1 Introduction
Understanding the phase structure and, more generally, the many-body physics of quantum chromodynamics (QCD) is one of the most outstanding challenges in contemporary physics. At low temperature and low quark chemical potential, the QCD matter is formed by hadrons; at high temperature and/or quark chemical potentials, we expect the QCD matter where quarks and gluons are liberated from the confinement. The "critical temperature" of the transition from the hadronic matter to the quark-gluon matter is expected to be at the order of the QCD confinement scale, ΛQCD∼200 MeV, at zero quark chemical potential. The unique terrestrial experiments that can achieve such an extremely high temperature are the relativistic heavy-ion collisions which have been carried out in the Relativistic Heavy Ion Collider (RHIC) at Brookhaven National Laboratory (BNL) and in the Large Hadron Collider (LHC) at CERN. The experimental data produced in RHIC and LHC have shown many features that support the formation of the hot quark-gluon plasma (QGP) and have also revealed a number of interesting properties of the QGP.
In noncentral heavy-ion collisions, strong magnetic fields can also be created [1-4]. This is because that the two colliding nuclei generate two electric currents in opposite directions and thus produce a magnetic field perpendicular to the reaction plane (the plane defined by the impact parameter and the beam direction). The typical strength of the magnetic fields can be roughly estimated by using the Biot-Savart formula,
In addition to the above mentioned subjects, strong magnetic fields may also drive a number of other phenomena which will not be discussed in this article. These phenomena include the magnetic catalysis of chiral symmetry breaking [43-46], the inverse magnetic catalysis or magnetic inhibition at finite temperature and density [47-63], the rotating system under a strong magnetic field [64-66], the possible ρ meson condensation in strong magnetic field [67-71], the neutral pion condensation in vaccum [72], the magnetic-field induced anisotropic viscosities in hydrodynamic equations [73-78], and the magnetic-field induced particle production [79-89]. Some of these intriguing effects are reviewed in Refs. [90-92].
In Sec. 2, we briefly review some features of the magnetic fields in heavy-ion collisions. In Sec. 3, we discuss the chiral magnetic effect, chiral separation effect, chiral magnetic wave, and the current status of the experimental search of these anomalous transport phenomena. In Sec. 4, we discuss how the magnetic fields modify the properties of the heavy-flavor mesons and the heavy-quark transport in a hot medium. Phenomenological implications are also discussed. We summarize in Sec. 5.
2 Magnetic fields in heavy-ion collisions
In relativistic heavy-ion collisions, due to the fast motion of the ions which carry positive charges, large magnetic fields can be generated in the reaction zones. In this section, we briefly review some general properties of such-generated magnetic fields by both numerical and analytic models. More detailed information of the numerical study can be found in the original references and in Ref. [93].
2.1 Numerical approach
A number of numerical simulations have revealed that the magnitudes of the magnetic fields in heavy-ion collisions can be very large [5-16]. In Fig. 1, the different spatial components of the magnetic fields at r= 0 (i.e., the center of the overlapping region) and t=0 (i.e., the time when the two colliding nuclei completely overlap) are presented as functions of the impact parameter, b. The fields are calculated by using the Lienard-Wiechert potential on the event-by-event basis where the events are generated by the HIJING event generater [10]. Please find the discussions on the quantum correction to the otherwise classical Leinard-Wiechert potential in Refs. [16, 93]. The most important information from Fig. 1 are three folds.
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F001.jpg)
First, it shows that heavy-ion collision experiments at RHIC and LHC generate very strong magnetic fields. The event-averaged magnetic fields at Au + Au collisions at
Second, although we do see that only the y-component (y axis is set to be perpendicular to the reaction plane) of the magnetic fields remains after the event average (due to the left-right symmetry of the noncentral collisions), we can find that the event-by-event fluctuation of the nuclear distribution results in nonzero x-component of the magnetic fields as well. These are reflected in the averaged absolute values of the fields in Fig. 1 and is most evident for central collisions [8, 10].
