodynamics is expected to have broad applications in various aspects of heavy-
ion collisions. In this section, we shall discuss several examples.
The study of fluctuation dynamics may help us understand the behavior of trans-
port coefficients in the crossover regime of the QCD phase diagram. This is because
the magnitude of fluctuations would be enhanced around the phase boundary. For
example, the evolution of the fluctuations of conserved charges is investigated in
Refs. [127, 128] in order to constrain the charge diffusive constant of the quark–
gluon plasma from experimental measurements of the balance function at top RHIC
energy. In Ref. [121], the relation between the non-trivial behavior of the bulk vis-
cosity in the crossover region and the hydrodynamic fluctuations has been analyzed.
3 Dynamical Evolution of Heavy-Ion Collisions 159

The correlators of stress–energy tensor and charge current are important for char-
acterizing both the static and transport properties of a medium. Based on the theory
of hydrodynamic fluctuations, one can study the effects of fluctuations on those
correlators. The results of doing so may be viewed as finite frequency and finite
gradient corrections to the transport coefficients and equation of state. The investi-
gation along this line helps reveal how those properties would evolve as a function
of relevant scale as well as non-trivial analytical structure induced by fluctuation
dynamics. Those effects might be relevant for the initial stages of the fireball as well
as for small colliding systems.
Finally, we note that with suitable generalizations, the formalism of fluctuating
hydrodynamics can also be used to study hydrodynamics with chiral anomaly which
couple non-conserved axial charge densities to hydrodynamic modes [142].

3.3.4 Open Questions and Future Directions

We have reviewed the application of the fluctuation dynamics to the study of the
QCD phase diagram, by examining the problem in two broad classes—near the
QCD critical point and away from the critical point. In both cases, we see that
the competition between the expansion and the equilibration leads to emergent new
features for both theoretical exploration and experimental observation. We anticipate
the development of fluctuating hydrodynamics would be particularly useful for the
search of signatures of the QCD phase boundary as well as for analyzing the off-
equilibrium properties of QCD matter.
So far, most of the studies are focused on the crossover region of the phase
diagram. The exploration of the high baryon density regime surely would bring out
new ingredients which require further studies. For the search for the QCD critical
point, the extension of the current framework to include non-Gaussian fluctuations
is imperative, though important first steps have been taken [126, 134]. A dynamical
framework describing pertinent physics of the first-order transition is also crucially
needed; see Ref. [47] on related developments. For cold and dense nuclear matter,
the effects of quantum fluctuations might be significant. We anticipate that through
a combination of the present and future theoretical developments together with an
abundance of experimental data from beam energy scan and system size scan, the
community will witness substantial progress in characterizing the phase structure
and dynamic evolution of QCD matter as well as mapping out a full picture of its
evolution as a function of temperature, density, and resolution scale.
160 H. Elfner et al.

3.4 Experimental Results on Flow

Before focusing on lower beam energies, let us review shortly the state-of-the-art
flow measurements at the highest energies at RHIC and LHC. At the LHC and top
RHIC energies, the dynamics of the heavy-ion collisions and produced QGP are
well described in terms of relativistic fluid dynamics. Due to the extremely short
crossing time, the energy deposition and entropy production happen very quickly
τ ! 0.1 fm/c, producing a hot, dense, and nearly thermalized QGP on the time scale of
τ ! 0.5 fm/c. Driven by the large pressure gradients, the QGP undergoes collective,
Hubble-like expansion, converting the spatial non-uniformities in the initial state into
the collective radial flow and azimuthally anisotropic flow in the momentum space.
They are characterized by the Fourier series in azimuthal angle:
+ ∞
-
d2 N 1 ,
= N ( pT ) 1 + 2 vn ( pT ) cos(n(φ − %n ) (3.29)
pT dpT dφ 2π i=1

where vn and %n represent the magnitude and the phase (referred to as the event
plane or EP) of the nth-order anisotropic flow, and the slope of the particle momen-
tum spectrum N ( pT ) reflects the strength of the radial flow, i.e., a flatter spectrum
or larger * pT + implies a stronger radial flow [143]. The flow develops mostly in the
early partonic phase when the pressure gradients are large, but persists all the way
through phase transition and up until freeze-out in the hadronic phase. The strength
of the radial and anisotropic flow are sensitive to not only the initial state, but also
the EOS and transport properties of the medium at various stages [144]. With the
extensive and precise measurements of vn and * pT + data, we are able to place quanti-
tative constraints on the bulk properties of the QGP such as shear viscosity η/s, bulk
viscosity ζ/s, and EOS. For example, recent state-of-the-art Bayesian analysis over
a space of hydrodynamic model parameters, obtained via fitting to experimental data
from both RHIC and the LHC, gives η/s ∼ (0.07+0.05 −0.04 ) + c(T − Tc ) for temperatures
T > Tc = 154 MeV with Tc corresponding to the crossover temperature [145–147].
The success of hydrodynamic description in high energy largely is due to the
fact that the system is nearly boost invariant and relatively long-lived, such that the
system evolution can be divided into different stages. But in the lower end of the

BES energies of sNN < 20 GeV, the boost-invariance assumption is no longer valid,
and the initial state and the space–time picture become much more complex [148].
Due to the increased crossing time, nuclear stopping in the longitudinal direction
and particle production in the transverse direction happens simultaneously, and there
is no clear distinction between the initial state and final state [8, 9]. The system,
the QGP phase, in particular, is more short-lived, and there is a large conceptual
uncertainty in the space–time dynamics. Furthermore, the lower energy region also
probes a medium at low T and large µ B for which one expects very different bulk
properties. Describing such a system requires full 3+1D models, either hydrodynamic
model or transport approach, which have been the focus of many theoretical efforts.
From the experimental point of view, all flow observables are also expected to show
3 Dynamical Evolution of Heavy-Ion Collisions 161

qualitatively different behavior if compared to higher energies. In particular, the


radial flow of identified particles is sensitive to the freeze-out condition, and rapidity
dependence of v1 is sensitive to nuclear EOS, and a number of quark scaling of
elliptic and higher-order harmonics v2 , v3 , etc. are sensitive to the transition from
collectivity dominated by QGP to collectivity dominated by hadronic matter. This
section summarizes these and other experimental measurements of collectivity at
low energies.

3.4.1 Energy Dependence of Collectivity

The information on radial flow is often represented by transverse radial flow velocity
β and kinetic freeze-out temperature Tkin , extracted by a blast-wave model fit to the
pT spectra of identified particles [149]:
# R ) * ) *
dN pT sinh ρ(r ) m T cosh ρ(r )
∝ r dr m T I0 × K1 (3.30)
pT dpT 0 Tkin Tkin

where m T is the transverse mass of a hadron, ρ(r ) = tanh−1 fi, and I0 and K 1 are
the modified Bessel functions. Hadrons are assumed to reach local thermalization at
the temperature Tkin and to move with a common transverse velocity β. The blast-
wave model describes simultaneously the measured pT spectra of π ± , K ± , p, and
p̄ from RHIC and the LHC energies [3, 150, 151] in all centrality ranges. The
extracted Tkin values show anti-correlation with extracted *β+ values as shown in
Fig. 3.10: higher values of *β+ correspond to lower values of Tkin and vice versa. The
steady decrease of Tkin from peripheral to central collisions suggests a longer lived

Fig. 3.10 The extracted Tkin 200


and *β+ from different Pb+Pb 2.76 TeV
collision energies and 180 Au+Au 200 GeV
Au+Au 62.4 GeV
centralities [3]
160 peripheral Au+Au 39 GeV

140
Tkin (MeV)

120

100

80
Au+Au 27 GeV
Au+Au 19.6 GeV central
60 Au+Au 11.5 GeV
Au+Au 7.7 GeV
40
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
〈β〉
162 H. Elfner et al.

Fig. 3.11 The slope of v1 at 0.03


mid-rapidity (dv1 /dy) as a
function of collision energy 0.02
in 10−40% central Au+Au
collisions for net-proton, 0.01

y=0
net-!, and net-K [152, 153]

dv1/dy|
0

−0.01 STAR BES-I


Net - proton
Net - Λ
−0.02 NICA Net - Kaon
FAIR

10 102
sNN (GeV)

fireball in more central collisions, while the increase of *β+ from peripheral to central
collisions suggests stronger radial flow in more central collisions. As a function of
collision energy, the *β+ shows a rapid increase at low energies (<5 GeV), then a
slight increase across the RHIC BES energies (7.7–39 GeV), after which it again
increases very slowly up to LHC energy [3].
Figure 3.11 summarizes collision energy dependence of directed flow slope rel-
ative to rapidity (dv1 /dy). The directed flow is sensitive to the expansion of the
created medium in the early stage of collisions as demonstrated in nuclear trans-
port and hydrodynamic models [154–156]. The appearance of local minimum of

dv1 /dy( sNN ) has been predicted to be a robust signature of first-order phase tran-
sition from quark matter to hadron matter [157, 158]. A hybrid calculation has
shown that the change in the dynamics from the hot and dense hydrodynamic
stage to the hadronic rescattering stage could influence the energy dependence of
directed flow and potentially hide the effects of changes in the EOS [159]. There-
fore, the direct flows of conserved charges, such as net-charge, net-strangeness,
or net-baryon number, are more sensitive probes of the phase transition. Here,
we define net-particle as the excess yield of a particle type over its anti-particle
type [152, 153]. On this basis, the directed flow of net-particle is defined as fol-
lows: v1X = r (y)v1 X̄ + [1 − r (y)]v1net−X , where X is particle type, X̄ represents the
corresponding anti-particle, and r (y) represents the ratio of particle to anti-particle
yield. Figure 3.11 presents the collision energy dependence of directed slope at mid-
rapidity (dv1 /dy) for net-proton, net-!, and net-kaon. The non-monotonic energy
dependence is observed for the net-baryon data, proton, and ! as shown in the plot,
which is consistent with the hydrodynamical calculations with first-order phase tran-
sition [157, 158]. The values of net-proton and net-! dv1 /dy are also consistent

within uncertainties. Below s N N < 20 GeV, there is a large divergence between
dv1 /dy of net-kaon and net-proton (net-!), whereas all three agree well at and
above 20 GeV. It suggests that the properties of the medium would be different
above/below 20 GeV. At lower beam energies, the signature of phase transition is
modified by other dynamical effects, such as baryon stopping and transportation.
3 Dynamical Evolution of Heavy-Ion Collisions 163

