1 Introduction
It is well known that a strong interaction binds quarks and gluons together to form hadrons such as protons and neutrons. The contemporary theory of strong interaction is governed by quantum chromodynamics (QCD), which is an SU(3) quantum gauge theory. The non-Abelian nature of QCD has important consequences such as color confinement at a low-energy scale and asymptotic freedom at a high-energy scale. Color confinement means that at low-energy scales, the color carriers (i.e., quarks and gluons) are always confined in color singlet hadrons; thus, no isolated quark and gluon can be observed. However, when the energy scale grows (e.g., when the temperature or the baryon density of the hadronic matter is increased), QCD undergoes a deconfinement phase transition, and quarks and gluons are liberated from the hadrons. When the energy scale is very high, the coupling constant of QCD becomes small and the system goes into the perturbative regime of QCD. In this regime, the coupling constant decreases with increasing energy scale, a phenomenon known as asymptotic freedom. Reliable perturbative calculation can apply in this regime.
In reality, the conditions for the deconfinement phase transition are difficult to achieve. Moreover, the confinement energy scale of QCD is approximately ΛQCD 200 MeV, which, in terms of temperature, is approximately Tc ΛQCD 1012 K. This high temperature may have once existed in the early universe (e.g., according to modern cosmology, this occurred immediately following the Big Bang) and can currently only be realized experimentally on earth by relativistic heavy-ion collisions. Current operating facilities of heavy-ion collisions include the Relativistic Heavy-Ion Collider (RHIC) at Brookhaven National Laboratory in the United States of America and the Large Hadron Collider (LHC) at the European Organization for Nuclear Research (CERN). RHIC has been operational since 2000 and its current top colliding energy for Au + Au collisions is
In addition to the abovementioned phenomena, in recent years, researchers have realized that relativistic heavy-ion collisions can also generate strong EM fields and fluid vorticity. More importantly, under these strong EM fields and vorticity, numerous intriguing macroscopic quantum phenomena may occur. These phenomena provide us opportunities to study the nontrivial chiral properties of quark-gluon matter, particularly those related to quantum anomaly, as well as the spin dynamics of QGP. Moreover, these phenomena are closely related to other subfields of physics, such as particle physics, condensed matter physics, astrophysics, and cold atomic physics, and thus give rise to a new interdisciplinary research area. Some review articles are already available, including Refs. [2-9]. In the following section, we introduce the EM field and vorticity that occur in heavy-ion collisions.
2 EM field and vorticity
Let us consider a noncentral collision between two nuclei. The collision geometry is depicted in Fig. 1. The z direction is along the motion of the projectile, the x direction is along the impact parameter b (from the target to the projectile); and the y direction is along
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F001.jpg)
where
This is a huge magnetic field, considerably larger than the squared masses of the electron and light quarks (u, d quarks), and thus may induce significant quantum effects in systems composed of electrons and light quarks. Moreover, this is the strongest known magnetic field in the current universe; it is several orders stronger than the surface magnetic fields of neutron stars, including magnetars (eB∼1014-1015 Gauss) [10]. The result in Eq. (1) is very rough. More advanced simulations can be performed using transport models such as HIJING, AMPT, UrQMD [11-26]. In such simulations, one can determine the positions and momenta of each charged particle before and after the collision and then use, for example, the Lienard–Wiechert formula to calculate the EM fields. The possible quantum correction to the Lienard–Wiechert formula can be estimated (which was found to be insignificant) [3, 16]. Many aspects of the EM field were studied through this approach, such as the event-by-event fluctuations of the strength and orientation of the EM fields [13, 15, 16], azimuthal correlation between the EM field and matter geometry [16, 17], EM fields in different collision systems [18, 22], and influence of the charge distribution of nucleons [16, 17] (please see the reviews [3, 4]). In Fig. 2, we show the impact parameter dependence of the EM fields computed using the HIJING model for Au + Au and Pb + Pb collisions at RHIC and LHC energies, respectively. It is seen that the strength of the fields is roughly proportional to the collision energy
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F002.jpg)
Let us consider once again a noncentral collision of energy
where A is the mass number of the nucleus. For RHIC Au + Au collisions at
where v is the flow velocity. From this definition, it is clear that the physical meaning of the vorticity is the local angular velocity of the fluid cell. In relativistic hydrodynamics, according to different physical contexts, different vorticities can be defined. The commonly used ones are the kinematic, temperature, and thermal vorticities. The kinematic vorticity is a natural generalization of the nonrelativistic vorticity:
where uμ=γ(1, v) is the flow four velocity. In many situations, it is more convenient to use its tensorial representation ωmn=(1/2)(∂vuμ-∂μuv), which is related to ωμ by ωμ=-(1/2)єμvρσuvωρσ. The temperature vorticity is defined as
where T is the temperature. The special property of the temperature vorticity is that, for an ideal neutral fluid, it satisfies the Carter–Lichnerowicz equation
where β=1/T is the inverse temperature. The importance of thermal vorticity relies on the fact that it characterizes the global equilibrium of a rotating fluid and determines the spin polarization of the constituent particles in the fluid at the global thermal equilibrium [29, 30]. We will discuss the spin polarization in detail in Sec. 5.
In Fig. 3, we present the numerical results of the nonrelativistic and relativistic kinematic vorticities in Au + Au collisions at
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F003.jpg)
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F004.jpg)
3 Chiral anomaly and transport phenomena
What are the consequences of strong EM fields and vorticity in heavy-ion collisions? During the past decade, many discussions have addressed this question and considerable interesting effects have been studied. Among the most intriguing effects are the quantum phenomena that are closely related to the spin dynamics of quarks. For massless fermions, these phenomena are also deeply related to the chiral anomaly of QCD and quantum electrodynamics (QED) and can be called anomalous chiral transports (ACTs). For a massive case, the spin polarization of hyperons by vorticity is a remarkable example. Of course, in general, both ACTs and spin polarization could occur with both massless and massive particles, but they manifest mostly with massless and massive particles, respectively. In this section, we focus on ACTs. The noticeable examples of ACTs are the chiral magnetic effect (CME), chiral vortical effects (CVEs), chiral separation effect (CSE), and chiral electric separation effect (CESE). We give a pedagogical discussion of the underlying mechanisms of the ACTs [39, 40].
Consider a massless Dirac fermion of charge e>0 in a strong constant magnetic field along the z direction. This is the usual Landau problem in quantum mechanics. The energy spectrum can be obtained by solving the Dirac equation, and the result is presented as Landau levels,
where n labels the Landau levels. The lowest Landau level (LLL), which corresponds to n=0, is special; see Fig. 5 (left). First, the LLL is gapless, whereas all the higher Landau levels are gapped by
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F005.jpg)
Now suppose an electric field is imposed in the same direction as the magnetic field; see Fig. 5 (right). Near the level crossing node pz=0, the downward moving particles can be easily flipped by the electric field to move upward, and thus some LH fermions are tuned to RH fermions. This is a typical spectral flow phenomenon. Therefore, NV=NR+NL, the total number of RH and LH fermions, is still conserved, whereas the difference NA=NR-NL is not. We can calculate the time derivative of NA in the following manner. Let
where eB/(2π) is the transverse density of state and V is the volume of the system. The electric force gives
or equivalently, dNV/dt=0 and dNA/dt=Ve2EB/(2π2). In differential forms, they yield
This is the well-known chiral or axial anomaly [41, 42]. We note that although we obtain Eq. (10) by considering the strong magnetic field so that only the LLL is occupied, the result is actually true for an arbitrary magnetic field, as the higher Landau levels are degenerate in chirality and do not contribute to Eq. (10).
With the previous preparation, we now remove the electric field and calculate the RH and LH currents along the magnetic field; see Fig. 6. A current is equal to the carrier density times the velocity of the constituent particles. For massless particles, the velocity is the speed of light such that
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F006.jpg)
where the minus sign is necessary because LH fermions move opposite to the direction of the magnetic field. We can rewrite Eq. (11) as
and
where we have defined the vector and axial chemical potentials as
In classical physics, the Larmor theorem establishes that the motion of a charged particle of mass m in a magnetic field is equivalent to the motion in a rotating frame with frequency eB/(2m). This suggests the existence of analogous effects to CME and CSE but induced by rotation or vorticity. Consider a massless particle in a rotating frame. The particle feels a Coriolis force
These are the vector and axial CVEs [47-50]. A more rigorous consideration shows that an additional term, namely, T2ω/6 exists in JA, which may be related to the global gravitational anomaly [51, 52].