Third, the fields in LHC Pb + Pb at 2.76 TeV are roughly 13.8=2.76/0.2 times larger than that in RHIC Au + Au collisions at 200 GeV. In fact, as investigated in Refs. [8, 10], to high precision, the magnitudes of the magnetic fields linearly depend on the collision energy,
The simulations depicted in Fig. 1 are only for fields at t=0. As the hot medium generated in heavy-ion collisions is expanding fast, a natural question is: how do the magnetic fields evolve in time in the expanding medium? This is still an unsolved problem. We summarize the recent progress briefly here and in the next subsection.
As we explained, the strong initial magnetic fields are consequences of the smallness of the system and the relativistically fast motion of the colliding nuclei. The main contributors to the initial magnetic fields are the spectator protons (which do not participate into the collision) in the colliding uclei. They leave the collision zone very fast in a time scale characterized by the Lorentz-suppressed radius of the nucleus, i.e.,
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F002.jpg)
2.2 Analytical approach
To understand the basic picture of the creation of the magnetic fields, we compared the numerical results shown in the previous section with a simple analytical model proposed in Ref. [28]. With the relativistic beam energy, the momenta of the colliding nuclei are so large that the nuclear stopping power is not really strong enough to stop them: the incident nuclei pass each other and recede from the collision point with almost the same rapidity as the incident beam. Therefore, in this simple model, we assume that a magnetic field is created by protons freely streaming with the beam rapidity, and thus do not distinguish the spectators and participants. From the comparison between the numerical and analytic modeling, we will also discuss effects of the nuclear stopping power.
As we have specified with the classical motion of charged particles, the magnitude of the magnetic field can be estimated by using the Lienard-Wiechert potential
where Ybeam is the absolute value of the beam rapidity. The positive and negative signs correspond to the nuclei moving in the positive and negative z-direction, respectively. The coordinate system is illustrated in Fig. 3 (Left). We have defined the coordinate vector
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F003.jpg)
The Lienard-Wiechert potential in Eq. (1) is obtained by boosting the electro-static potential in the rest frame of the nuclei (see Fig. 3). As is well known, the net electro-static potential can be calculated from the effective point-like charge sitting at the center of the sphere. Therefore, the effective charge Zeff appearing in Eq. (1) should be the total amount of the charges inside the sphere, i.e.,
where ρ(r′) is the charge distribution inside the nuclei. Note that the upper boundary of the integral is given by the distance r± from the center to the observer. When the observer is in the exterior of the nucleus r±≥RA, one simply has the total charge of the nucleus Zeff(r±>RA)=Z. The impact parameter and time dependences of the effective charge is shown in the right panel of Fig. 3. We find significant dependence on both the impact parameter and the time collapsed after the collision, which will be important to reproduce the results from the numerical simulation shown in the previous section.
Plugging the effective charge (2) to the Lienard-Wiechert potential (1), we obtain the magnitude of the magnetic field shown in Figs. 4, 5 and 6. Here, we have taken the Woods-Saxon form for the charge distribution ρ(r)=NZ/{1+exp[(r-RA)/a]} with
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F004.jpg)
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F005.jpg)
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F006.jpg)
Figure 4 shows the time dependence of the magnetic field obtained from the analytic model (solid lines) and the event-averaged numerical simulations (dots). The latter has been already shown in Fig. 2. We find a good agreement between the results from these two models in the time range before and after the collision except for the late time. The origin of the deviation in the late time is, as can be guessed from the asymmetric time dependence in the numerical result with respect to the origin of time, attributed to the effect of the nuclear stopping power. Because of the deceleration by the nuclear stopping power, participant nucleons (also called remnants) will stay in the collision region longer than in the free streaming case so that they will maintain the magnetic field for a longer time.
Figure 5 shows the impact parameter dependence of the magnetic field. Again, the sold lines and dots show the results from the analytic model and the event-averaged numerical simulations, respectively, and we find a good agreement between them. The slight deviation may be originated from the energy loss due to the nuclear stopping power in the numerical simulation. In addition to these results, we show the magnetic field created by point-like nuclei carrying the total charge Z, i.e., Zeff (t,x)=Z in Eq. (1). We find that this simplest model completely fails to reproduce the numerical results when there is an overlap between the nuclei (
Finally, Fig. 6 shows the beam energy dependence of the magnetic field from the analytic model (left panel) and the numerical simulation (right panel) [7]. While we find a good agreement, the nuclear stopping power will play an important role in the low energy collisions such as those in the beam energy scan program in RHIC, FAIR, and J-PARC.