(a) 0.06 (b) STAR +BES


-
Ξ -Ξ
0.05 0.05 Λ-Λ
p-p -
+
0.04 K -K

v2(X)-v (X)
π+-π-
0 0.03

2
v2
0.02
CERES ALICE
−0.05 E877 STAR 0.01
E895 PHENIX
FOPI PHOBOS 0
NICA
NICA
−0.1 FAIR −0.01 FAIR

10 10 2 103 10 10 2
sNN (GeV) sNN (GeV)

Fig. 3.12 a The pT -integrated v2 in 20−30% most central collisions (or similar centrality) as
a function of collision energy. The data points are from E895 for protons; NA49 for pions; and
FOPI, E877, CERES, STAR, ALICE PHENIX, and PHOBOS for charged hadrons [162–174]. b
The v2 difference between particles and the corresponding anti-particles as a function of collision
energy [175–177]

The detailed centrality dependence study of dv1 /dy of net-particle could offer more
information, constrain model calculations, and possibly separate the signal of phase
transition. The dv1 /dy of φ mesons has also been measured by STAR [153], and
they are found to be larger than those of pions and kaons at and above 14.5 GeV.
More interestingly, the dv1 /dy of φ meson seems to increase sharply at 11.5 GeV.
The current data have large uncertainty. Within the large uncertainty of current data,

the AMPT model seems to reproduce the monotonic sNN dependence φ dv1 /dy of
φ [160]. This study is expected to be significantly improved with data from RHIC
BES program phase II. The previous measurements from AGS/SPS and earlier time
are not included here. In Sect. 4.3.2, a recent data from Ca+Ca to Au+Au over the
SIS energy range from 0.09 to 1.5 GeV/u is discussed in detail [161].
Figure 3.12 shows the pT -integrated elliptic flow and elliptic flow difference
between particles and anti-particles as a function of collision energy. Several theoreti-
cal studies suggest that a non-monotonic collision energy dependence of elliptic flow
could be observed around the “softest point of EOS” [178, 179], as there is a strong
drop in the speed of sound or reduction in the pressure of the created medium during
the dynamic evolution. Panel (a) of Fig. 3.12 shows the data from E895 for protons;
NA49 for pions; and FOPI, E877, CERES, STAR, ALICE PHENIX, and PHOBOS
for charged hadrons [162–174] for 20–30% or similar centrality. The elliptic flow

values are negative around sNN = 2–4 GeV, which is known as the “squeeze-out”
effect [164]: the spectators hinder the particles moving along the direction of impact
parameter, thus the most particles emit in the direction perpendicular to the reac-
tion plane. A steady increasing trend of elliptic as collision energy is observed from

sNN = 4−2760 GeV, and the decrease of v2 is faster for 3–7.7 GeV compared to
7.7–2760 GeV.
164 H. Elfner et al.

On the other hand, the pT differential elliptic flow of charged hadrons changes
little from RHIC BES to LHC energies [173]. Therefore, the increase of integral v2
is mainly due to the increase of mean pT which is expected from a larger radial flow.
As shown in [180], there is a non-trivial interplay between the contribution from the
hot and dense hydrodynamic stage and the contribution from the hadronic transport
evolution. Within the ideal hydrodynamic+UrQMD transport hybrid approach, it has
been demonstrated that the elliptic flow in the SPS energy regime is not very sensitive
to initial-state fluctuations and changes in the EOS [181]. The currently available data

over 4 < sNN < 2760 GeV do not show the non-monotonic behavior predicted by
the softening of the EOS for a system close to the critical temperature [178].
The elliptic flow of identified particles reveals more information of properties of
the system. The values of v2 for particles and the corresponding anti-particles show

significant differences at sNN < 20 GeV as shown in panel (b) of Fig. 3.12 [175–
177]. The difference is larger between baryons, protons, !, &, and corresponding
antibaryons. The number of constituent quark (NCQ) scaling is broken between par-
ticles and anti-particles below 20 GeV, whereas the scaling still holds separately for
particles and anti-particles down to 7.7 GeV, except for multi-strange hadrons and φ
meson at 11.5 and 7.7 GeV. The mechanism for the v2 splitting between particle and
anti-particle has been investigated in models [182–186]. A model calculation based
on hydrodynamics + UrQMD hadronic transport reproduces the proton data, indi-
cating the transport effect would be important in low energies, but it fails to describe
the meson data [182]. A calculation based only on hydrodynamics quantitatively
reproduces the π, K , and proton v2 data, but fails to reproduce the apparent order-
p
ing for strange and multi-strange baryons, i.e., #v2 > #v2! > #v2& > #v2' [183].
A Nambu–Jona-Lasino (NJL) model describes the data qualitatively, suggesting the
hadronic and/or partonic potential plays an important role [184, 185]. Future detailed
centrality dependence and high precise data of multi-strange hadrons from BES-II
program could better constrain and guide the model calculation.
One of the signals of QGP formation in high-energy heavy-ion collisions is the
NCQ scaling on elliptic flow, which has been used to support that the collectivity
is generated at the parton level. RHIC experiments measured elliptic flow for many
different types of identified particles, from copious produced π ± , K ± , p, and p̄,
to strange hadrons, multi-strange hadrons, φ mesons, and D 0 mesons in Au+Au

collisions at sNN = 200 GeV. The NCQ scaling is universal (holds within 10%)
for all measured identified particles except for pions which are more affected by the
resonance decays [187–190]. The multi-strange hadrons (& and ') and φ mesons
show a similar magnitude of elliptic flow as light hadrons in the intermediate pT
range of 2–5 GeV/c. As multi-strange hadrons and φ mesons are less sensitive to late
hadronic interactions compared to light hadrons, the apparent NCQ scaling suggests
that the majority of v2 is built-up during the partonic phase. The LHC experiments
observe a similar scaling behavior, although showing a somewhat larger deviation
than the RHIC top energy [191]. Naturally, as the collision energy goes down, the
conditions for the formation of QGP will not be satisfied below certain energy. Such
disappearance of the partonic collectivity pattern should be reflected by the energy
dependence of NCQ scaling of v2 . In particular, multi-strange hadrons and φ mesons
3 Dynamical Evolution of Heavy-Ion Collisions 165

0.1 7.7 GeV p +


+
11.5 GeV 19.6 GeV
Au+Au, 0-80% K
- 0
-sub EP -
Ks

0.05

v2/nq 0.1 27 GeV 39 GeV 62.4 GeV

0.05

0 0.5 1 1.5 20 0.5 1 1.5 2 0 0.5 1 1.5 2


(m -m0)/n (GeV/c2)
T q

Fig. 3.13 The Number-of-Constituent Quark (n q ) scaled elliptic flow as a function of (m T −



m 0 )/n q for 0–80% most central Au+Au collisions at s N N = 7.7–62.4 GeV for selected anti-
particles [176]

are better probes of the parton level interactions. Figure 3.13 shows the NCQ scaling

of v2 for 0–80% most central Au+Au collisions at sNN = 7.7–62.4 GeV from RHIC
BES phase I [176]. Here, we show the results for the anti-particles as they can only
be produced, instead of being transported, in collisions. Except for & and φ mesons
from 7.7 and 11.5 GeV collisions, the NCQ scaling of v2 works well down to collision
energy 7.7 GeV, although the measured pT range is limited due to limited statistics
at 7.7 and 11.5 GeV. The φ meson data point for 11.5 GeV and 7.7 is 2.3 and 1.8
σ lower than those of the other hadrons, respectively, at the highest (m T − m 0 )/n q
values. More precise measurements are expected to become available from the RHIC
BES-II program.

3.4.2 Collectivity in High Baryon Density Region

The fixed-target program (FXT) of RHIC BES phase II extends the coverage of
collisions energy down to 3 GeV. The main additional energies of FXT are 7.7 GeV,
6.2 GeV, 5.2 GeV, 4.5 GeV, 3.9 GeV, 3.5 GeV, 3.2, and 3 GeV. As the lowest energy
achieved, 3 GeV is extremely important to determine the possible phase boundary.
The left panel of Fig. 3.14 shows the recent measurement of the NCQ scaling on

v2 for π + , K + , and p from 10 to 40% mid-central Au+Au collisions at sNN = 3
GeV [192]. Colored dashed-lines indicate the fits by a polynomial function to the

NCQ-scaled v2 data at sNN = 200, 54.4, 27, 14.5, and 7.7 GeV with the colliding

mode [175, 193]. The same fit is also applied to the v2 data at sNN = 4.5 GeV
in the 0–30% Au+Au collisions from the fixed-target mode [194]. Results of v2
from 4.5 to 200 GeV are all positive and follow the NCQ scaling, although the
166 H. Elfner et al.