In Fig. 2, we see that, in addition to the strong magnetic field, heavy-ion collisions also create a strong electric field because of the fluctuation of the proton distribution. In geometrically asymmetric collisions such as Cu + Au collisions, a strong electric field can also exist that points from the Au to the Cu nucleus with a strength comparable to the magnetic field [18, 53, 54]. The electric field can also lead to anomalous transport (i.e., the CESE [55]; see also the derivation in holographic models [56, 57] and discussion in Weyl semimetal [58]). The CESE is not directly related to the chiral anomaly and its appearance requires both P and C violations. The CESE represents an axial current along the direction of the electric field. Its expression for two flavor QCD up to leading-log accuracy is given by [59]
where Qe and QA are the charge and axial matrices in flavor space, and g is the strong coupling constant. Of course, in addition to the CESE, the electric field induces the Ohm current JV=σE, where σ is the electric conductivity, which, for QGP, is actually very high, meaning that the QGP is a good conducting matter [60].
Interesting collective modes emerge from the coupled evolution of the axial and vector charges through CME and CSE, vector CVE and axial CVE, or CESE and the usual Ohm’s law. For example, the continuity equations for vector and axial charges can be written in terms of RH and LH charges:
Substituting the CME and CSE expressions and considering small fluctuations in
where
eE | eB | ω | |
---|---|---|---|
JV | σ | ||
JA | |||
Collective mode | chiral electric wave | chiral magnetic wave | chiral vortical wave |
4 ACTs in heavy-ion collisions
ACTs have attracted considerable attention in many subfields of physics, including nuclear physics, particle physics, astrophysics, condensed matter physics, atomic physics, and quantum optics. For heavy-ion collisions, in particular, ACTs provide a valuable means to detect the possible P and CP violations of QCD at high temperatures. It is a well-known experimental fact that the strong interaction is P and CP invariant in vacuum, although QCD itself permits the existence of P and CP violating θ term. This lacks a natural explanation and is one of the main puzzles in contemporary physics. It has been proposed that in a high-temperature environment produced by heavy-ion collisions, metastable domains leading to P and CP violations could be produced through, for example, sphaleron-induced transition between gauge field vacua of different topological winding numbers [63-65]. In these domains, the interaction between gluons and quarks (through triangle anomaly) can induce chirality imbalance in quarks, which can be characterized by the parameter μA. Thus, the EM fields or vorticity exerting to these domains cause the CME, CVE, and CESE. Therefore, the detection of ACTs is highly demanded in heavy-ion collisions.
4.1 Experimental search of CME
Because the magnetic field is roughly perpendicular to the reaction plane, the CME would drive a current that finally causes a charge separation with respect to the reaction plane. However, the production of μA has strong spatial fluctuation (among the metastable P-violating domains) and event-by-event fluctuation such that the event-averaged CME-induced charge separation vanishes. What can be observed is the fluctuation of the charge separation. This can be done by designing appropriate hadronic observables. One commonly used observable is the γ correlation introduced by Voloshin [66]:
where α, β=± denote the charge signs, φα and φβ are the corresponding azimuthal angles, ΨRP is the reaction plane angle, and 〈⋅〉 is the event average. It is easy to see that a charge separation with respect to the reaction plane results in positive γ+- and γ-+ (denoted as γOS) and negative γ++ and γ– (denoted as γSS). In real experiments, one additional reference hadron (of arbitrary charge) is required to determine ΨRP. Therefore, Eq. (20) is practically a three-particle correlation.
The correlation γαβ was first measured by the STAR Collaboration at RHIC for Au + Au collisions at
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F007.jpg)
The main challenge remaining with the experiments is to disentangle the elliptic-flow-driven background effects and the magnetic-field-driven CME signal. One important experimental progress is the measurement of the γ correlation in small systems such as p(d) + A collisions. In p(d) + A collisions, although the magnetic field could be large, its orientation is not correlated to the participant plane (or v2 plane). Thus, the magnetic field is not expected to drive a strong γ correlation measured with respect to the v2 plane. Therefore, the p(d) + A collisions can serve as a baseline for the background contributions. The recent results from CMS [71, 72] and STAR [78] Collaborations showed that the γ correlation in p(d) + A collisions is comparable to or even larger than that in A + A collisions at the same energy and multiplicity. This suggests that the γ correlation contains a large portion of background contribution for peripheral A + A collisions; see additional discussions in Refs. [7, 71, 72, 78].
Another important experimental progress, namely, the isobar collision was made in 2018 at RHIC. In this experimental program, two sets of collisions are operated, one for
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F008.jpg)
Recently, other methods have been proposed for the purpose of disentangling the CME signal and the backgrounds. They include the pair invariant mass dependence of the γ correlation [86, 87], a comparative measurement of the γ correlation with respect to reaction and participant planes [87, 88], the signed balance functions [89], and the charge-sensitive in-event correlations [90]. A detailed discussion can be found in the cited studies.
4.2 Experimental search of other ACTs
The chiral magnetic wave can transport both the vector and axial charges and can lead to an electric quadrupole in the QGP with more positive charges on the tips of the fireball and more negative charges in the equator of the fireball [91, 92]. Therefore, hydrodynamic expansion of the fireball drives a larger v2 for negative charges (for example, π-) than the positive charges (for example, π+) [92-96]. The difference Δv2=v2(π-)-v2(π+) is proportional to the net charge asymmetry Ach=(N+-N-)/(N++N-); this is because the CSE is proportional to μV. This charge dependence of v2 was measured by the STAR Collaboration [97] at RHIC and by ALICE Collaboration [98] and CMS Collaboration [99] at LHC. Tv data show an elliptic-flow difference v2 linear in Ach with a positive slope whose centrality dependence is consistent with the expectation of the CMW. However, we should emphasize that, similar to the measurement of the γ correlation, non-CMW background effects exist, that contribute to Δv2 [15, 100-106]. A conclusive claim about the experimental results for the CMW search can be made only after we can successfully subtract the background effects, which we are unable to do now.
In heavy-ion collisions, the transverse space-averaged vorticity at the mid-rapidity region is roughly perpendicular to the reaction plane. Therefore, similar to the CME case, the vector CVE induces a baryon number separation with respect to the reaction plane. We can use a correlation similar to the γ correlation for CME to detect the vector-CVE-induced baryon number separation (i.e., ηαβ= (ϕα+ϕβ-2ΨRP), where α, β=± denote baryons or anti-baryons and ϕα,β is the corresponding azimuthal angle). However, similar to what occurs with the CME search, it would be challenging to subtract the possible background contributions as with the transverse momentum conservation and local baryon number conservation in the η correlation. The implication of the CVW in heavy-ion collisions is that it could induce a baryon quadrupole in the QGP in such a manner that more baryons and anti-baryons are distributed on the tips and in the equator of the fireball, respectively. After the collective expansion of the fireball, the baryons (for example, Λ) would have smaller v2 than the anti-baryons (
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F009.jpg)
The non-central Cu + Au collisions may be used to test the CESE, as they generate a persistent electric field orientating from the Au to Cu nuclei [18]. As illustrated in Fig. 10, the CESE induces an axial charge separation along the impact parameter direction (e.g., RH and LH chiralities on the near-Cu and near-Au sides, respectively), the CME in turn induces a charge separating pattern as shown in the last step (which is superposed by an Ohm-current-induced in-plane charge separation). A possible observable for this special quadrupolar pattern of charge distribution can be the charge dependence of the event planes, namely, a finite
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F010.jpg)
5 Spin polarization in heavy-ion collisions
A remarkable effect of vorticity is that it could polarize the spin of the constituent particles [109-112]. This is simply due to the quantum mechanical spin–orbit coupling. The motion of the fluid cell with finite vorticity generates an orbital angular momentum that can be transferred to the spin degree of freedom of the particles that constitute the fluid. If the system attains thermal equilibrium, we can use statistical mechanics to estimate the spin polarization. The density operator is
where nF(p0) with
Thus, we obtain the polarization vector in the rest frame of the particle as
In the following, without confusion, we simply use P to denote the polarization vector in the rest frame.