In the above discussions, we omit one important issue which may substantially modify the time evolution of the magnetic fields, that is, the finite electric conductivity, σ, of the produced quark-gluon matter. If σ is large, the Faraday induction effect may be strong as well and the lifetime of the magnetic fields can be significantly prolonged. As a matter of fact, the equilibrium QGP is a very good conductor according to the theoretical and lattice QCD studies [96-105]; see also Ref. [106] for review.
Now let us understand how the large σ induces the strong Faraday induction and prolongs the lifetime of the magnetic fields; more discussions can be found in review [93]. Let us write down the Maxwell’s equations in the background of a charged flow characterized by velocity field v,
where J is the electric current which according to the Ohm’s law is determined by
Substituting J into Eq. (4) and eliminating E by using Eq. (3), one can obtain an equation that governs the time evolution of the magnetic field
where we have regarded the QGP as locally neutral (this is a good assumption for high-energy collisions) so that ∇⋅E=ρ=0.
The Faraday prolongation of the lifetime of the magnetic fields is most easily seen in two limits. The first limit is the no-flow limit, i.e., the limit when v can be neglected. In this case, if the electric conductivity is so large that
Clearly, in this case, the decay of the magnetic field is due to diffusion, and the diffusion time is given by
where L characterizes the length scale of the QGP system. For sufficiently large σ, the diffusion time, tD, could be larger than τB [79].
The second limit is the ideal magnetohydrodynamic limit, i.e., when σ is very large and at the same time the flow velocity, v, is also large so that we can neglect the term divided by σ in Eq. (6) and obtain
This equation is known to lead to the so-called frozen-in theorem, i.e., the magnetic lines are frozen in the ideally conducting plasma. Thus in this limit, the decay of the mangnetic fields is totally due to the expansion of the QGP. For high-energy heavy-ion collisions, the longitudinal expansion is well described by the Bjorken flow, vz=z/t, which makes the area of cross section of the overlapping region on the reaction plane grow linearly in time. Thus the frozen-in theorem leads us to the conclusion that in the ideal magnetohydrodynamic limit, before the transverse expansion play significant role, the magnetic fields will decay inversely proportional to time, By(t)∝(t0/t)By(t0), which is much slower than the 1/t3-type decay in the insulating case. The transverse expansion of the system can happen because of the transverse gradient of the fluid pressure. When this transverse expansion becomes fast, the decay of the fields will become fast as well, see discussions in Refs. [10, 93].
So far, we discussed two special limiting cases from which we can analytically extract the effect of Faraday induction. However, the realistic environment produced by heavy-ion collisions is not the case as described by the two limits. Moreover, the system is essentially non-equilibrium. Thus, the precise time evolution of the magnetic fields is really a hard task to achieve, and some recent progresses can be found in Refs. [10, 15, 79, 82, 107-113].
In the above discussions, we focus only on collisions of the same spherical nuclei, i.e., the Au + Au collisions and Pb + Pb collisions. The Heavy-ion program at RHIC also performed collisions of different nuclei, like Cu + Au collisions, as well as collisions of prolate nuclei like the U + U collisions. Of course, these collisions can also generate strong magnetic fields; see for example, the studies in Refs [12, 114-118] and recent review [93]. It is worth mentioning that, owing to the charge number asymmetry between Cu and Au nuclei, the Cu + Au collisions can generate a strong electric field pointing from Au nucleus to Cu nucleus [114-117, 119]. Such E field may lead to a splitting between the directed flow v1 of positively and negatively charged hadrons, which has been recently test by STAR Collaboration at RHIC [120].
3 Magnetic-field induced anomalous transports in heavy-ion collisions
The strong magnetic fields can induce a variety of novel quantum effects in the quark-gluon plasma which we will discuss in the following sections. In this section, we will focus on the ones that are deeply related to the chiral anomaly in QCD and QED, namely, the chiral magnetic effect (CME) and its relatives, the chiral separation effect, the chiral magnetic wave, etc. All these effects represent currents or collective modes that transport vector charges or axial charges in QGP which appear only at quantum level so that they do not have classical counterparts. The recent reviews of these anomalous transports can be found in Refs. [93, 121-126].