Au+Au Collisions (10-40%)


0.4
Au+Au Collisions at RHIC
0.08 200
54.4 p
0.3 STAR Λ
0.06 27
14.5 π

y=0
7.7 3 27 54.4 (GeV)
π 10-40% K
0.04 K 0.2

dv1/dy|
φ
v2 /nq
p

0.02
Mean-field Cascade
p
10-30%
π p 0.1 UrQMD
3 GeV
π-
0 4.5 GeV 0-30% +
K

−0.02 0

0 0.2 0.4 0.6 0.8 2 3 5 10 20 30


2
(mT - m0)/nq (GeV/c ) Collision Energy sNN (GeV)

Fig. 3.14 a The Number-of-Constituent Quark (n q ) scaled elliptic flow as a function of (m T −



m 0 )/n q for π + , K + , and p from Au+Au collisions in 10–40% centrality at sNN = 200–3 GeV.
Colored dash lines represent the fit to data in 200–4.5 GeV collisions from STAR experiment at
RHIC. b The slope of v1 at mid-rapidity (dv1 /dy) as a function of collision energy for p, !, π
(combined from π ± ), K (combined from K ± and K S0 ), and φ in heavy-ion collisions. The collision
centrality is 10–30% for 4.5 GeV and 10–40% for other energies [192]

strength of the collectivity becomes weaker as the colliding energy reduces. For 3
GeV collisions, it is apparent that all the values of v2 /n q are negative, which are
very different from high energies. Furthermore, the NCQ scaling seems to disappear
at 3 GeV, pointing to the creation of matter for which the partonic interactions no
longer dominate and the hadronic interactions are more important. The right panel
of Fig. 3.14 shows the collision energy dependence of dv1 /dy for π, K , p, !, and
φ mesons from Au+Au collisions for 10–40% centrality [192]. It is well known that
at top RHIC energy, all mid-rapidity dv1 /dy slopes are negative [195]; however,

at sNN = 3 GeV, they are all positive. This behavior suggests that the dominant
degrees of freedom at 3 GeV are nucleons. In addition, mesons are expected to flow
along with baryons in the high baryon density region since they are dominantly
produced through resonance decays. Experimental signatures for the transition from
partonic dominated to hadronic dominated and to baryonic dominated matter have
been discussed in Ref. [6] in terms of the ratios of K + /π + and net-particle dv1 /dy.
The production of light nuclei, such as d, t,3 He, and 4 He, is believed to be mainly
through the coalescence of produced or transported nucleons. Since the binding ener-
gies of light nuclei are small, they are expected to be mainly produced at a later stage
of the system evolution. The anisotropic flow of light nuclei is a unique probe of
the light-nuclei production mechanism and freeze-out properties. The STAR Col-
laboration has observed that all light nuclei v2 generally follow scaling by atomic

mass number in Au+Au collisions at sNN = 7.7–200 GeV [197]. This finding indi-
cates that the coalescence of nucleons is indeed the underlying mechanism of the
production of light nuclei in these collisions.
The left panel of Fig. 3.15 shows the light nucleus v1 scaled by atomic mass
as a function of pT in four different rapidity intervals from Au+Au collisions at

sNN = 3 GeV [196]. One can see that the v1 /A of all the light nuclei follow the
3 Dynamical Evolution of Heavy-Ion Collisions 167

Au+Au 3 GeV
0
10-40%
3 GeV Au+Au Collisions
-0.1 < y < 0 10-40%
0

−0.2 -0.2 < y< -0.1


(× 1.5)
−0.05
(a) -0.1 < y < 0 (b) -0.2 < y < -0.1

v2 / A
v1 / A

-0.3 < y < -0.2


(× 2.0) 0 p0.5 d
1 1.5t 2 0 0.5 1 1.5 2
3 4
−0.4
p 0.05
He He

d
t 0
3
−0.6 He -0.4 < y < -0.3
(× 2.5)
4
He −0.05 (c) -0.3 < y < -0.2 (d) -0.4 < y < -0.3

0 0.5 1 1.5 2 2.5


0 0.5 1 1.5 2 0 0.5 1 1.5 2
p / A (GeV/c)
p / A (GeV/c)
T
T

Fig. 3.15 The v1 and v2 as a function of pT with the atomic mass scale for p, d, t, 3 He, and 4 He
in four rapidity intervals in 10-40% mid-central Au+Au collisions [196]

A scaling at mid-rapidity of −0.3 < y < 0, supporting the coalescence mechanism.


The scaling is slightly broken in the rapidity window −0.4 to −0.3, which might be
due to contamination from the nuclear fragmentation of the target. The right panel
of Fig. 3.15 shows the v2 as a function of pT in different rapidity intervals from

Ref. [196]. The light nucleus v2 exhibits departure from A scaling at sNN = 3 GeV,

opposite to the observation in higher energies 7.7 < sNN < 200 GeV. JAM model
with the baryon mean field and coalescence of nucleons reproduce the v1 and v2
data of light nuclei qualitatively, indicating that baryonic interaction dominates the
collective dynamics at 3 GeV.
One important topic of low-energy heavy-ion collisions is their longitudinal
dynamics, in particular, their baryon stopping mechanism. Recently, HADES Col-
laboration carried out an impressive measurement of v1 , v2 , and higher-order flow
harmonics up to sixth order for proton, deuteron, and triton as a function of pT and
rapidity [198]. These flow harmonics show rich structures as a function of rapid-
ity as summarized in Fig. 3.16, namely the odd harmonics are odd functions of y,
while the even harmonics are even functions of y. All three species show similar
rapidity dependence trends, consistent with a picture in which the deuterons and
tritons are produced at a later stage through coalescing of protons and neutrons. The
sign-change pattern of the flow harmonics carries important information about the
baryon stopping mechanism and should be sensitive to EOS and hadronic transport.
The sign-change of v2 at large rapidity to positive values suggests a transition of
elliptic flow from squeeze out to bounce off, in a way very similar to one scan v2 as

a function of sNN to lower energy. Based on this detailed measurement, HADES
Collaboration reconstructs a three-dimensional representation of the baryon angular
emission pattern as shown in the right panel of Fig. 3.16. The fireball in momentum
168 H. Elfner et al.

Fig. 3.16 The left two columns show the rapidity dependence of vn for proton, deuteron, and triton
for n = 1–6 from HADES experiments. Each panel shows results for one harmonics. The cartoon
to the right illustrates a three-dimensional view of the proton emission pattern d N /dφ by summing
all the harmonics together as a function of rapidity. Figure revised from Ref. [198]

space evolves from a left-shifted pear-like shape in positive y, to a peanut-like shape


at y = 0, and to a right-shifted pear-like shape in negative y which is a reflection of
shape in positive y. Such very detailed measurements of anisotropic flow patterns
are extremely valuable for the tuning of the hydrodynamics and transport models.

3.4.3 Multi-particle Correlations and Flow Fluctuations

There are two major contributions to the fluctuations in the observed harmonic flow:
Initial-state fluctuations and thermal (hydrodynamical) fluctuations. In high-energy
heavy-ion collisions, flow fluctuations are dominated by initial-state fluctuations, and
the influence from hydrodynamical fluctuations has not yet been fully established.
Initial-state fluctuations are quantum fluctuations associated with the process
of nucleon–nucleon collisions and the subsequent energy deposition, on an event-
by-event basis. With these initial-state fluctuations, geometrical configurations of
the generated QGP fluctuate from event to event, giving rise to fluctuating spatial
anisotropies of the density distribution, !0 , !1 , !2 , !3 , etc. As a result of the well-
3 Dynamical Evolution of Heavy-Ion Collisions 169

known linear response, vn ∼ !n , initial-state fluctuations are expected to the source


for odd-order flow, such as triangular flow v3 and pentagonal flow v5 , etc. [199, 200],
as well as the non-trivial correlations among these flow harmonics [201, 202]. Since
initial-state fluctuations are introduced to the QGP medium at the very early stage, the
induced flow phenomena must be long-range in rapidity. As a result, effects of initial-
state fluctuations can be isolated by multi-particle correlations and/or by requiring a
large rapidity gap via the called sub-event method [200, 203–206], which efficiently
suppresses nonflow correlations. Since the strength of the initial-state fluctuations
can be crudely estimated . via the number of participant nucleons (or sources) in the
collision event N p , as 1/ N p [207, 208], initial-state fluctuations are expected to
be stronger in small collision systems. In high-energy heavy-ion collisions, a large
number of measurements have been accumulated, which allow us to investigate in
detail the properties of initial-state fluctuations. Deviations from the Gaussian dis-
tribution [209], for instance, have been measured and quantified through cumulants
of flow, vn {m} [210]. It is notable that, in all of the high-energy collisions, the non-
Gaussianity of the observed flow cumulants can be well understood in terms of
initial-state fluctuations [211].
In each collision event, given one specified QGP density profile, the evolution of
the medium suffers from thermal fluctuations, as has been described in Sect. 3.3.
These are known also as the hydrodynamical fluctuations. Unlike initial-state fluctu-
ations, whose strength depends on the number of sources of energy deposition from
the colliding nuclei, hydrodynamical fluctuations depend on the thermodynamical
properties of the QGP medium, such as temperature and dissipative corrections. At
an earlier stage, the QGP medium has higher temperatures and larger dissipation, and
the effects of hydrodynamical fluctuations are more pronounced. Also, the higher-
order flow harmonics are more affected by hydrodynamical fluctuations [212, 213]
since these higher flow harmonics are more sensitive to small-length scales of the
QGP medium. On the other hand, hydrodynamical fluctuations can be produced
throughout the whole space–time evolution of the QGP medium, and their contribu-
tion to the flow could be short-range in relative rapidity. Therefore, to some extent,
effects from hydrodynamical fluctuations on the flow observables have already been
suppressed by the rapidity-gap requirement in the experimental analysis.
At low collision energies, the initial state of the generated QGP medium is no
longer given at one particular proper time, but spans over space and time. When the
collision energy becomes sufficiently low, it is even possible that there is no distinct
separation between the initial state and the final state for the system evolution, so that
the flow response paradigm, vn ∼ !n , is expected to be no longer valid. However,
effects of initial-state fluctuations on flow observables must still exist, in a way
different from those at high collision energies. One may expect that the dominant
flow fluctuations are from the initial state, as long as the system evolution is far
away from the QCD critical region. Conversely, hydrodynamical fluctuations will be
important if the QCD medium is close to the QCD critical region.
170 H. Elfner et al.