Before discussing the experimental measurements and numerical computations, let us explain the relation and distinction between the spin polarization of hyperons and ACTs in heavy-ion collisions. The ACTs are closely related to the chiral anomaly of QCD and/or QED, which is critical in modern physics. Detecting the ACTs also provides strong evidence for chiral symmetry restoration in the hot QGP. However, the underlying mechanism of the spin polarization is not related to the chiral anomaly but to quantum mechanical spin-orbit coupling. Importantly, spin polarization measurements provide a new probe for the QGP, that is, the spin probe, which is complementary to the usual probes using, for example, the charges. The ACTs and spin polarization of hyperons are also closely related to each other. First, they all represent the responses of the hot medium to the external vortical or EM field. In fact, as we will see in the following, the spin polarizations of Λ and
Substituting the theoretically calculated thermal vorticity shown in Fig. 4 into Eqs. (21)-(23), we can obtain the y component of the spin polarization. This reflects the global angular momentum of the collision system and is called the global spin polarization; see Fig. 11 for global spin polarization of Λ and
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F011.jpg)
Recently, the STAR Collaboration published the measurements of differential spin polarization, namely, the dependence of Λ polarization on the kinematic variables such as the azimuthal angle and transverse momentum [116, 128]. In describing the differential spin polarization, the theoretical calculations thus far have been unsatisfactory. In particular, the calculations based on hydrodynamic and transport models show that Py(ϕ) (ϕ: azimuthal angle) at mid-rapidity increases when ϕ grows from 0 to π/2. However, the experimental data show the opposite; see Fig. 12 [32]. Similarly, for noncentral collisions, a nonzero longitudinal Λ polarization Pz(ϕ) is observed in experiments (where this polarization vanishes when integrated over all the angles ϕ), indicating a ϕ dependence that is also qualitatively opposite to the theoretical calculations of thermal vorticity [32, 129, 130]; see Fig. 13. Expressed in formula as
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F012.jpg)
-202006/1001-8042-31-06-004/alternativeImage/1001-8042-31-06-004-F013.jpg)
the second-order harmonic coefficient f2z (and g2y) has the opposite sign in current theoretical calculations and in experimental data (i.e.,
To conclude this section, we explain how the special pattern of the thermal vorticity shown in Fig. 13 emerges. Although in Sec. 2 we discussed the fact that the global angular momentum of the collision system is the cause of vorticity, it is not the only cause. There are many other sources of vorticity. One important source is the inhomogeneous expansion of the fireball. Because in the non-central collisions, the fireball is almond shaped, the gradient of pressure would more strongly drive the fireball expanse along the reaction plane, and this is why we observe positive elliptic flow v2. In this type of expansion, we can easily imagine that a vortical structure with four vortices in four quadrants of the x-y plane (z=0) would appear. Of course, the temperature is also inhomogeneous and its gradient also contributes to the thermal vorticity, which together with the gradient of the velocity field gives the pattern shown in Fig. 13.
6 Spin hydrodynamics
Many attempts have been made to solve the spin sign problem. However, thus far, no satisfactory solution has been found. From a theoretical point of view, a key step forward would be to develop new theoretical frameworks to describe the spin polarization beyond the global equilibrium assumption. One promising framework is hydrodynamics, which very effectively describes the bulk evolution of the fireball in heavy-ion collisions, with the dynamical spin degree of freedom encoded. This type of framework is the spin hydrodynamics in which the spin polarization density (or equivalently the spin chemical potential) is treated on the same level as temperature T and flow velocity uμ [145-149].
In first-order spin hydrodynamics, the energy-momentum and spin current tensors are given by
in which we have chosen the Landau–Lifshitz frame. Here, e is the energy density, P is the pressure,
where X[αβ]=(Xαβ-Xβα)/2 is anti-symmetrization in indices α, β; X〈αβ〉=(Xαβ+Xβα)/2-Xμ μΔαβ/3 is traceless symmetrization in indices α, β; θ=∂μuμ is the expansion rate; Δμv=gμv-uμuv is the spatial projection; D=u·∂ is the co-moving time derivative;
To make the aforementioned equation close, we also need the equation of state, which links e,P,Sαβ.
In practical use, the aforementioned first-order theory has non-physical modes at the ultraviolet region, which violates the relativistic causality and leads to numerical instability. This problem stems from the constitutive relations Eqs. (26)-(29) that represent simple proportionality between the responses of the fluid (i.e., the LHSs) and the corresponding forces (i.e., the RHSs). The simplest means of overcoming this drawback of first-order hydrodynamics is to amend Eqs. (26)-(29) to the Israel-Stewart form:
where (⋅s)⊥ means taking the components transverse to uμ (e.g.,
7 Chiral and spin kinetic theories
In addition to hydrodynamics, kinetic theory is another commonly used method to study many-body systems in and out of equilibrium. Let us start with a short review of classical kinetic theory.
7.1 Classical kinetic theory
Classically, kinetic theory is built based on a single particle distribution function, which is a scalar function defined in the phase space. The physical meaning of the single particle distribution, which we denote as f(t,x,p), is the number of particles with a specific space location x and momentum p at time t. The kinetic equation determines the time evolution of f(t,x,p) and was first proposed by Boltzmann in the following form:
where u p/m is the single particle velocity with particle mass m, F is the external force, and C(t, x,p) is the collision term, which is a functional of f. The LHS of the aforementioned equation is the evolution of f due to streaming in the phase space with the existence of the external force field. In other words, the particle at the phase space point (x,p) moves with the velocity
Considering the special relativity, we can generalize Eq. (36) into a relativistic kinetic equation [136]. We adopt the Minkowski metric ημν=diag{1,-1,-1,-1} and the convention c=e=kB=1, and we define the eight-dimensional phase space coordinates as (x,p), where x=xμ=(t,x) and p=pμ=(p0,p), with p0 being the energy coordinate. Particles satisfy the following on-shell condition
where uμ pμ/p0 is the single particle four velocity and Fμ=(F0, F) is the four external force. The external force is called mechanical if it satisfies the condition
we reproduce the form of the LHS of Eq. (36) in the nonrelativistic kinetic representation.
The relations between physical quantities and the distribution function are readily obtained. The most elementary quantity is the particle density n(t,x), which is expressed as
where the delta function ensures that the particles are on-shell. Next, we consider the energy-momentum tensor. Classically, the energy-momentum tensor can be explained as the covariant current of the four momentum and thus reads as
The energy-momentum tensor is symmetric because the four velocity is proportional to the momentum uμ=pμ/p0. The entropy density is defined as
The entropy current satisfies the second law of thermodynamics (the Boltzmann H-theorem) ∂μsμ 0, where the equality holds in the global equilibrium state.