3.1 The chiral magnetic effect
The CME represents the generation of vector current induced by an external magnetic field in chirality-imbalanced (P-odd) medium [5, 17]. The CME composes a very general class of anomalous transport phenomena spreading in a wide contexts in physics, ranging from astrophysics, condensed matter physics, nuclear physics, to particle physics. We apologize for not being able to mention all the relevant works. We refer the readers to other review articles as mentioned above for more information. The CME can be expressed as
where e is the charge of the fermion, JV is the vector current defined by
Although the CME stems from the profound chiral anomaly in gauge theory, its physical origin can be understood in an intuitive way by adopting the Landau quantization picture. We know that the electrically charged fermions in a magnetic field occupy a tower of Landau levels (i.e., the energy spectrum of charged fermions in a magnetic field is discrete) in such a way that the lowest Landau level contains fermions with only one spin polarization (i.e., the spin of positively (negatively) charged fermions is polarized to be along (opposite to) the direction of the magnetic field) while higher Landau levels are occupied by equal numbers of up-polarized and down-polarized fermions. Now, let us consider the situation that the number of right-handed (RH) and the number of left-handed (LH) fermions are unequal, say, for example, NR>NL. Thus there are more RH fermions than LH fermions confined in the lowest Landau level which means that a current consisting of positively charged fermions will flow along the magnetic field as RH fermions lock their momenta with their spins, while a current consisting of negatively charged fermions will flow opposite to the magnetic field. The net effect is the appearance of an electric current along the magnetic field. This is the CME current. Note that the higher Landau levels do not contribute to the CME current as higher Landau levels are occupied symmetrically by spin-up and spin-down fermions and their contribution to the CME current cancel.
Now let us quantify the above intuitive argument to see how we can arrive at Eq. (10). To simplify our lives, we consider a system of massless Dirac fermions with only positive charge, e, in an constant magnetic field, B, pointing along z direction which is supposed to be so strong that all the fermions are confined in the lowest Landau level. The number densities of RH and LH fermions are given by
where the prefactor eB/(2π) is the density of states of the lowest Landau level in the transverse directions and pF denotes the Fermi momentum (which is the momentum that separates the empty states and occupied states). Moreover, chiral fermions on the lowest Landau level move at the speed of light in a way correlated to their spins which tell us that the RH fermions generate a current along z direction, JR=enR, while LH fermions generate a current along -z direction, JL=-enL. Therefore the net vector current reads
which, once we equate
We emphasize that although the above argument is for a non-interacting system and relies on Landau quantization picture, the CME conductivity σCME=e2μA/(2π2) is actually fixed by the chiral anomaly equation and universal in the sense that it does not involve concrete microscopic scattering among fermions. This is in parallel to the fact that the chiral anomaly equation itself is universal so that it does not receive perturbative correction from scattering between fermions.
Apparently, the occurrence of CME requires an environmental parity violation as being characterized by μA. A nonzero μA may be induced by different mechanisms in different physical systems. For example, the electro-weak plasma can have a nonzero μA because the weak interaction does not respect parity symmetry [127]. Recently, it is realized in condensed matter experiments the so-called Weyl or Dirac semimetals, whose band structures permit level crossing around the low energy excitations, are Weyl or Dirac fermions. The CME was observed in these semimetals by applying parallel electric and magnetic fields, which induce a difference between the chemical potentials of RH and LH fermions via the chiral anomaly [128-131]; see also discussions in Refs. [92, 93, 126, 132]. In the hot quark-gluon matter produced in heavy-ion collisions, μA may appear due to the topological transition between the degenerate vacua of the gluonic sector of QCD characterized by different winding numbers [5, 133, 134]. The rate of such topological transition is exponentially suppressed at zero temperature as two neighbor vacua are separated by a finite energy barrier of the order of QCD confinement scale ΛQCD∼200 MeV. However, at high temperature, the topological transition can be induced by another classical and thermal excitation called sphaleron [135, 136] which, instead of tunneling through the barrier, can take the vacuum over the barrier at a large rate. At equilibrium, the probability of having a topological transition that drives the winding number to increase is equal to the probability of having an opposite topological transition that decreases the winding number. Thus, what really matters is the fluctuation of the topological transition above global equilibrium which results in local domains in the quark-gluon matter within which the gluonic configuration has a nonzero winding number. The quarks coupled with such gluons via the chiral anomaly can acquire nonzero chirality and thus a nonzero μA. Furthermore, μA in heavy-ion collisions may also be induced by other mechanisms. For example, the parallel chromoelectric and chromomagnetic fields, Ea·Ba≠0, in the initial glasma can also induce a nonzero μA [137-139].