3.5 Summary

Flow signatures in experiments are essentially consequences of the space–time


dynamics of the system created in heavy-ion experiments. Unlike the top RHIC
energies and the LHC where the system evolution is dominated by the paronic
stage, at low energies, contributions from the hadronic phase become substantial.
Accordingly, theoretical modeling should be improved. As discussed in detail in the
previous sections, it is either the hybrid model that contains a fluid-dynamics descrip-
tion accompanied by the subsequent hadronic transport models, such as UrQMD,
JAM, and SMASH, or a transport description that characterizes from partons to
hadrons, such as AMPT and PHSD, or purely a transport modeling for the hadron
or baryon scatterings at sufficiently low energies. For high-energy experiments, var-
ious theoretical modelings have shown that reproducing the observed flow requires
various ingredients, including initial geometrical configurations, the medium trans-
port properties captured by transport coefficients in QGP and hadron gas, EOS, and
particlization scheme [214]. These ingredients are of great physical significance, as
they influence the extracted properties of the QCD matter, in particular those related
to the QCD phase transition.
In low-energy collisions, flow observables are expected to be influenced by these
ingredients as well. Therefore, flow measurements can constrain parameters in theo-
retical modelings, and correspondingly the properties of the system. For the particle
spectra, for instance, the most sensitive parameters in theoretical modelings are the
initial system size and the freeze-out temperature. This explains the extracted Tkin
and *β+ from the blast-wave fit. The NCQ scaling of v2 can always be modeled as a
coalescence of flowing partonic medium, with effects of hadronic cascade properly
taken into account. Hence, the NCQ scaling of flow is a natural observable to dis-
tinguish the parton dominant system from the hadronic dominant system. Although
a consistent description of v2 from high to low energies can be achieved by tuning
parameters in theoretical modelings, some of the flow patterns are more subtle. The
change of slope of the proton direct flow dv1 /dy at low energies is not yet qualitatively
described by the existing models [159]. However, it is realized that the inclusion of
a softening EOS associated with first-order QCD phase transition, which effectively
gives rise to attractive scattering in the hadronic phase, is necessary in order to pro-
duce a negative slope. Interestingly, the softening effect in the EOS does not change
the prediction on v2 [79]. At high-energy collisions, higher-order flow harmonics are
more sensitive to the initial-state geometrical fluctuations and the medium transport
properties. However, at low energies, especially when the initial stage cannot be sep-
arable from subsequent expansions, the correlation between higher-order momentum
anisotropies and initial-state spatial anisotropies is suppressed [215].
Looking into the future, the recently completed BES-II program at RHIC has accu-

mulated high-statistics datasets from sNN = 3–200 GeV. The STAR experiment will

be able to improve all the flow measurements mentioned above, including the sNN ,
y, PID, and light nuclei with the strangeness sector, pT , and higher-order harmon-

ics. Together with the new data from HADAS at sNN < 3 GeV, they will provide
3 Dynamical Evolution of Heavy-Ion Collisions 171

a solid foundation on which the model can be tuned and improved. We also look
forward to new flow observables, traditionally not explored at low energy, such as
the multiple-particle flow cumulants, symmetric cumulants, mixed-harmonics [143],
higher-order pT fluctuations [216], vn – pT correlations [217], and longitudinal flow
decorrelations [148, 218]. Another important aspect of heavy-ion collisions is defor-
mations and radial structure of the atomic nuclei, and their influences on the initial
state and subsequent system evolution [219–224]. The MPD experiment at NICA is

expected to start data-taking in 2023 over the energy range of sNN = 3–11 GeV.
NICA has the ability to collide different species, which will provide an additional
handle on the longitudinal dynamics. All of these are expected to greatly improve
our understanding of the space–time dynamics of the heavy-ion collisions at low
energy and ultimately improve the constraints of the medium properties and QCD
phase diagram.
Heavy-ion collisions at O(10) GeV provide us with experimental access to study
hot nuclear matter in a high baryon density environment and quantify the phase struc-
ture of QCD matter. To establish definitive links between observables and structures
in the phase diagram, we need detailed dynamical modeling of all stages of heavy-ion
collisions. Because relativistic heavy-ion collisions go through a complex multi-stage
dynamics, fully integrated theoretical frameworks play a central role to provide reli-
able estimates of the dynamical evolution of the collisions and all relevant sources
of fluctuations. Different types of initial condition models have been developed to
consider the finite overlapping of the two nuclei at low collision energies. They open
a new venue to study the longitudinal dynamics of heavy-ion collisions. (3+1)D rela-
tivistic viscous hydrodynamics with multiple conserved charges is needed to study the
charge fluctuations and detailed hadronic chemistry in the high baryon density region.
Hadronic transport models with the inclusion of mean fields are essential. Searching
for the conjectured QCD critical point has been a main driving force for the Beam
Energy Scan program at RHIC and future experimental program at FAIR and NICA.
Quantifying the non-equilibrium evolution of stochastic fluctuations near the QCD
critical point remains a challenge to be incorporated into the full three-dimensional
dynamical frameworks. Further going beyond Gaussian fluctuations and theoretical
description of first-order phase transition requires more theoretical developments.

Acknowledgements This work is partly supported by the Chinese National Natural Science Foun-
dation under Grant Nos: 11861131009 and 12075098, and in part by the National Science Founda-
tion (NSF) under grant numbers PHY-2012922 and PHY-2012947 (Z.W.L.), and by the U.S. Depart-
ment of Energy, Office of Science, Office of Nuclear Physics, under contract numbers DE-SC001346
and DE-SC0021969, and within the framework of the Beam Energy Scan Theory (BEST) Topical
Collaboration, and Grants-in-Aid for Scientific Research from JSPS (JP21K03577). J.Y. Jia research
is supported in part by the U.S. Department of Energy, Office of Science, Office of Nuclear Physics,
under contract number DEFG0287ER40331, and by the National Science Foundation (NSF) under
grant number PHY-1913138. Y. Y. would like to acknowledge financial support from the Strate-
gic Priority Research Program of the Chinese Academy of Sciences, Grant No. XDB34000000.
M. Stephanov is supported by D.O.E with grant No. DE-FG0201ER41195. S.S. Shi would like
to acknowledge financial support from the National Natural Science Foundation of China under
Grant Nos. 11890710(11890711) and 12175084 and the National Key Research and Development
Program of China under Grant No. 2020YFE0202002. L.Y. is supported in part by the National
Natural Science Foundation of China under grant number 11975079.
172 H. Elfner et al.

References

1. Mohanty B (2011) STAR experiment results from the beam energy scan program at RHIC. J
Phys G 38:124023. arXiv: 1106.5902
2. Luo X, Xu N (2017) Search for the QCD critical point with fluctuations of conserved quantities
in relativistic heavy-ion collisions at RHIC: an overview. Nucl Sci Tech 28(8):112. arXiv:
1701.02105
3. Adamczyk L et al (2017) bulk properties of the medium produced in relativistic heavy-ion
collisions from the beam energy scan program. Phys Rev C 96(4):044904. arXiv: 1701.07065
4. Keane D (2017) The beam energy scan at the relativistic heavy ion collider. J Phys: Conf Ser
878(1):012015
5. Stephanov MA (2011) On the sign of kurtosis near the QCD critical point. Phys Rev Lett
107:052301. arXiv: 1104.1627
6. Bzdak A, Esumi S, Koch V, Liao J, Stephanov M, Xu N (2020) Mapping the phases of quantum
chromodynamics with beam energy scan. Phys Rep 853:1–87. arXiv: 1906.00936
7. Karpenko IA, Huovinen P, Petersen H, Bleicher M (2015) Estimation of the shear viscosity

at finite net-baryon density from A + A collision data at sNN = 7.7 − 200 GeV. Phys Rev
C 91(6):064901. arXiv: 1502.01978
8. Shen C, Schenke B (2018) Dynamical initial state model for relativistic heavy-ion collisions.
Phys Rev C 97(2):024907. arXiv: 1710.00881
9. Lin Z-W (2018) Extension of the Bjorken energy density formula of the initial state for
relativistic heavy ion collisions. Phys Rev C 98(3):034908. arXiv: 1704.08418
10. Gale C, Jeon S, McDonald S, Paquet J-F, Shen C (2019) Photon radiation from heavy-ion col-

lisions in the s N N = 19 − 200 GeV regime. Nucl Phys A 982:767–770. arXiv: 1807.09326
11. Mendenhall T, Lin Z-W (2021) Calculating the initial energy density in heavy ion collisions
by including the finite nuclear thickness. Phys Rev C 103(2):024907. arXiv: 2012.13825
12. Shen C, Yan L (2020) Recent development of hydrodynamic modeling in heavy-ion collisions.
Nucl Sci Tech 31(12):122. arXiv: 2010.12377
13. Denicol GS, Gale C, Jeon S, Monnai A, Schenke B, Shen C (2018) Net baryon diffusion
in fluid dynamic simulations of relativistic heavy-ion collisions. Phys Rev C 98(3):034916.
arXiv:1804.10557
14. Li M, Shen C (2018) longitudinal dynamics of high baryon density matter in high energy
heavy-ion collisions. Phys Rev C 98(6):064908. arXiv: 1809.04034
15. Mohs J, Ryu S, Elfner H (2020) Particle production via strings and baryon stopping within a
hadronic transport approach. J Phys G 47(6):065101. arXiv:1909.05586
16. Bjorken JD (1983) Highly relativistic nucleus-nucleus collisions: the central rapidity region.
Phys Rev D 27:140–151
17. Adcox K et al (2005) Formation of dense partonic matter in relativistic nucleus-nucleus
collisions at RHIC: experimental evaluation by the PHENIX collaboration. Nucl Phys A
757:184–283 nucl-ex/0410003
18. Oliinychenko D, Petersen H (2016) Deviations of the energy-momentum tensor from equi-
librium in the initial state for hydrodynamics from transport approaches. Phys Rev C
93(3):034905. arXiv: 1508.04378
19. Okai M, Kawaguchi K, Tachibana Y, Hirano T (2017) New approach to initializing hydrody-
namic fields and mini-jet propagation in quark-gluon fluids. Phys Rev C 95(5):054914. arXiv:
1702.07541
20. Lin Z-W, Ko CM, Li B-A, Zhang B, Pal S (2005) A Multi-phase transport model for relativistic
heavy ion collisions. Phys Rev C 72:064901. arXiv:nucl-th/0411110
21. Nara Y, Otuka N, Ohnishi A, Niita K, Chiba S (2000) Study of relativistic nuclear collisions at
AGS energies from p + Be to Au + Au with hadronic cascade model. Phys Rev C 61:024901.
arXiv:nucl-th/9904059
22. Weil J et al (2016) Particle production and equilibrium properties within a new hadron transport
approach for heavy-ion collisions. Phys Rev C 94(5):054905. arXiv:1606.06642
3 Dynamical Evolution of Heavy-Ion Collisions 173