7.2 Wigner function in non-relativistic physics
When quantum mechanics is in action, the aforementioned kinetic theory must be modified. Quantum kinetic theory can be built based on the Wigner function method [152]. The Wigner function is the quantum correspondence of the classical distribution function first proposed by Wigner in 1932. In quantum mechanics, the properties of a particle are described by the wave function φ(t, x). The dynamics of a non-relativistic particle is governed by the Schrödinger equation:
where V=V(t, x) is the external potential. After the second quantization, we define the Wigner function as
where
The dynamics of the Wigner function are derived from the Schrödinger equation (42). Define
where we have integrated by parts. Next, we suppose that the gradient of the potential V is small so that we can make a gradient expansion. At the first order in ∂x, we have V(x+,t) - V(x-,t)=y·∂xV(x,t) and thus Eq. (44) reduces to
We thus identify the Wigner function as the single particle distribution function f(t,x,p)=W(t,x,p) and identify the external force F=-∂xV(x,t). Thus, Eq. (45) is reduced to the classical kinetic equation (36) without the collision term. To obtain the collision term, we must start with an interacting theory rather than the Schrödinger equation. The Wigner function method is particularly useful in performing the semiclassical approach to the quantum kinetic theory of spinful particles. Therefore, we next discuss quantum kinetic theory as related to spin-
7.3 Kinetic theory for Spin- fermions
With the aforementioned warmup preparation, we now consider the Dirac fermions. We not only introduce the Wigner function for the spinor field [153] but also review the derivation of the kinetic theory available in curved spacetime and the external EM field for Dirac fermions [115, 154-157]. In quantum field theory in Minkowski spacetime, the spin-
where the Dirac matrices satisfy γμ, γν=2gμν, ▽μψ=(∂μ + Γμ)ψ with the spin connection
Next, we establish the phase space in curved spacetime to introduce the Wigner function and the kinetic theory. We use the cotangent vector pμ to denote the momentum in curved spacetime with yμ as its conjugate variable. Thus, the momentum space is the cotangent space of the spacetime manifold at a given point. The local inner product of the momentum space and the spacetime manifold constitute the phase space, which is the cotangent bundle [158]. {yμ} constitutes the tangent space at a given point of the spacetime manifold, and the tangent bundle is locally the inner product of the tangent space and the spacetime manifold. We introduce the horizontal lifts of the covariant derivative in the cotangent bundle
The covariant Wigner operator under the U(1) gauge, local Lorentz transformation, and diffeomorphism are defined as [157]
with
with
where Rμν=Rρμρν is the Ricci tensor. We find that the spacetime curvature comes at O(ℏ2) at least. The Wigner function for the Dirac field is a 4 × 4 matric, which is different from the scalar case discussed in the previous subsection. Thus, the relation between the Wigner function and the semiclassical distribution function is less obvious in the spinor case. Equation (48) holds 16 scalar equations if we separate its matrix components, which can be decomposed into hermitian and antihermitian parts further.
Thus, we decompose the Wigner function based on Clifford algebra:
7.3.1 Chiral kinetic theory
For massless fermions, in the classical limit, not only Vμ but also Aμ is parallel to the momentum, and up to O(ℏ), they read as
where f=f(x, p) and f5=f5(x,p) are two scalar coefficients,
We define the right- and left-hand distribution functions as
where the mass-shell condition is corrected by the interaction between spin and the external EM field at O(ℏ). The flat spacetime version of the aforementioned chiral kinetic equation has been under intensive investigations recently [159-174], which can be written in the following form (for right-hand particles only) after p0 being integrated out:
where we have chosen nμ=(1,0,0,0),
Similarly, the kinetic equation and current for left-hand particles can be readily derived.
The kinetic theory in curved spacetime can be used to study the rotating frame. We consider the frame as rotating with the angular velocity Ω in the inertial frame, and we choose the frame vector nμ=(1, x×Ω). The kinetic equation reads as [142]
where
The equilibrium state can be derived from the kinetic equation (51). We suppose the local equilibrium distribution functions depend on a linear combination of the collisional conserved quantities: the particle number, energy and momentum, and angular momentum. Therefore, we have
where ϕ(x) is an arbitrary function that arises as a result of the conformal invariance in the massless case. We define the four velocity of the fluid as Uμ Tβμ with T being the temperature, and the chemical potential μR/L -TαR/L. Substituting the global equilibrium condition into Eqs. (53) and (55) and considering also the current of left-hand particles, we derive the CME and CSE as
where
We should note that the results for the CME and CVE currents are independent of the choice of the frame vector nμ.
7.3.2 Spin kinetic theory
For massive fermions, the particle spin is perpendicular to its momentum up to O(ℏ). The expressions of the vector and the axial vector are as follows:
where f=f(x, p) and fA=fA(x,p) are two scalar functions and θμ is the unit spacelike spin vector that is perpendicular to momentum pμθμ=0. We define
where
The evolution equation for the spin-direction vector θμ is given by [115]
We emphasize that the third term on the right-hand side is actually O(ℏ) order. From the kinetic equations, we can extract the Mathisson–Papapetrou–Dixon equations as
where τ is the proper time along the trajectory of the particle and dxμ/dτ=pμ/m. The previous two equations describe the spin dynamics for a single particle in curved spacetime and the external EM field.
We can derive the equilibrium state for massive fermions using the same method as in the massless case. Supposing
where we use
where
which, after integrating p0 over 0 to ∞, yields formula (21) for s=1/2 and with the contribution from the EM field added. (Note that in Eq. (21), the approximation
8 Summary
We discussed some intriguing properties of the strong EM fields and vorticity in heavy-ion collisions. We provided a heuristic introduction to the anomalous chiral transport phenomena and spin polarization in heavy-ion collisions. We briefly reviewed the recent progress in both theory and experiments toward understanding these novel quantum phenomena in heavy-ion collisions. The ACTs could be used to detect the nontrivial topological structure of the QCD gauge sector and the possible P and CP violations of strong interaction in a high-temperature environment. The spin polarization of hadrons provides us a probe to the (local) rotating properties and to the spin dynamics of the quark-gluon matter. This opens a door to a new era of subatomic spintronics.