3.2 Experimental signals of the chiral magnetic effect
If the CME occurs in the quark-gluon matter produced in the heavy-ion collisions, what will be its experimental signals? We have already seen that the magnetic fields in heavy-ion collisions are, on average over events, perpendicular to the reaction plane; thus the CME is expected to induce an electric current which on average is also perpendicular to the reaction plane. Such an electric current will drive a charge dipole in the fireball which, after the hydrodynamic evolution of the fireball, will be converted to a charge dipole in momentum space. However, as μA can flip the sign from event to event, the charge dipole could also flip its direction from event to event, and thus the dipolar distribution in different events cancel each other after event average and cannot be measured directly. Instead, one should look for an observable which measures the dipolar fluctuation over events. Such an observable can be chosen as the following charge-dependent two-particle azimuthal correlation [133, 140]:
where the indices α and β denote the charge of the hadrons, φα and φβ are the azimuthal angles of hadrons of charge α and β, respectively, ΨRP is the reaction plane angle, and <…> denotes average over events; see Fig. 7 for illustration. It is easy to understand that a charge separation with respect to the reaction plane, will give a positive γ+- (and γ-+) and a negative γ++ or γ–. We will refer to γ+-=γ-+ as γOS and γ++≈γ– as γSS where "OS" stands for "opposite-sign" and "SS" stands for "same-sign". In real experiments, the measurements are done with three-particle correlations where the third hadron (of arbitrary charge) is used to reconstruct the reaction plane.
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F007.jpg)
The correlation γαβ was measured by STAR Collaboration at RHIC for Au + Au and Cu + Cu collisions at
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F008.jpg)
Although the experimental data supports the presence of the CME, the interpretation of the data is not unique owing to possible background effects that are not related to the CME. First, the transverse momentum conservation (TMC) can contribute to γαβ [148-150]. The momentum conservation enforces the sum of the momenta projected to the transverse plane (the plane that is perpendicular to the beam direction) of all the particles to be zero at any time during the evolution of the system. Thus, the detection of a hadron of transverse momentum, p⊥, requires the presence of other produced hadrons whose transverse momenta must be summed to -p⊥. This enforces an intrinsic back-to-back two-particle correlation:
where we used that
However, there are other background effects contributing to Δγ. The most notable one may be the local charge conservation (LCC) [148, 151, 152] which assumes that the pairs of opposite charges are created locally (i.e., within a small volume) in space at the late stage of the heavy-ion collisions. Then, due to the elliptic flow, opposite-sign pairs are strongly correlated while the same-sign pairs are not correlated; thus LCC gives a finite contribution to Δγ. The detailed analysis reveals that the LCC contribute to Δγ the amount of [93]
where M is the average number of hadrons in a typical local neutral cell. As M is expected to be not very large, so we expect the LCC contribution is at the order of 10-4 or even larger in, e.g., RHIC Au + Au collisions at 200 GeV. This is comparable to the experimental results in Fig. 8.