23. Bass SA et al (1998) Microscopic models for ultrarelativistic heavy ion collisions. Prog Part
Nucl Phys 41:255–369. arXiv: nucl-th/9803035
24. Bleicher M et al (1999) Relativistic hadron hadron collisions in the ultrarelativistic quantum
molecular dynamics model. J Phys G 25:1859–1896 hep-ph/9909407
25. Cassing W, Bratkovskaya EL (2009) Parton-hadron-string dynamics: an off-shell transport
approach for relativistic energies. Nucl Phys A 831:215–242. arXiv: 0907.5331
26. Buss O, Gaitanos T, Gallmeister K, van Hees H, Kaskulov M, Lalakulich O, Larionov AB,
Leitner T, Weil J, Mosel U (2012) Transport-theoretical description of nuclear reactions. Phys
Rep 512:1–124 arXiv: 1106.1344
27. Pang L, Wang Q, Wang X-N (2012) Effects of initial flow velocity fluctuation in event-by-
event (3+1)D hydrodynamics. Phys Rev C 86:024911. arXiv: 1205.5019
28. Shen C, Denicol G, Gale C, Jeon S, Monnai A, Schenke B (2017) A hybrid approach
to relativistic heavy-ion collisions at the RHIC BES energies. Nucl Phys A 967:796–799.
arXiv:1704.04109
29. Du L, Heinz U, Vujanovic G (2019) Hybrid model with dynamical sources for heavy-ion
collisions at BES energies. Nucl Phys A 982:407–410. arXiv:1807.04721
30. Akamatsu Y, Asakawa M, Hirano T, Kitazawa M, Morita K, Murase K, Nara Y, Nonaka C,
Ohnishi A (2018) Dynamically integrated transport approach for heavy-ion collisions at high
baryon density. Phys Rev C 98(2):024909. arXiv:1805.09024
31. Shen C, Schenke B (2018) Initial state and hydrodynamic modeling of heavy-ion collisions
at RHIC BES energies. PoS CPOD2017:006. arXiv: 1711.10544
32. Bialas A, Bzdak A, Koch V (2018) Stopped nucleons in configuration space. Acta Phys Polon
B 49:103. arXiv:1608.07041
33. Shen C, Alzhrani S (2020) Collision-geometry-based 3D initial condition for relativistic
heavy-ion collisions. Phys Rev C 102(1):014909. arXiv:2003.05852
34. Kharzeev D (1996) Can gluons trace baryon number? Phys Lett B 378:238–246. arXiv:
nucl-th/9602027
35. Shen C, Schenke B (2019) Dynamical initialization and hydrodynamic modeling of relativistic
heavy-ion collisions. Nucl Phys A 982:411–414. arXiv:1807.05141
36. Wang X-N, Gyulassy M (1991) HIJING: a Monte Carlo model for multiple jet production in
p p, p A and A A collisions. Phys Rev D 44:3501–3516
37. Andersson B, Gustafson G, Ingelman G, Sjostrand T (1983) Parton fragmentation and string
dynamics. Phys Rep 97:31–145
38. Zhang B, Ko CM, Li B-A, Lin Z-W (2000) A multiphase transport model for nuclear collisions
at RHIC. Phys Rev C 61:067901. arXiv:nucl-th/9907017
39. Lin Z-W, Pal S, Ko CM, Li B-A, Zhang B (2001) Charged particle rapidity distributions at
relativistic energies. Phys Rev C 64:011902. arXiv: nucl-th/0011059
40. Zhang B (1998) ZPC 1.0.1: A Parton cascade for ultrarelativistic heavy ion collisions. Comput
Phys Commun 109:193–206. arXiv:nucl-th/9709009
41. Lin Z-W, Ko CM (2002) Partonic effects on the elliptic flow at RHIC. Phys Rev C 65:034904.
arXiv:nucl-th/0108039
42. Lin Z-W (2014) Evolution of transverse flow and effective temperatures in the parton phase
from a multi-phase transport model. Phys Rev C 90(1):014904
43. Pang L-G, Petersen H, Qin G-Y, Roy V, Wang X-N (2016) Decorrelation of anisotropic flow
along the longitudinal direction. Eur Phys J A 52(4):97. arXiv:1511.04131
44. Andersson B, Gustafson G, Soderberg B (1983) A general model for jet fragmentation. Z
Phys C 20:317
45. Petersen H, Steinheimer J, Burau G, Bleicher M, Stocker H (2008) A fully integrated trans-
port approach to heavy ion reactions with an intermediate hydrodynamic stage. Phys Rev C
78:044901. arXiv:0806.1695
46. Werner K (2007) Core-corona separation in ultra-relativistic heavy ion collisions. Phys Rev
Lett 98:152301. arXiv:0704.1270
47. Steinheimer J, Bleicher M (2011) Core-corona separation in the UrQMD hybrid model. Phys
Rev C 84:024905. arXiv:1104.3981
174 H. Elfner et al.

48. Kanakubo Y, Okai M, Tachibana Y, Hirano T (2018) Enhancement of strange baryons


in high-multiplicity proton–proton and proton–nucleus collisions. PTEP 2018(12):121D01.
arXiv:1806.10329
49. Kanakubo Y, Tachibana Y, Hirano T (2020) Unified description of hadron yield ratios from
dynamical core-corona initialization. Phys Rev C 101(2):024912. arXiv:1910.10556
50. Bertsch GF, Kruse H, Gupta SD (1984) Boltzmann equation for heavy ion collisions. Phys
Rev C 29:673–675. [Erratum: Phys. Rev. C 33, 1107–1108 (1986)]
51. Bertsch GF, Das Gupta S (1988) A Guide to microscopic models for intermediate-energy
heavy ion collisions. Phys Rep 160:189–233
52. Cassing W, Metag V, Mosel U, Niita K (1990) Production of energetic particles in heavy ion
collisions. Phys Rep 188:363–449
53. Welke GM, Prakash M, Kuo TTS, Das Gupta S, Gale C (1988) Azimuthal distributions in
heavy ion collisions and the nuclear equation of state. Phys Rev C 38:2101–2107
54. Gale C, Welke GM, Prakash M, Lee SJ, Das Gupta S (1990) Transverse momenta, nuclear
equation of state, and momentum-dependent interactions in heavy-ion collisions. Phys Rev
C 41:1545–1552
55. Aichelin J, Stoecker H (1986) Quantum molecular dynamics. A Novel approach to N body
correlations in heavy ion collisions. Phys Lett B 176:14–19
56. Aichelin J (1991) ‘Quantum’ molecular dynamics: a Dynamical microscopic n body approach
to investigate fragment formation and the nuclear equation of state in heavy ion collisions.
Phys Rep 202:233–360
57. Sorge H, Stoecker H, Greiner W (1989) Poincare invariant hamiltonian dynamics: modeling
multi - hadronic interactions in a phase space approach. Ann Phys 192:266–306
58. Sorge H (1995) Flavor production in Pb (160-A/GeV) on Pb collisions: effect of color ropes
and hadronic rescattering. Phys Rev C 52:3291–3314. arXiv:nucl-th/9509007
59. Maruyama T, Niita K, Maruyama T, Chiba S, Nakahara Y, Iwamoto A (1996)Relativistic
effects in the transverse flow in the molecular dynamics framework. Prog Theor Phys 96:263–
268. arXiv:nucl-th/9601010
60. Isse M, Ohnishi A, Otuka N, Sahu PK, Nara Y (2005) Mean-field effects on collective lows
in high-energy heavy-ion collisions from AGS to SPS energies. Phys Rev C 72:064908.
arXiv:nucl-th/0502058
61. Ko CM, Li Q, Wang R-C (1987) Relativistic Vlasov equation for heavy ion collisions. Phys
Rev Lett 59:1084–1087
62. Ko C-M, Li Q (1988) Relativistic Vlasov-Uehling-Uhlenbeck model for heavy-ion collisions.
Phys Rev C 37:2270–2273
63. Li Q, Wu JQ, Ko CM (1989) Relativistic Vlasov-Uehling-Uhlenbeck equation for nucleus-
nucleus collisions. Phys Rev C 39:849–852
64. Elze HT, Gyulassy M, Vasak D, Heinz H, Stoecker H, Greiner W (1987) Towards a relativistic
selfconsistent quantum transport theory of hadronic matter. Mod Phys Lett A 2:451–460
65. Blattel B, Koch V, Cassing W, Mosel U (1988) Covariant Boltzmann-Uehling-Uhlenbeck
approach for heavy-ion collisions. Phys Rev C 38:1767–1775
66. Blaettel B, Koch V, Mosel U (1993) Transport theoretical analysis of relativistic heavy ion
collisions. Rep Prog Phys 56:1–62
67. Fuchs C, Wolter HH (1995) The Relativistic Landau-Vlasov method in heavy ion collisions.
Nucl Phys A 589:732–756
68. Fuchs C, Lehmann E, Sehn L, Scholz F, Kubo T, Zipprich J, Faessler A (1996) Heavy ion
collisions and the density dependence of the local mean field. Nucl Phys A 603:471–485
69. Nara Y, Stoecker H (2019) Sensitivity of the excitation functions of collective flow to rela-
tivistic scalar and vector meson interactions in the relativistic quantum molecular dynamics
model RQMD.RMF. Phys Rev C 100(5):054902. arXiv:1906.03537
70. Nara Y, Maruyama T, Stoecker H (2020) Momentum-dependent potential and collective flows
within the relativistic quantum molecular dynamics approach based on relativistic mean-field
theory. Phys Rev C 102(2):024913. arXiv:2004.05550
3 Dynamical Evolution of Heavy-Ion Collisions 175