Some challenges remain. Noticeably, the experimental observables for the ACTs (e.g., the CME) contain strong background contributions, which call for more efforts and new ideas from both the theoretical and experimental sides to be resolved. The experimental data for the azimuthal-angle dependence of spin polarization show a qualitatively opposite trend as compared to the thermal vorticity based on theoretical calculations, which gives rise to a spin sign problem. It is promising that new theoretical frameworks with spin as an independent dynamical variable may provide important insight into the spin sign problem. Presently, two of these frameworks, namely, spin hydrodynamics and spin kinetic theory, are progressing rapidly, and hopefully in the near future, the numerical simulations based on them could be achieved.
Chiral magnetic and vortical effects in high-energy nuclear collisions-A status report
. Prog. Part. Nucl. Phys. 88, 1-28 (2016). arXiv:1511.04050, https://doi.org/10.1016/j.ppnp.2016.01.001Electromagnetic fields and anomalous transports in heavy-ion collisions — A pedagogical review
. Rept. Prog. Phys. 79, 076302 (2016). arXiv:1509.04073, https://doi.org/10.1088/0034-4885/79/7/076302Novel quantum phenomena induced by strong magnetic fields in heavy-ion collisions
. Nucl. Sci. Tech. 28, 26 (2017). arXiv:1609.00747, https://doi.org/10.1007/s41365-016-0178-3Status of the chiral magnetic effect and collisions of isobars
. Chin. Phys. 41, 072001 (2017). arXiv:1608.00982, https://doi.org/10.1088/1674-1137/41/7/072001Experimental searches for the chiral magnetic effect in heavy-ion collisions
. Prog. Part. Nucl. Phys. 107, 200-236 (2019). arXiv:1906.11413, https://doi.org/10.1016/j.ppnp.2019.05.001Chiral Magnetic Effects in Nuclear Collisions
. arXiv:2002.10397Polarization and Vorticity in the Quark Gluon Plasma
. arXiv:2003.03640, https://doi.org/10.1146/annurev-nucl-021920-095245Vorticity and Spin Polarization — A Theoretical Perspective
. arXiv:2002.07549Magnetars
. Ann. Rev. Astron. Astrophys. 55, 261-301 (2017). arXiv:1703.00068, https://doi.org/10.1146/annurev-astro-081915-023329Estimate of the magnetic field strength in heavy-ion collisions
. Int. J. Mod. Phys. 24, 5925-5932 (2009). arXiv:0907.1396, https://doi.org/10.1142/S0217751X09047570(Electro-)Magnetic field evolution in relativistic heavy-ion collisions
. Phys. Rev. 83, 054911 (2011). arXiv:1103.4239, https://doi.org/10.1103/PhysRevC.83.054911Event-by-event fluctuations of magnetic and electric fields in heavy ion collisions
. Phys. Lett. 710, 171-174 (2012). arXiv:1111.1949, https://doi.org/10.1016/j.physletb.2012.02.065Magnetic effects in heavy-ion collisions at intermediate energies
. Phys. Rev. 84, 064605 (2011). arXiv:1107.3192, https://doi.org/10.1103/PhysRevC.84.064605Event-by-event generation of electromagnetic fields in heavy-ion collisions
. Phys. Rev. 85, 044907 (2012). arXiv:1201.5108, https://doi.org/10.1103/PhysRevC.85.044907Azimuthally fluctuating magnetic field and its impacts on observables in heavy-ion collisions
. Phys. Lett. 718, 1529-1535 (2013). arXiv:1209.6594, https://doi.org/10.1016/j.physletb.2012.12.030Charge-dependent azimuthal correlations from AuAu to UU collisions
. Nucl. Phys. 939, 85-100 (2015). arXiv:1311.5451, https://doi.org/10.1016/j.nuclphysa.2015.03.012Electric fields and chiral magnetic effect in Cu+Au collisions
. Phys. Lett. 742, 296-302 (2015). arXiv:1411.2733, https://doi.org/10.1016/j.physletb.2015.01.050A systematic study of magnetic field in Relativistic Heavy-ion Collisions in the RHIC and LHC energy regions
. Adv. High Energy Phys. 2014, 193039 (2014). arXiv:1408.5694, https://doi.org/10.1155/2014/193039Spatial distributions of magnetic field in the RHIC and LHC energy regions
. Chin. Phys. 39, 104105 (2015). arXiv:1410.6349, https://doi.org/10.1088/1674-1137/39/10/104105In search of chiral magnetic effect: separating flow-driven background effects and quantifying anomaly-induced charge separations
. Nucl. Phys. 956, 661-664 (2016). arXiv:1512.06602, https://doi.org/10.1016/j.nuclphysa.2016.01.064Test the chiral magnetic effect with isobaric collisions
. Phys. Rev. 94, 041901 (2016). arXiv:1607.04697, https://doi.org/10.1103/PhysRevC.94.041901Chiral magnetic effect in isobaric collisions
. Nucl. Phys. 967, 736-739 (2017). arXiv:1704.04382, https://doi.org/10.1016/j.nuclphysa.2017.05.071Electromagnetic fields in small systems from a multiphase transport model
. Phys. Rev. 97, 024910 (2018). arXiv:1709.05962, https://doi.org/10.1103/PhysRevC.97.024910Electromagnetic field effects on nucleon transverse momentum for heavy ion collisions
. Nucl. Sci. Tech. 28, 182 (2017). https://doi.org/10.1007/s41365-017-0337-1Electromagnetic field from asymmetric to symmetric heavy-ion collisions at 200 GeV/c
. Phys. Rev. 99, 054906 (2019). arXiv:1909.03160, https://doi.org/10.1103/PhysRevC.99.054906A study of vorticity formation in high energy nuclear collisions
. Eur. Phys. J. 75, 406 (2015). [Erratum: Eur. Phys. J.C78,no.5,354(2018)]. arXiv:1501.04468, https://doi.org/10.1140/epjc/s10052-015-3624-1, https://doi.org/10.1140/epjc/s10052-018-5810-4Vorticity in Heavy-Ion Collisions
. Phys. Rev. 93, 064907 (2016). arXiv:1603.06117, https://doi.org/10.1103/PhysRevC.93.064907Covariant statistical mechanics and the stress-energy tensor
. Phys. Rev. Lett. 108, 244502 (2012). arXiv:1201.5278, https://doi.org/10.1103/PhysRevLett.108.244502Relativistic distribution function for particles with spin at local thermodynamical equilibrium
. Annals Phys. 338, 32-49 (2013). arXiv:1303.3431, https://doi.org/10.1016/j.aop.2013.07.004Global Λ hyperon polarization in nuclear collisions: evidence for the most vortical fluid
. Nature 548, 62-65 (2017). arXiv:1701.06657, https://doi.org/10.1038/nature23004Thermal vorticity and spin polarization in heavy-ion collisions
. Phys. Rev. 99, 014905 (2019). arXiv:1810.00151, https://doi.org/10.1103/PhysRevC.99.014905Rotating quark-gluon plasma in relativistic heavy ion collisions
. Phys. Rev. 94, 044910 (2016). [Erratum: Phys. Rev.C95,no.4,049904(2017)]. arXiv:1602.06580, https://doi.org/10.1103/PhysRevC.94.044910, https://doi.org/10.1103/PhysRevC.95.049904Vorticity and hydrodynamic helicity in heavy-ion collisions in the hadron-string dynamics model
. Phys. Rev. 92, 014906 (2015). https://doi.org/10.1103/PhysRevC.92.014906Λ polarization in peripheral collisions at moderate relativistic energies
. Phys. Rev. 94, 054907 (2016). arXiv:1610.08678, https://doi.org/10.1103/PhysRevC.94.054907Vorticity in heavy-ion collisions at the JINR Nuclotron-based Ion Collider fAcility
. Phys. Rev. 95, 054915 (2017). arXiv:1701.01319, https://doi.org/10.1103/PhysRevC.95.054915Vorticity and hyperon polarization at energies available at JINR Nuclotron-based Ion Collider fAcility
. Phys. Rev. 97, 064902 (2018). arXiv:1801.07610, https://doi.org/10.1103/PhysRevC.97.064902Vorticity in low-energy heavy-ion collisions
. arXiv:2001.01371Notes on Anomaly Induced Transport
. Acta Phys. Polon. 47, 2617 (2016). arXiv:1610.04413, https://doi.org/10.5506/APhysPolB.47.2617Phenomenology of anomalous chiral transports in heavy-ion collisions
. EPJ Web Conf. 172, 01003 (2018). arXiv:1802.00172, https://doi.org/10.1051/epjconf/201817201003Axial vector vertex in spinor electrodynamics
. Phys. Rev. 177, 2426-2438 (1969). https://doi.org/10.1103/PhysRev.177.2426A PCAC puzzle: π0→γγ in the σ model
. Nuovo Cim. 60, 47-61 (1969). https://doi.org/10.1007/BF02823296The Effects of topological charge change in heavy ion collisions: ’Event by event P and CP violation’