The background effects mask the possible CME signals. To disentangle the possible CME signal and the flow-related backgrounds, one can setup experiments to either vary the backgrounds with the signal fixed or vary the signal with the backgrounds fixed (It is also important to setup delicate analysis methods to extract the elliptic flow on the event-by-event basis [153, 154]). The former approach can be carried out by using the prolate shape of the uranium nuclei [155]. In central U + U collisions, there would be sizable v2 due to the prolate shape of the colliding nuclei while the magnetic field would be very tiny [12]. The STAR Collaboration collected 0-1% most central events from U + U collisions at
Recently, the CMS Collaboration at LHC reported the measurement of γαβ in high multiplicity events in p + Pb and Pb + Pb collisions at
3.3 Chiral separation effect and chiral magnetic wave
By looking at the CME (10), one may wonder whether there is a quantum-anomaly induced axial current in response to the external magnetic field and under what condition such current can appear. The answer is yes, and such a current can be caused by the chiral separation effect (CSE) which is the dual effect to the CME [18, 19, 161]:
where the axial current is defined by
We can understand the CSE from an intuitive picture. In parallel to the discussion of CME in Sec.3.1, let us consider a massless fermionic system under a strong magnetic field, and let us focus on the lowest Landau level on which the spin of positively charged fermions are polarized to be parallel to the magnetic field (we call its direction the "up" direction and its opposite direction the "down" direction) while the spin of negatively charged fermions are polarized to be anti-parallel to the magnetic field. If a positively charged fermion on the lowest Landau level moves upward then it contributes to the RH current while if it moves downward it contributes to the LH current; thus no matter which direction it moves it contributes to an axial current along the magnetic field. Similarly, it is easy to see that the negatively charged fermions contributes an axial current in the direction opposite to the magnetic field. Now, if the system contains more positively charged fermions than negatively charged fermions, namely, if μV>0, the downward-moving axial current cannot compensate the upward-moving axial current and thus we get a net axial current along the direction of the magnetic field. This is the CSE current. The fermions occupying the higher Landau levels do not contribute to the CSE current as their spins are not polarized.
Let us formulate the above argument in a more precise way. For simplicity, we consider a system of massless Dirac fermions with only positive charge, e, in an constant magnetic field, B, in z direction which is supposed to be so strong that all the fermions are confined in the lowest Landau level. The number densities of RH and LH fermions are given by Eq. (12). As the massless fermions on the lowest Landau level move at the speed of light in a way correlated to their spins which means that the RH fermions generate a current along z direction, JR=enR, while LH fermions generate a current along -z direction, JL=-enL. Therefore the net axial current reads
which, once we equate
The CME and CSE couple together the the vector and axial densities and currents in the presence of the magnetic field and lead to collective modes called chiral magnetic waves (CMWs). This is most easily seen when we express CME and CSE in the chiral basis:
where the RH and LH quantities are defined as JR/L=JV±JA)/2 and μR/L=μV±μA. Substituting JR and JL in to the continuation equations and considering a small fluctuation in the density
where
3.4 Experimental signals for chiral magnetic wave
The CMWs are able to transport both chirality and electric charge and can induce an electric quadrupole in the QGP [21, 22]; see Fig. 9 for an illustration. Owing to strong in-plane pressure gradient, such an electric quadrupole can cause a splitting between the v2 of π+ and π- [22]; see also Refs. [139, 172-174] for more advanced numerical implementations. The CMW-induced charge dependent elliptic flows can be parameterized as
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F009.jpg)
where v2 is the average elliptic flow and Ach=(N+-N-)/(N++N-) is the net charge asymmetry parameter which is usually small (the events that can generate large Ach are rare). The slop parameter, r, is predicted to be proportional to <(eB)2cos[2(ψB-ΨRP)]> where ψB is the azimuthal direction of the magnetic field [11, 22, 170].
Recently, the CMW has been tested by STAR Collaboration [146, 175, 176] at RHIC and by ALICE Collaboration [177] at LHC. The data shows an elliptic-flow difference, v2(π-)-v2(π+), linear in Ach with a positive slope whose centrality dependence can be fitted by the CMW estimation. Furthermore, the slope parameter, r, displays no obvious trend of the beam energy dependence for 10-60% centrality at
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F010.jpg)
However, ambiguities exist in the interpretation of the experimental results as there are background effects which can also lead to v2 splitting between π+ and π-. A variety of possible background effects were proposed, see Refs. [10, 178-183] and review articles, Refs. [93, 126], for discussions. It should be noted that none of the above backgrounds can successfully explain all the features of the experimental data. It is quite plausible that various effects (including CMW) are coupled to produce the measured pattern. Until we can successfully subtract the background effects, we cannot make a conclusive claim. On the experimental side, it will be important to perform more detailed measurements, such as the measurement of the v2 splitting versus transverse momentum and rapidity, the measurement for other hadrons like kaons, the measurement of splitting in other harmonic flows like v3, and the measurement of the cross-correlation of different observables. On the theoretical side, an anomalous hydrodynamic simulation or a simulation based on kinetic theory for the CMW that incorporate both the dynamically evolving magnetic fields and the realistic initial charge distribution will be very useful.