71. Li B-A, Ko CM (1998) Probing the softest region of nuclear equation of state. Phys Rev C
58:1382–1384. arXiv:nucl-th/9807088
72. Danielewicz P, Gossiaux PB, Lacey RA (1999) Hadronic transport model with a phase tran-
sition. Fundam Theor Phys 95:69–84. arXiv:nucl-th/9808013
73. Danielewicz P (1999) Nuclear phase transitions in transport theory. In: 27th international
workshop on the gross properties of nuclei and nuclear excitations (Hirschegg 99), pp 263–
272. arXiv:nucl-th/9902043
74. Sorensen A, Koch V (2021) Phase transitions and critical behavior in hadronic transport with a
relativistic density functional equation of state. Phys Rev C 104(3):034904. arXiv:2011.06635
75. Kim M, Jeon S, Kim Y-M, Kim Y, Lee C-H (2020) Extended parity doublet model with a
new transport code. Phys Rev C 101(6):064614. arXIv:2006.02023
76. Halbert EC (1981) Density patterns and energy-angle distributions from a simple cascade
scheme for last Ne-20 + U-238 collisions. Phys Rev C 23:295–330
77. Gyulassy M, Frankel KA, Stoecker H (1982) Do nuclei flow at high-energies? Phys Lett B
110:185–188
78. Kahana DE, Keane D, Pang Y, Schlagel T, Wang S (1995) Collective flow from the intranuclear
cascade model. Phys Rev Lett 74:4404–4407. arXiv:nucl-th/9405017
79. Nara Y, Niemi H, Ohnishi A, Stöcker H (2016) Examination of directed flow as a signa-
ture of the softest point of the equation of state in QCD matter. Phys Rev C94(3):034906.
arXiv:1601.07692
80. Sorge H (1999) Highly sensitive centrality dependence of elliptic flow: a novel signature of
the phase transition in QCD. Phys Rev Lett 82:2048–2051. arXiv: nucl-th/9812057
81. Nara Y, Niemi H, Steinheimer J, Stöcker H (2017) Equation of state dependence of directed
flow in a microscopic transport model. Phys Lett B 769:543–548. arXiv:1611.08023
82. Nara Y, Niemi H, Ohnishi A, Steinheimer J, Luo X, Stöcker H (2018) Enhancement of elliptic
flow can signal a first order phase transition in high energy heavy ion collisions. Eur Phys J
A 54(2):18. arXIv: 1708.05617
83. Nara Y, Steinheimer J, Stoecker H (2018) The enhancement of v4 in nuclear collisions
at the highest densities signals a first-order phase transition. Eur Phys J A 54(11):188.
arXiv:1809.04237
84. He Y, Lin Z-W (2017) Improved quark coalescence for a multi-phase transport model. Phys
Rev C 96(1):014910. arXiv: 1703.02673
85. Zhang L-Y, Chen J-H, Lin Z-W, Ma Y-G, Zhang S (2018) Two-particle angular correlations
in pp and p-Pb collisions at energies available at the CERN Large Hadron Collider from a
multiphase transport model. Phys Rev C 98(3):034912. arXiv:1808.10641
86. Borsanyi S, Fodor Z, Hoelbling C, Katz SD, Krieg S, Szabo KK (2014) Full result for the
QCD equation of state with 2+1 flavors. Phys Lett B 730:99–104. arXiv:1309.5258
87. Bazavov A et al (2014) Equation of state in ( 2+1 )-flavor QCD. Phys Rev D 90:094503.
arXiv:1407.6387
88. Monnai A, Schenke B, Shen C (2019) Equation of state at finite densities for QCD matter in
nuclear collisions. Phys Rev C 100(2):024907. arXiv:1902.05095
89. Noronha-Hostler J, Parotto P, Ratti C, Stafford JM (2019) Lattice-based equation of state
at finite baryon number, electric charge and strangeness chemical potentials. Phys Rev C
100(6):064910. arXiv: 1902.06723
90. Monnai A, Schenke B, Shen C (2021) QCD Equation of state at finite chemical potentials for
relativistic nuclear collisions. 1. arXiv:2101.11591
91. Denicol GS, Niemi H, Molnar E, Rischke DH (2012) Derivation of transient relativistic fluid
dynamics from the Boltzmann equation. Phys Rev D 85:114047. arXiv:1202.4551. [Erratum:
Phys. Rev. D 91, 039902 (2015)]
92. Denicol GS, Jeon S, Gale C (2014) Transport coefficients of bulk viscous pressure in the
14-moment approximation. Phys Rev C 90(2):024912 1403.0962
93. Greif M, Fotakis JA, Denicol GS, Greiner C (2018) Diffusion of conserved charges in rela-
tivistic heavy ion collisions. Phys Rev Lett 120(24):242301. arXiv:1711.08680
176 H. Elfner et al.

94. Rose J-B, Greif M, Hammelmann J, Fotakis JA, Denicol GS, Elfner H, Greiner C (2020)
Cross-conductivity: novel transport coefficients to constrain the hadronic degrees of freedom
of nuclear matter. Phys Rev D 101(11):114028. arXiv:2001.10606
95. Fotakis JA, Greif M, Greiner C, Denicol GS, Niemi H (2020) Diffusion processes involving
multiple conserved charges: a study from kinetic theory and implications to the fluid dynamical
modeling of heavy ion collisions. Phys Rev D 101(7):076007. arXiv:1912.09103
96. Ivanov YB, Russkikh VN, Toneev VD (2006) Relativistic heavy-ion collisions within 3-fluid
hydrodynamics: Hadronic scenario. Phys Rev C 73:044904. arXiv:nucl-th/0503088
97. Batyuk P, Blaschke D, Bleicher M, Ivanov YuB, Karpenko I, Merts S, Nahrgang M, Petersen
H, Rogachevsky O (2016) Event simulation based on three-fluid hydrodynamics for collisions
at energies available at the Dubna Nuclotron-based Ion Collider Facility and at the Facility
for Antiproton and Ion Research in Darmstadt. Phys Rev C 94:044917. arXiv:1608.00965
98. Petersen H (2014) Anisotropic flow in transport + hydrodynamics hybrid approaches. J Phys
G 41(12):124005 1404.1763
99. Pratt S (2014) Accounting for backflow in hydrodynamic-Boltzmann interfaces. Phys Rev C
89(2):024910 1401.0316
100. Oliinychenko D, Huovinen P, Petersen H (2015) Systematic investigation of negative cooper-
frye contributions in heavy ion collisions using coarse-grained molecular dynamics. Phys Rev
C 91(2):024906 1411.3912
101. Schwarz C, Oliinychenko D, Pang LG, Ryu S, Petersen H (2018) Different realizations of
Cooper–Frye sampling with conservation laws. J Phys G 45(1):015001. arXiv:1707.07026
102. Oliinychenko D, Koch V (2019) Microcanonical particlization with local conservation laws.
Phys Rev Lett 123(18):182302. arXiv:1902.09775
103. Oliinychenko D, Shi S, Koch V (2020) Effects of local event-by-event conservation
laws in ultrarelativistic heavy-ion collisions at particlization. Phys Rev C 102(3):034904.
arXiv:2001.08176
104. Vovchenko V, Koch V (2020) Particlization of an interacting hadron resonance gas with
global conservation laws for event-by-event fluctuations in heavy-ion collisions. 12 2020.
arXiv:2012.09954
105. Stoecker H, Greiner W (1986) High-energy heavy ion collisions: probing the equation of state
of highly excited hadronic matter. Phys Rept 137:277–392
106. Pang L-G, Zhou K, Su N, Petersen H, Stöcker H, Wang X-N (2018) An equation-of-state
meter of quantum chromodynamics transition from deep learning. Nat Commun 9(1):210.
arXiv:1612.04262
107. Du Y-L, Zhou K, Steinheimer J, Pang L-G, Motornenko A, Zong H-S, Wang X-N, Stöcker
H (2020) Identifying the nature of the QCD transition in relativistic collision of heavy nuclei
with deep learning. Eur Phys J C 80(6):516. arXiv:1910.11530
108. Steinheimer J, Pang L, Zhou K, Koch V, Randrup J, Stoecker H (2019) A machine learn-
ing study to identify spinodal clumping in high energy nuclear collisions. JHEP 12:122.
arXiv:1906.06562
109. Fupeng LI, Yongjia WANG, Qingfeng LI (2020) Using deep learning to study the equation
of state of nuclear matter. Nucl Phys Rev 37(4):825–832
110. Wang R, Ma Y-G,Wada R, Chen L-W, HeW-B, Liu H-L, Sun K-J (2020) Nuclear liquid-gas
phase transition with machine learning. Phys Rev Res 2(4):043202. arXiv:2010.15043
111. Huang Y, Pang L-G, Luo X, Wang X-N (2022) Probing criticality with deep learning in
relativistic heavy-ion collisions. Phys Lett B 827:137001. arXiv:2107.11828
112. Pratt S, Sangaline E, Sorensen P, Wang H (2015) Constraining the eq. of state of super-hadronic
matter from heavy-ion collisions. Phys Rev Lett 114:202301. arXiv:1501.04042
113. Stephanov MA, Rajagopal K, Shuryak EV (1998) Signatures of the tricritical point in QCD.
Phys Rev Lett 81:4816–4819. arXiv:hep-ph/9806219
114. Stephanov MA, Rajagopal K, Shuryak EV (1999) Event-by-event fluctuations in heavy ion
collisions and the QCD critical point. Phys Rev D 60:114028. arXiv:hep-ph/9903292
115. Berdnikov B, Rajagopal K (2000) Slowing out-of-equilibrium near the QCD critical point.
Phys Rev D 61:105017 hep-ph/9912274
3 Dynamical Evolution of Heavy-Ion Collisions 177