. Nucl. Phys. 803, 227-253 (2008). arXiv:0711.0950, https://doi.org/10.1016/j.nuclphysa.2008.02.298The Chiral Magnetic Effect
. Phys. Rev. 78, 074033 (2008). arXiv:0808.3382, https://doi.org/10.1103/PhysRevD.78.074033Quantum anomalies in dense matter
. Phys. Rev. 70, 074018 (2004). arXiv:hep-ph/0405216, https://doi.org/10.1103/PhysRevD.70.074018Anomalous axion interactions and topological currents in dense matter
. Phys. Rev. 72, 045011 (2005). arXiv:hep-ph/0505072, https://doi.org/10.1103/PhysRevD.72.045011MACROSCOPIC PARITY VIOLATING EFFECTS: NEUTRINO FLUXES FROM ROTATING BLACK HOLES AND IN ROTATING THERMAL RADIATION
. Phys. Rev. 20, 1807-1812 (1979). https://doi.org/10.1103/PhysRevD.20.1807Fluid dynamics of R-charged black holes
. JHEP 01, 055 (2009). arXiv:0809.2488, https://doi.org/10.1088/1126-6708/2009/01/055Hydrodynamics from charged black branes
. JHEP 01, 094 (2011). arXiv:0809.2596, https://doi.org/10.1007/JHEP01(2011)094Hydrodynamics with Triangle Anomalies
. Phys. Rev. Lett. 103, 191601 (2009). arXiv:0906.5044, https://doi.org/10.1103/PhysRevLett.103.191601Gravitational Anomaly and Transport
. Phys. Rev. Lett. 107, 021601 (2011). arXiv:1103.5006, https://doi.org/10.1103/PhysRevLett.107.021601Global Anomalies, Discrete Symmetries, and Hydrodynamic Effective Actions
. JHEP 01, 043 (2019). arXiv:1710.03768, https://doi.org/10.1007/JHEP01(2019)043Estimation of electric conductivity of the quark gluon plasma via asymmetric heavy-ion collisions
. Phys. Rev. 90, 021903 (2014). arXiv:1211.1114, https://doi.org/10.1103/PhysRevC.90.021903Charge-dependent directed flow in asymmetric nuclear collisions
. Phys. Rev. 90, 064903 (2014). arXiv:1410.1402, https://doi.org/10.1103/PhysRevC.90.064903Axial Current Generation from Electric Field: Chiral Electric Separation Effect
. Phys. Rev. Lett. 110, 232302 (2013). arXiv:1303.7192, https://doi.org/10.1103/PhysRevLett.110.232302Holographic Chiral Electric Separation Effect
. Phys. Rev. 89, 085024 (2014). arXiv:1401.6972, https://doi.org/10.1103/PhysRevD.89.085024Holographic Charged Fluid with Chiral Electric Separation Effect
. JHEP 09, 083 (2018). arXiv:1803.08389, https://doi.org/10.1007/JHEP09(2018)083Chiral electric separation effect in Weyl semimetals
. Phys. Rev. 98, 165205 (2018). arXiv:1803.00723, https://doi.org/10.1103/PhysRevB.98.165205Chiral electric separation effect in the quark-gluon plasma
. Phys. Rev. 91, 045001 (2015). arXiv:1409.6395, https://doi.org/10.1103/PhysRevD.91.045001Thermal dilepton rates and electrical conductivity of the QGP from the lattice
. Phys. Rev. 94, 034504 (2016). arXiv:1604.06712, https://doi.org/10.1103/PhysRevD.94.034504Chiral Magnetic Wave
. Phys. Rev. 83, 085007 (2011). arXiv:1012.6026, https://doi.org/10.1103/PhysRevD.83.085007Chiral vortical wave and induced flavor charge transport in a rotating quark-gluon plasma
. Phys. Rev. 92, 071501 (2015). arXiv:1504.03201, https://doi.org/10.1103/PhysRevD.92.071501Possibility of spontaneous parity violation in hot QCD
. Phys. Rev. Lett. 81, 512-515 (1998). arXiv:hep-ph/9804221, https://doi.org/10.1103/PhysRevLett.81.512Anomalous chirality fluctuations in the initial stage of heavy ion collisions and parity odd bubbles
. Phys. Lett. 545, 298-306 (2002). arXiv:hep-ph/0109253, https://doi.org/10.1016/S0370-2693(02)02630-8Parity violation in hot QCD: Why it can happen, and how to look for it
. Phys. Lett. 633, 260-264 (2006). arXiv:hep-ph/0406125, https://doi.org/10.1016/j.physletb.2005.11.075Parity violation in hot QCD: How to detect it
. Phys. Rev. 70, 057901 (2004). arXiv:hep-ph/0406311, https://doi.org/10.1103/PhysRevC.70.057901Azimuthal Charged-Particle Correlations and Possible Local Strong Parity Violation
. Phys. Rev. Lett. 103, 251601 (2009). arXiv:0909.1739, https://doi.org/10.1103/PhysRevLett.103.251601Observation of charge-dependent azimuthal correlations and possible local strong parity violation in heavy ion collisions
. Phys. Rev. 81, 054908 (2010). arXiv:0909.1717, https://doi.org/10.1103/PhysRevC.81.054908Charge separation relative to the reaction plane in Pb-Pb collisions at sNN=2.76 TeV
. Phys. Rev. Lett. 110, 012301 (2013). arXiv:1207.0900, https://doi.org/10.1103/PhysRevLett.110.012301Constraining the magnitude of the Chiral Magnetic Effect with Event Shape Engineering in Pb-Pb collisions at sNN = 2.76 TeV
. Phys. Lett. 777, 151-162 (2018). arXiv:1709.04723, https://doi.org/10.1016/j.physletb.2017.12.021Observation of charge-dependent azimuthal correlations in p-Pb collisions and its implication for the search for the chiral magnetic effect
. Phys. Rev. Lett. 118, 122301 (2017). arXiv:1610.00263, https://doi.org/10.1103/PhysRevLett.118.122301Constraints on the chiral magnetic effect using charge-dependent azimuthal correlations in pPb and PbPb collisions at the CERN Large Hadron Collider
. Phys. Rev. 97, 044912 (2018). arXiv:1708.01602, https://doi.org/10.1103/PhysRevC.97.044912Beam-energy dependence of charge separation along the magnetic field in Au+Au collisions at RHIC
. Phys. Rev. Lett. 113, 052302 (2014). arXiv:1404.1433, https://doi.org/10.1103/PhysRevLett.113.052302Effects of Momentum Conservation and Flow on Angular Correlations at RHIC
. Phys. Rev. 84, 024909 (2011). arXiv:1011.6053, https://doi.org/10.1103/PhysRevC.84.024909Azimuthal correlations from transverse momentum conservation and possible local parity violation
. Phys. Rev. 83, 014905 (2011). arXiv:1008.4919, https://doi.org/10.1103/PhysRevC.83.014905Charge conservation at energies available at the BNL Relativistic Heavy Ion Collider and contributions to local parity violation observables
. Phys. Rev. 83, 014913 (2011). arXiv:1009.4283, https://doi.org/10.1103/PhysRevC.83.014913Effects of Cluster Particle Correlations on Local Parity Violation Observables
. Phys. Rev. 81, 064902 (2010). arXiv:0911.1482, https://doi.org/10.1103/PhysRevC.81.064902Charge-dependent pair correlations relative to a third particle in p + Au and d+ Au collisions at RHIC
. Phys. Lett. 798, 134975 (2019). arXiv:1906.03373, https://doi.org/10.1016/j.physletb.2019.134975Testing the Chiral Magnetic Effect with Central U+U collisions
. Phys. Rev. Lett. 105, 172301 (2010). arXiv:1006.1020, https://doi.org/10.1103/PhysRevLett.105.172301Importance of isobar density distributions on the chiral magnetic effect search
. Phys. Rev. Lett. 121, 022301 (2018). arXiv:1710.03086, https://doi.org/10.1103/PhysRevLett.121.022301Multiphase transport model predictions of isobaric collisions with nuclear structure from density functional theory
. Phys. Rev. 98, 054907 (2018). arXiv:1808.06711, https://doi.org/10.1103/PhysRevC.98.054907Chiral kinetic approach to the chiral magnetic effect in isobaric collisions
. Phys. Rev. 98, 014911 (2018). arXiv:1803.06043, https://doi.org/10.1103/PhysRevC.98.014911Examination of the observability of a chiral magnetically driven charge-separation difference in collisions of the 4496Ru+ 4496Ru and 4096Zr+ 4096Zr isobars at energies available at the BNL Relativistic Heavy Ion Collider
. Phys. Rev. 98, 061902 (2018). https://doi.org/10.1103/PhysRevC.98.061902Signatures of Chiral Magnetic Effect in the Collisions of Isobars
. arXiv:1910.14010Methods for a blind analysis of isobar data collected by the STAR collaboration