116. Grossi E, Soloviev A, Teaney D, Yan F (2020) Transport and hydrodynamics in the chiral
limit. Phys Rev D 102(1):014042. arXiv:2005.02885
117. Grossi E, Soloviev A, Teaney D, Yan F (2021) Soft pions and transport near the chiral critical
point. 1. arXiv:2101.10847
118. Akamatsu Y, Mazeliauskas A, Teaney D (2017) A kinetic regime of hydrodynamic fluctuations
and long time tails for a Bjorken expansion. Phys Rev C 95(1):014909. arXiv:1606.07742
119. Landau LD, Lifšic EM, Lifshitz EM, P L, Pitaevskii LP, Sykes JB, Kearsley MJ (1980)
Statistical physics: theory of the condensed state. Course of theoretical physics. Elsevier
Science
120. Kapusta JI, Muller B, Stephanov M (2012) Relativistic theory of hydrodynamic fluctuations
with applications to heavy ion collisions. Phys Rev C85:054906. arXiv:1112.6405
121. Akamatsu Y, Mazeliauskas A, Teaney D (2018) Bulk viscosity from hydrodynamic fluctua-
tions with relativistic hydrokinetic theory. Phys Rev C 97(2):024902. arXiv:1708.05657
122. Stephanov M, Yin Y (2018) Hydrodynamics with parametric slowing down and fluctuations
near the critical point. Phys Rev D98(3):036006. arXiv:1712.10305
123. Akamatsu Y, Teaney D, Yan F, Yin Y (2018) Transits of the QCD Critical Point. 1811:05081
124. Martinez M, Schäfer T (2019) Stochastic hydrodynamics and long time tails of an expanding
conformal charged fluid. Phys Rev C 99(5):054902. arXiv:1812.05279
125. An X, Başar G, Stephanov M, Yee H-U (2020) Fluctuation dynamics in a relativistic fluid
with a critical point. Phys Rev C 102(3):034901. arXiv:1912.13456
126. An X, Basar G, Stephanov M, Yee H-U (2020) Evolution of non-Gaussian hydrodynamic
fluctuations. 9. arXiv:2009.10742
127. Pratt S, Plumberg C (2019) Evolving charge correlations in a hybrid model with both hydro-
dynamics and hadronic boltzmann descriptions. Phys Rev C 99(4):044916. arXiv:1812.05649
128. Pratt S, Plumberg C (2020) Determining the diffusivity for light quarks from experiment.
Phys Rev C 102(4):044909. arXiv:1904.11459
129. Crossley M, Glorioso P, Liu H (2017) Effective field theory of dissipative fluids. JHEP
1709:095. arXiv:1511.03646
130. Crossley M, Glorioso P, Liu H (2017) Effective field theory of dissipative fluids (II): classical
limit, dynamical KMS symmetry and entropy current. JHEP 1709:096. arXiv:1701.07817
131. Liu H, Glorioso P (2018) Lectures on non-equilibrium effective field theories and fluctuating
hydrodynamics. PoS, TASI2017:008. arXiv:1805.09331
132. Chen-Lin X, Delacrétaz LV, Hartnoll SA (2019) Theory of diffusive fluctuations. Phys Rev
Lett 122(9):091602. arXiv:1811.12540
133. Delacretaz LV, Glorioso P (2020) Breakdown of diffusion on chiral edges. Phys Rev Lett
124(23):236802. arXiv:2002.08365
134. Mukherjee S, Venugopalan R, Yin Y (2015) Real time evolution of non-Gaussian cumulants
in the QCD critical regime. Phys Rev C 92(3):034912. arXiv:1506.00645
135. Yin Y (2018) The QCD critical point hunt: emergent new ideas and new dynamics. 11.
arXiv:1811.06519
136. Zurek WH (1996) Cosmological experiments in condensed matter systems. Phys Rep
276:177–221. arXiv:cond-mat/9607135
137. Mukherjee S, Venugopalan R, Yin Y (2016) Universal off-equilibrium scaling of critical
cumulants in the QCD phase diagram. Phys Rev Lett 117(22):222301. arXiv:1605.09341
138. Rajagopal K, Ridgway G, Weller R, Yin Y (2020) Understanding the out-of-equilibrium
dynamics near a critical point in the QCD phase diagram. Phys Rev D 102(9):094025.
arXiv:1908.08539
139. Du L, Heinz U, Rajagopal K, Yin Y (2020) Fluctuation dynamics near the QCD critical point.
4. arXiv:2004.02719
140. Martinez M, Schäfer T, Skokov V (2019) Critical behavior of the bulk viscosity in QCD. Phys
Rev D 100(7):074017. arXiv:1906.11306
141. Nahrgang M, Bluhm M, Schaefer T, Bass SA (2019) Diffusive dynamics of critical fluctuations
near the QCD critical point. Phys Rev D 99(11):116015. arXiv:1804.05728
178 H. Elfner et al.

142. Sogabe N, Yamamoto N, Yin Y (2021) Positive magnetoresistance induced by hydrodynamic


fluctuations in chiral media. 5. arXiv:2105.10271
143. Jia J (2014) Event-shape fluctuations and flow correlations in ultra-relativistic heavy-ion
collisions. J Phys G 41(12):124003 1407.6057
144. Heinz U, Snellings R (2013) Collective flow and viscosity in relativistic heavy-ion collisions.
Ann Rev Nucl Part Sci 63:123–151. arXiv:1301.2826
145. Bernhard JE, Scott Moreland J, Bass SA, Liu J, U Hei (2016)z. Applying Bayesian parameter
estimation to relativistic heavy-ion collisions: simultaneous characterization of the initial state
and quark-gluon plasma medium. Phys Rev C 94(2):024907. arXiv:1605.03954
146. Everett D et al (2020) Multi-system Bayesian constraints on the transport coefficients of QCD
matter. 11. arXiv:2011.01430
147. Nijs G, van der Schee W, Gürsoy U, Snellings R (2020) A transverse momentum differential
global analysis of heavy ion collisions. 10. arXiv:2010.15130
148. Jia J, Zhang C, Xu J (2020) Centrality fluctuations and decorrelations in heavy-ion collisions
in a Glauber model. Phys Rev Res 2(2):023319. arXiv:2001.08602
149. Schnedermann E, Sollfrank J, Heinz UW (1993) Thermal phenomenology of hadrons from
200A GeV S+S collisions. Phys Rev C 48:2462

150. Adamczyk L et al (2020) Bulk properties of the system formed in Au+Au collisions at sNN =
14.5 GeV. Phys Rev C 101(2):024905. arXiv:1908.03585
151. Abelev B et al (2013) Centrality dependence of π, K, p production in Pb-Pb collisions at

s N N = 2.76 TeV. Phys Rev C 88:044910. arXiv:1303.0737
152. Adamczyk L et al (2014) Beam-energy dependence of the directed flow of protons, antiprotons,
and pions in Au+Au collisions. Phys Rev Lett 112(16):162301 1401.3043
153. Adamczyk L et al (2018) Beam-energy dependence of directed flow of !, !, ¯ K ± , K s0 and φ
in Au+Au collisions. Phys Rev Lett 120(6):062301. arXiv:1708.07132
154. Heinz UW (2010) Relativistic heavy ion physics. Landolt- Bornstein data collection series,
vol 23(I)
155. Bass SA et al (1998) Microscopic models for ultrarelativistic heavy ion collisions. Prog Part
Nucl Phys 41:255
156. Bleicher M et al (1999) Relativistic hadron-hadron collisions in the ultra-relativistic quantum
molecular dynamics model (UrQMD). J Phys G 25:1859
157. Rischke D et al (1996) The phase transition to the quark-gluon plasma and its effect on
hydrodynamic flow. Heavy Ion Phys 1:209
158. Stocker H (2005) Collective flow signals the quark gluon plasma. Nucl Phys A 750:121
159. Steinheimer J, Auvinen J, Petersen H, Bleicher M, Stöcker H (2014) Examination of directed
flow as a signal for a phase transition in relativistic nuclear collisions. Phys Rev C 89(5):054913
1402.7236
160. Nayak K et al (2019) Energy dependence study of directed flow in Au+Au collisions using
an improved coalescence in a multiphase transport model. Phys Rev C 100:054903
161. Reisdorf W et al (2012) Systematics of azimuthal asymmetries in heavy ion collisions in the
1 A GeV regime. Nucl Phys A 876:1–60. arXiv:1112.3180
162. Pinkenburg C et al (1999) elliptic flow: transition from out-of-plane to in-plane emission in
Au+Au collisions. Phys Rev Lett 83:1295
163. Alt C et al (2003) Directed and elliptic flow of charged pions and protons in Pb+Pb collisions
at 40A and 158A GeV. Phys Rev C 68:034903
164. Andronic A et al (2005) Excitation function of elliptic flow in Au + Au collisions and the
nuclear matter equation of state. Phys Lett B 612:173
165. Braun-Munzinger P, Stachel J (1998) Dynamics of ultra-relativistic nuclear collisions with
heavy beams: an experimental overview. Nucl Phys A 638:3c
166. Appelshauser H (2002) New results from CERES. Nucl Phys A 698:253c

167. Aamodt K et al (2010) Elliptic flow of charged particles in Pb+Pb collisions at s N N = 2.76
TeV. Phys Rev Lett 105:252302
168. Voloshin SA, Poskanzer AM, Snellings R (2010) Collective phenomena in non-central nuclear
collisions. Relativistic heavy ion. Springer, Berlin/Heidelberg, Germany, pp 293–333
3 Dynamical Evolution of Heavy-Ion Collisions 179


169. Adams J et al (2005) Azimuthal anisotropy in Au+Au collisions at s N N = 200 GeV. Phys
Rev C 72:014904
170. Adler C et al (2002) Elliptic flow from two- and four-particle correlations in Au+Au collisions

at s N N = 130 GeV. Phys Rev C 66:034904
171. Adare A et al (2007) Scaling properties of azimuthal anisotropy in Au+Au and Cu+Cu colli-

sions at s N N = 200 GeV. Phys Rev Lett 98:162301
172. Alver B et al (2007) System size, energy, pseudorapidity, and centrality dependence of elliptic
flow. Phys Rev Lett 98:242302

173. Adamczyk L et al (2012) Inclusive charged hadron elliptic flow in Au+Au collisions at s N N
= 7.7–39 GeV. Phys Rev C 86:054908

174. Shi S (2013) Event anisotropy v2 in Au+Au collisions at s N N = 7.7–62.4 GeV with STAR.
Nucl Phys A 904-905:895c
175. Adamczyk L et al (2013) Observation of an energy-dependent difference in elliptic flow
between particles and antiparticles in relativistic heavy ion collisions. Phys Rev Lett
110:142301

176. Adamczyk L et al (2013) Elliptic flow of identified hadrons in Au + Au collisions at s N N
= 7.7–62.4 GeV. Phys Rev C 88:014902
177. Adamczyk L et al (2016) Centrality dependence of identified particle elliptic flow in relativistic

heavy-ion collisions at s N N = 7.7–62.4 GeV. Phys Rev C 93:014907
178. Kolb PF, Sollfrank J, Heinz U (2000) Anisotropic transverse flow and the quark-hadron phase
transition. Phys Rev C 62:054909
179. Sorge H (1999) Highly sensitive centrality dependence of elliptic flow: a novel signature of
the phase transition in QCD. Phys Rev Lett 82:2048

180. Auvinen J, Petersen H (2013) Evolution of elliptic and triangular flow as a function of s N N
in a hybrid model. Phys Rev C 88(6):064908 1310.1764
181. Petersen H, Bleicher M (2010) Eccentricity fluctuations in an integrated hybrid approach:
Influence on elliptic flow. Phys Rev C 81:044906. arXiv:1002.1003
182. Steinheimer J, Koch V, Bleicher M (2012) Hydrodynamics at large baryon densities: under-
standing proton versus anti-proton v2 and other puzzles. Phys Rev C 86:044903
183. Hatta Y, Monnai A, Xiao B-W (2015) Flow harmonics vn at finite density. Phys Rev D
92:114010
184. Xu J, Song T, Ko C, Li F (2014) Elliptic flow splitting as a probe of the QCD phase structure
at finite baryon chemical potential. Phys Rev Lett 112:012301
185. Liu H et al (2019) Isospin splitting of pion elliptic flow in relativistic heavy-ion collisions.
Phys Lett B 798:135002
186. Tu B, Shi S, Liu F (2019) Elliptic flow of transported and produced protons in Au+Au collisions
with the UrQMD model. Chin Phys C 43:054106
187. Adamczyk L et al (2016) Centrality and transverse momentum dependence of elliptic flow

of multistrange hadrons and φ Meson in Au + Au collisions at s N N = 200 GeV. Phys Rev
Lett 116:062301
188. Adamczyk L et al (2017) Measurement of D 0 azimuthal anisotropy at midrapidity in Au +

Au collisions at s N N = 200 GeV. Phys Rev Lett 118:212301
189. Luo X, Shi S, Xu N, Zhang Y (2020) A study of the properties of the QCD phase diagram in
high-energy nuclear collisions. Particles 3(2):278–307. arXiv:2004.00789
190. Shi S (2016) An experimental review on elliptic flow of strange and multistrange hadrons in
relativistic heavy ion collisions. Adv High Energy Phys 2016:1987432. arXiv:1607.04863
191. Bezverkhny Abelev B et al (2015) Elliptic flow of identified hadrons in Pb-Pb collisions at

sNN = 2.76 TeV. JHEP 06:190. arXiv:1405.4632

192. Abdallah M et al (2021) Disappearance of partonic collectivity in s N N = 3 GeV Au+Au
collisions at RHIC. 8. arXiv: 2108.00908
193. Dong X, Esumi S, Sorensen P, Xu N, Xu Z (2004) Resonance decay effects on anisotropy
parameters. Phys Lett B 597:328–332. arXiv: nucl-th/0403030

194. Adam J et al (2020) Flow and interferometry results from Au+Au collisions at s N N = 4.5
GeV. 7. arXiv:2007.14005
180 H. Elfner et al.

195. Snellings R et al (2000) Novel rapidity dependence of directed flow in high-energy heavy ion
collisions. Phys Rev Lett 84:2803–2805 nucl-ex/9908001

196. Abdallah M et al (2021) Light nuclei collectivity from sNN = 3 GeV Au+Au collisions at
RHIC. 12. arXiv:2112.04066

197. Adamczyk L et al (2016) Measurement of elliptic flow of light nuclei at s N N = 200, 62.4,
39, 27, 19.6, 11.5, and 7.7 GeV at the BNL relativistic heavy ion collider. Phys Rev C
94(3):034908. arXiv:1601.07052
198. Adamczewski-Musch J et al (2020) Directed, elliptic, and higher order flow harmonics of

protons, deuterons, and tritons in Au+Au collisions at s N N = 2.4GeV. Phys Rev Lett
125:262301. arXiv:2005.12217
199. Alver B, Roland G (2010) Collision geometry fluctuations and triangular flow in heavy-ion
collisions. Phys Rev C 81:054905. arXiv: 1003.0194. [Erratum: Phys. Rev. C 82, 039903
(2010)]
200. Aad G et al (2012) Measurement of the azimuthal anisotropy for charged particle production

in s N N = 2.76 TeV lead-lead collisions with the ATLAS detector. Phys Rev C 86:014907.
arXiv: 1203.3087

201. Aad G et al (2014) Measurement of event-plane correlations in s N N = 2.76 TeV lead-lead
collisions with the ATLAS detector. Phys Rev C 90(2):024905. arXiv:1403.0489
202. Adam J et al (2016) Correlated event-by-event fluctuations of flow harmonics in Pb-Pb col-

lisions at sNN = 2.76 TeV. Phys Rev Lett 117:182301. arXiv:1604.07663
203. Aad G et al (2014) Measurement of long-range pseudorapidity correlations and azimuthal

harmonics in s N N = 5.02 TeV proton-lead collisions with the ATLAS detector. Phys Rev
C 90(4):044906. arXiv:1409.1792
204. Jia J, Zhou M, Trzupek A (2017) Revealing long-range multiparticle collectivity in small
collision systems via subevent cumulants. Phys Rev C 96(3):034906. arXiv:1701.03830
205. Huo P, Gajdošová K, Jia J, Zhou Y (2018) Importance of non-flow in mixed-harmonic multi-
particle correlations in small collision systems. Phys Lett B 777:201–206. arXiv:1710.07567
206. Zhang C, Jia J, Xu J (2019) Non-flow effects in three-particle mixed-harmonic azimuthal
correlations in small collision systems. Phys Lett B 792:138–141. arXiv:1812.03536
207. Alver B et al (2008) Importance of correlations and fluctuations on the initial source eccen-
tricity in high-energy nucleus-nucleus collisions. Phys Rev C 77:014906. arXiv:0711.3724
208. Aaboud M et al (2018) Measurement of long-range multiparticle azimuthal correlations with
the subevent cumulant method in pp and p + Pb collisions with the ATLAS detector at the
CERN Large Hadron Collider. Phys Rev C 97(2):024904. arXiv:1708.03559
209. Aad G et al (2013) Measurement of the distributions of event-by-event flow harmonics in
lead-lead collisions at = 2.76 TeV with the ATLAS detector at the LHC. JHEP, 11:183.
arXiv:1305.2942
210. Chatrchyan S et al (2014) Measurement of higher-order harmonic azimuthal anisotropy in

PbPb collisions at s N N = 2.76 TeV. Phys Rev C 89(4):044906. arXiv:1310.8651
211. Yan L (2018) A flow paradigm in heavy-ion collisions. Chin Phys C 42(4):042001.
arXiv:1712.04580
212. Yan L, Grönqvist H (2016) Hydrodynamical noise and Gubser flow. JHEP 03:121. arXiv:
1511.07198
213. Sakai A, Murase K, Hirano T (2020) Rapidity decorrelation of anisotropic flow caused by
hydrodynamic fluctuations. Phys Rev C 102(6):064903. arXiv:2003.13496
214. Schenke B, Shen C, Tribedy P (2020) Running the gamut of high energy nuclear collisions.
Phys Rev C 102(4):044905. arXiv:2005.14682
215. Hillmann P, Steinheimer J, Bleicher M (2018) Directed, elliptic and triangular flow of protons
in Au+Au reactions at 1.23 A GeV: a theoretical analysis of the recent HADES data. J Phys
G 45(8):085101. arXiv:1802.01951
216. Giacalone G, Gardim FG, Noronha-Hostler J, Ollitrault J-Y (2021) Skewness of mean
transverse momentum fluctuations in heavy-ion collisions. Phys Rev C 103(2):024910.
arXiv:2004.09799
3 Dynamical Evolution of Heavy-Ion Collisions 181

217. Bozek P (2016) Transverse-momentum–flow correlations in relativistic heavy-ion collisions.


Phys Rev C 93(4):044908. arXiv:1601.04513
218. Jia J, Huo P (2014) Forward-backward eccentricity and participant-plane angle fluctuations
and their influences on longitudinal dynamics of collective flow. Phys Rev C 90(3):034915
1403.6077
219. Shou QY, Ma YG, Sorensen P, Tang AH, Videbæk F, Wang H (2015) Parameterization of
deformed nuclei for glauber modeling in relativistic heavy ion collisions. Phys Lett B 749:215–
220. arXiv:1409.8375
220. Giacalone G (2020) Observing the deformation of nuclei with relativistic nuclear collisions.
Phys Rev Lett 124(20):202301. arXiv:1910.04673
221. Xu H-j, Li H, Wang X, Shen C, Wang F (2021) Determine the neutron skin type by relativistic
isobaric collisions. Phys Lett B 819:136453. arXiv:2103.05595
222. Jia J (2022) Shape of atomic nuclei in heavy ion collisions. Phys Rev C 105(1):014905.
arXiv:2106.08768
223. Jia J (2022) Probing triaxial deformation of atomic nuclei in high-energy heavy ion collisions.
Phys Rev C 105(4):044905. arXiv:2109.00604
224. Zhang C, Jia J (2022) Evidence of quadrupole and octupole deformations in 96 Zr+96 Zr
and 96 Ru+96 Ru collisions at ultra-relativistic energies. Phys Rev Lett 128(2):022301.
arXiv:2109.01631

You might also like