. arXiv:1911.00596Isolating the chiral magnetic effect from backgrounds by pair invariant mass
. Eur. Phys. J. 79, 168 (2019). arXiv:1705.05410, https://doi.org/10.1140/epjc/s10052-019-6671-1Measurements of the chiral magnetic effect with background isolation in 200 GeV Au+Au collisions at STAR
. Nucl. Phys. 982, 535-538 (2019). arXiv:1807.09925, https://doi.org/10.1016/j.nuclphysa.2018.08.035Varying the chiral magnetic effect relative to flow in a single nucleus-nucleus collision
. Chin. Phys. 42, 084103 (2018). arXiv:1710.07265, https://doi.org/10.1088/1674-1137/42/8/084103Probe Chiral Magnetic Effect with Signed Balance Function
. arXiv:1903.04622New correlator to detect and characterize the chiral magnetic effect
. Phys. Rev. 97, 061901 (2018). arXiv:1710.01717, https://doi.org/10.1103/PhysRevC.97.061901Normal ground state of dense relativistic matter in a magnetic field
. Phys. Rev. 83, 085003 (2011). arXiv:1101.4954, https://doi.org/10.1103/PhysRevD.83.085003Chiral magnetic wave at finite baryon density and the electric quadrupole moment of quark-gluon plasma in heavy ion collisions
. Phys. Rev. Lett. 107, 052303 (2011). arXiv:1103.1307, https://doi.org/10.1103/PhysRevLett.107.052303From the chiral magnetic wave to the charge dependence of elliptic flow
. arXiv:1208.2537Chiral magnetic wave in an expanding QCD fluid
. Phys. Rev. 91, 024902 (2015). arXiv:1310.0193, https://doi.org/10.1103/PhysRevC.91.024902Realistic Implementation of Chiral Magnetic Wave in Heavy Ion Collisions
. Phys. Rev. 89, 044909 (2014). arXiv:1311.2574, https://doi.org/10.1103/PhysRevC.89.044909The chiral magnetic effect in heavy-ion collisions from event-by-event anomalous hydrodynamics
. arXiv:1412.0311Observation of charge asymmetry dependence of pion elliptic flow and the possible chiral magnetic wave in heavy-ion collisions
. Phys. Rev. Lett. 114, 252302 (2015). arXiv:1504.02175, https://doi.org/10.1103/PhysRevLett.114.252302Charge-dependent flow and the search for the chiral magnetic wave in Pb-Pb collisions at sNN= 2.76 TeV
. Phys. Rev. 93, 044903 (2016). arXiv:1512.05739, https://doi.org/10.1103/PhysRevC.93.044903Probing the chiral magnetic wave in pPb and PbPb collisions at sNN=5.02TeV using charge-dependent azimuthal anisotropies
. Phys. Rev. 100, 064908 (2019). arXiv:1708.08901, https://doi.org/10.1103/PhysRevC.100.064908Charged elliptic flow at zero charge asymmetry
. Phys. Rev. 88, 014908 (2013). arXiv:1304.6410, https://doi.org/10.1103/PhysRevC.88.014908Constituent quark scaling violation due to baryon number transport
. Phys. Rev. 84, 044914 (2011). arXiv:1107.3078, https://doi.org/10.1103/PhysRevC.84.044914Effects of hadronic potentials on elliptic flows in relativistic heavy ion collisions
. Phys. Rev. 85, 041901 (2012). arXiv:1201.3391, https://doi.org/10.1103/PhysRevC.85.041901Partonic mean-field effects on matter and antimatter elliptic flows
. Nucl. Phys. 928, 234-246 (2014). arXiv:1211.5511, https://doi.org/10.1016/j.nuclphysa.2014.05.016Contributions to the event-by-event charge asymmetry dependence for the elliptic flow of pi+ and pi- in heavy-ion collisions
. Phys. Lett. 726, 239-243 (2013). arXiv:1303.1138, https://doi.org/10.1016/j.physletb.2013.08.003Elliptic flow difference of charged pions in heavy-ion collisions
. Nucl. Phys. 947, 155-160 (2016). arXiv:1507.04690, https://doi.org/10.1016/j.nuclphysa.2015.12.009Complications in the interpretation of the charge asymmetry dependent π flow for the chiral magnetic wave
. Phys. Rev. 101, 014913 (2020). arXiv:1910.02896, https://doi.org/10.1103/PhysRevC.101.014913Mapping the Phases of Quantum Chromodynamics with Beam Energy Scan
. arXiv:1906.00936Possible observables for the chiral electric separation effect in Cu + Au collisions
. Phys. Rev. 91, 054901 (2015). arXiv:1501.03903, https://doi.org/10.1103/PhysRevC.91.054901Globally polarized quark-gluon plasma in non-central A+A collisions
. Phys. Rev. Lett. 94, 102301 (2005). [Erratum: Phys. Rev. Lett.96,039901(2006)]. arXiv:nucl-th/0410079, https://doi.org/10.1103/PhysRevLett.94.102301Polarized secondary particles in unpolarized high energy hadron-hadron collisions?
. arXiv:nucl-th/0410089Global quark polarization in non-central A+A collisions
. Phys. Rev. 77, 044902 (2008). arXiv:0710.2943, https://doi.org/10.1103/PhysRevC.77.044902Quark Polarization in a Viscous Quark-Gluon Plasma
. Phys. Rev. 84, 054910 (2011). arXiv:1108.5649, https://doi.org/10.1103/PhysRevC.84.054910Polarization of massive fermions in a vortical fluid
. Phys. Rev. 94, 024904 (2016). arXiv:1604.04036, https://doi.org/10.1103/PhysRevC.94.024904Spin-dependent distribution functions for relativistic hydrodynamics of spin-1/2 particles
. Phys. Rev. 97, 116017 (2018). arXiv:1712.07676, https://doi.org/10.1103/PhysRevD.97.116017Covariant Spin Kinetic Theory I: Collisionless Limit
. arXiv:2002.03753Global polarization of Λ hyperons in Au+Au collisions at sNN = 200 GeV
. Phys. Rev. 98, 014910 (2018). arXiv:1805.04400, https://doi.org/10.1103/PhysRevC.98.014910Study of Λ polarization in relativistic nuclear collisions at sNN=7.7 - 200 GeV
. Eur. Phys. J. 77, 213 (2017). arXiv:1610.04717, https://doi.org/10.1140/epjc/s10052-017-4765-1Global Λ polarization in high energy collisions
. Phys. Rev. 95, 031901 (2017). arXiv:1703.03770, https://doi.org/10.1103/PhysRevC.95.031901Global Λ polarization in heavy-ion collisions from a transport model
. Phys. Rev. 96, 054908 (2017). arXiv:1704.01507, https://doi.org/10.1103/PhysRevC.96.054908Λ hyperon polarization in relativistic heavy ion collisions from a chiral kinetic approach
. Phys. Rev. 96, 024906 (2017). arXiv:1706.09467, https://doi.org/10.1103/PhysRevC.96.024906Searching for the Subatomic Swirls in the CuCu and CuAu Collisions
. Phys. Lett. 788, 409-413 (2019). arXiv:1712.00878, https://doi.org/10.1016/j.physletb.2018.09.066Λ and Λ¯ spin interaction with meson fields generated by the baryon current in high energy nuclear collisions
. Phys. Rev. 99, 021901 (2019). arXiv:1807.11521, https://doi.org/10.1103/PhysRevC.99.021901A study of thermal vorticity in PICR hydrodynamic model
. arXiv:1912.00209Vorticity and Particle Polarization in Relativistic Heavy-Ion Collisions
. 2019. arXiv:1910.01332Investigating different Λ and Λ¯ polarizations in relativistic heavy-ion collisions
. Phys. Lett. 786, 255-259 (2018). arXiv:1707.07262, https://doi.org/10.1016/j.physletb.2018.10.001Magnetic Field Induced Polarization Difference between Hyperons and Anti-hyperons
. Phys. Lett. 798, 134929 (2019). arXiv:1905.12613, https://doi.org/10.1016/j.physletb.2019.134929Magnetic Field in the Charged Subatomic Swirl
. arXiv:1904.04704Polarization of Λ (Λ¯) hyperons along the beam direction in Au+Au collisions at sNN = 200 GeV
. Phys. Rev. Lett. 123, 132301 (2019). arXiv:1905.11917, https://doi.org/10.1103/PhysRevLett.123.132301Collective Longitudinal Polarization in Relativistic Heavy-Ion Collisions at Very High Energy
. Phys. Rev. Lett. 120, 012302 (2018). arXiv:1707.07984, https://doi.org/10.1103/PhysRevLett.120.012302Probing vorticity structure in heavy-ion collisions by local Λ polarization
. Phys. Rev. 98, 024905 (2018). arXiv:1803.00867, https://doi.org/10.1103/PhysRevC.98.024905Feed-down effect on spin polarization
. Phys. Rev. 100, 014913 (2019). arXiv:1905.03120, https://doi.org/10.1103/PhysRevC.100.014913Polarization transfer in hyperon decays and its effect in relativistic nuclear collisions
. Eur. Phys. J. 79, 741 (2019). arXiv:1905.03123, https://doi.org/10.1140/epjc/s10052-019-7213-6Relaxation times for chiral transport phenomena and spin polarization in a strongly coupled plasma
. Phys. Rev. 98, 056018 (2018). arXiv:1805.04057, https://doi.org/10.1103/PhysRevD.98.056018A microscopic description for polarization in particle scatterings
. Phys. Rev. 100, 064904 (2019). arXiv:1904.09152, https://doi.org/10.1103/PhysRevC.100.064904Relaxation Time for Strange Quark Spin in Rotating Quark-Gluon Plasma
. Phys. Rev. 101, 024907 (2020). arXiv:1907.10750, https://doi.org/10.1103/PhysRevC.101.024907Relaxation time for quark spin and thermal vorticity alignment in heavy-ion collisions
. Phys. Lett. 801, 135169 (2020). arXiv:1909.00274, https://doi.org/10.1016/j.physletb.2019.135169Relaxation time for the alignment between the spin of a finite-mass quark/antiquark and the thermal vorticity in relativistic heavyion collisions
.arXiv:2003.06545Azimuthal angle dependence of the longitudinal spin polarization in relativistic heavy ion collisions
. Phys. Rev. 99, 011903 (2019). arXiv:1810.10359, https://doi.org/10.1103/PhysRevC.99.011903Spin polarizations in a covariant angular momentum conserved chiral transport model
. arXiv:1910.06774Longitudinal spin polarization in a thermal model
. Phys. Rev. 100, 054907 (2019). arXiv:1904.00002, https://doi.org/10.1103/PhysRevC.100.054907Local spin polarization in high energy heavy ion collisions
. Phys. Rev. Res. 1, 033058 (2019). arXiv:1906.09385, https://doi.org/10.1103/PhysRevResearch.1.033058Fluid Dynamics Study of the Λ Polarization for Au+Au Collisions at sNN = 200 GeV
. Eur. Phys. J. 80, 39 (2020). arXiv:1907.00773, https://doi.org/10.1140/epjc/s10052-019-7576-8Spin alignment of vector mesons in non-central A+A collisions
. Phys. Lett. 629, 20-26 (2005). arXiv:nucl-th/0411101, https://doi.org/10.1016/j.physletb.2005.09.060Signatures of the vortical quark-gluon plasma in hadron yields
. arXiv:2002.10082Relativistic fluid dynamics with spin
. Phys. Rev. 97, 041901 (2018). arXiv:1705.00587, https://doi.org/10.1103/PhysRevC.97.041901Causality and dissipation in relativistic polarizable fluids
. Phys. Rev. 100, 056011 (2019). arXiv:1807.02796, https://doi.org/10.1103/PhysRevD.100.056011Relativistic hydrodynamics for spin-polarized fluids
. Prog. Part. Nucl. Phys. 108, 103709 (2019). arXiv:1811.04409, https://doi.org/10.1016/j.ppnp.2019.07.001Fate of spin polarization in a relativistic fluid: An entropy-current analysis
. Phys. Lett. 795, 100-106 (2019). arXiv:1901.06615, https://doi.org/10.1016/j.physletb.2019.05.040Relativistic dissipative spin dynamics in the relaxation time approximation
. arXiv:2002.03937Quantum Kinetic Theory of Spin Polarization of Massive Quarks in Perturbative QCD: Leading Log
. Phys. Rev. 100, 056022 (2019). arXiv:1905.10463, https://doi.org/10.1103/PhysRevD.100.056022Relativistic Kinetic Theory
. Principles and Applications, (Quantum Transport Theory for Abelian Plasmas
. Annals Phys. 173, 462-492 (1987). https://doi.org/10.1016/0003-4916(87)90169-2WIGNER TRANSFORMATION IN CURVED SPACE-TIME AND THE CURVATURE CORRECTION OF THE VLASOV EQUATION FOR SEMICLASSICAL GRAVITATING SYSTEMS
. Phys. Rev. 32, 1871-1888 (1985). https://doi.org/10.1103/PhysRevD.32.1871Quantum Kinetic Field Theory in Curved Space-time: Covariant Wigner Function and Liouville-vlasov Equation
. Phys. Rev. 37, 2901 (1988). https://doi.org/10.1103/PhysRevD.37.2901Wigner function and quantum kinetic theory in curved space-time and external fields
. J. Math. Phys. 35, 2105-2129 (1994). arXiv:gr-qc/9309005, https://doi.org/10.1063/1.530542Chiral kinetic theory in curved spacetime
. Phys. Rev. 99, 085014 (2019). arXiv:1812.10127, https://doi.org/10.1103/PhysRevD.99.085014Lorentz Invariance in Chiral Kinetic Theory
. Phys. Rev. Lett. 113, 182302 (2014). arXiv:1404.5963, https://doi.org/10.1103/PhysRevLett.113.182302Berry Curvature, Triangle Anomalies, and the Chiral Magnetic Effect in Fermi Liquids
. Phys. Rev. Lett. 109, 181602 (2012). arXiv:1203.2697, https://doi.org/10.1103/PhysRevLett.109.181602Chiral Kinetic Theory
. Phys. Rev. Lett. 109, 162001 (2012). arXiv:1207.0747, https://doi.org/10.1103/PhysRevLett.109.162001Chiral Anomaly and Local Polarization Effect from Quantum Kinetic Approach
. Phys. Rev. Lett. 109, 232301 (2012). arXiv:1203.0725, https://doi.org/10.1103/PhysRevLett.109.232301Kinetic theory with Berry curvature from quantum field theories
. Phys. Rev. 87, 085016 (2013). arXiv:1210.8158, https://doi.org/10.1103/PhysRevD.87.085016Berry Curvature and Four-Dimensional Monopoles in the Relativistic Chiral Kinetic Equation
. Phys. Rev. Lett. 110, 262301 (2013). arXiv:1210.8312, https://doi.org/10.1103/PhysRevLett.110.262301Collisions in Chiral Kinetic Theory
. Phys. Rev. Lett. 115, 021601 (2015). arXiv:1502.06966, https://doi.org/10.1103/PhysRevLett.115.021601Simulating Chiral Magnetic and Separation Effects with Spin-Orbit Coupled Atomic Gases
. Sci. Rep. 6, 20601 (2016). arXiv:1506.03590, https://doi.org/10.1038/srep20601Relativistic Chiral Kinetic Theory from Quantum Field Theories
. Phys. Rev. 95, 091901 (2017). arXiv:1612.04630, https://doi.org/10.1103/PhysRevD.95.091901Worldline construction of a covariant chiral kinetic theory
. Phys. Rev. 96, 016023 (2017). arXiv:1702.01233, https://doi.org/10.1103/PhysRevD.96.016023Consistent relativistic chiral kinetic theory: A derivation from on-shell effective field theory
. Phys. Rev. 98, 076005 (2018). arXiv:1806.01684, https://doi.org/10.1103/PhysRevD.98.076005Complete and Consistent Chiral Transport from Wigner Function Formalism
. Phys. Rev. 98, 036010 (2018). arXiv:1801.03640, https://doi.org/10.1103/PhysRevD.98.036010Chiral Vortical Effect For An Arbitrary Spin
. JHEP 03, 084 (2019). arXiv:1805.08779, https://doi.org/10.1007/JHEP03(2019)084Chiral Kinetic Theory from Effective Field Theory Revisited
. JHEP 06, 060 (2019). arXiv:1901.01528, https://doi.org/10.1007/JHEP06(2019)060Chiral kinetic theory from Landau level basis
. Phys. Rev. 101, 034006 (2020). arXiv:1909.11514, https://doi.org/10.1103/PhysRevD.101.034006Photon Self-energy in Magnetized Chiral Plasma from Kinetic Theory
. arXiv:2002.07959Relativistic Quantum Kinetic Theory for Massive Fermions and Spin Effects
. Phys. Rev. 100, 056021 (2019). arXiv:1902.06510, https://doi.org/10.1103/PhysRevD.100.056021Kinetic theory for massive spin-1/2 particles from the Wigner-function formalism
. Phys. Rev. 100, 056018 (2019). arXiv:1902.06513, https://doi.org/10.1103/PhysRevD.100.056018Axial Kinetic Theory and Spin Transport for Fermions with Arbitrary Mass
. Phys. Rev. 100, 096011 (2019). arXiv:1903.01653, https://doi.org/10.1103/PhysRevD.100.096011Mass Correction to Chiral Kinetic Equations
. Phys. Rev. 100, 014015 (2019). arXiv:1903.03461, https://doi.org/10.1103/PhysRevD.100.014015Chiral kinetic theory from the on-shell effective theory: derivation of collision terms
. arXiv:1908.00561Quantum kinetic theory for spin transport: general formalism for collisional effects
. arXiv:2002.02612Chiral Radiation Transport Theory of Neutrinos
. arXiv:2002.11348