Before we end this section, we note that, in addition to the magnetic-field induced anomalous transports, the electric field and vorticity can also induce certain anomalous transport phenomena and corresponding collective modes. For example, electric field can induce an axial current in the medium with both μA and μV nonzero which is called chiral electric separation effect (CESE) and, once coupled with the normal Ohm’s current (the vector current induced by electric field), CESE leads to chiral electric wave and axial and vecor density waves [184-187]. The CESE may lead to observable effect in, Cu + Au collisions where persistant electric fields from Au to Cu nuclei exist [188]. The fluid vorticity can induce the so-called chiral vortical effects (CVEs), which are the vortical analogues of CME [189-192], and the corresponding collective mode, chiral vortical wave (CVW) [193]. The CVE and CVW may have important consequences in noncentral heavy-ion collisions where nonzero fluid vorticity can be generated as a result of angular momentum conservation [194-197]. Phenomenologically, the CVE can induce baryon number separation with respect to the reaction plane [198], and CVW may cause a v2 splitting between Λ and
4 Heavy-quark dynamics in magnetic fields
In the heavy ion collisions, the dominant process for the heavy-quark production is the initial hard scatterings among the partons from the colliding nuclei. This is because the thermal excitation is suppressed due to the large value of the heavy-quark mass compared to the temperature of the quark-gluon plasma (QGP). Therefore, the heavy quarks will serve as a probe of the dynamics in the magnetic field persisting in the early time of the collision events. In this section, we discuss effects of the magnetic field on the quarkonium spectra in Sec. 4.1 and the heavy-quark transport in a hot medium and magnetic field in Sec. 4.2.
The quarkonium spectra in magnetic fields have been investigated by using the Cornell potential model [27, 33, 38, 39], the QCD sum rules [26, 29, 30] with the help of the lattice QCD simulations for the heavy quark potential [32, 34], and the quark and gluon condensates [48-50, 200, 201]. The phenomenological consequences for the quarkonium production in the heavy-ion collisions have been discussed in Ref. [23-25, 27, 28, 37], which could be measured by the dilepton channel.
Most of the heavy quarks created in the hard scatterings will evolve in a hot medium and be observed as open heavy flavor mesons in the final state (see Ref. [202] for a review). Therefore, it is important to understand the transport property of the heavy quarks in a hot medium and magnetic field. In Sec. 4.2, we discuss the diagrammatic computation of the heavy-quark diffusion constant. The study on the heavy-quark dynamics in the magnetized medium has been just initiated [40, 42] and is awaiting for further developments.
4.1 Quarkonia in magnetic fields
4.1.1 Spin mixing and level repulsion
We shall begin with the simplest description of effects of magnetic fields on quarkonium spectra. In a weak magnetic field, the dominant effects of the magnetic fields can be described by the hadronic degrees of freedom. This contrasts the case of strong magnetic fields where modifications of nonperturbative vacuum structure and of the internal structure of hadrons are important (see Refs. [58, 52, 62, 68-70, 203-207] and references therein for the ongoing discussions). As long as the magnitude of the magnetic field is much smaller than the heavy quark mass, these effects will be subdominant contributions.
In the hadronic degrees of freedom, effects of magnetic fields appear in the mixing of spin eigenstates [25, 27, 29, 30, 33, 38, 39]. This mixing is caused by the breaking of the spatial rotation symmetry in magnetic fields: there remains only the azimuthal rotational symmetry with respect to the direction of the external magnetic field. This indicates that only the spin state along the magnetic field can persist as a good quantum number of the mesons, and thus that there is a mixing between the spin single and triplet states, Stotal, Sz) = (0,0) and (1,0), which are, for example, ηc and the longitudinal component of J/ψ.
We examine general mixing patterns among the pseudoscalar, P, vector, Vμ, scalar, S, and axial-vector, Aμ, quarkonia. Possible interaction vertices among those fields are informed from the Lorentz invariance and the parity and charge-conjugation symmetries. The vertices relevant for interactions among static charmonia are found to be [29, 30]
with m0=(mP+mV)/2, m1=(mS+mA)/2, and dimensionless effective coupling constants gPV, gVA, and gSA. The vertices proportional to
From the above vertices, one can identify two pairs of mixing partners: the mixing between P and V and between S and A are possible when they are at rest in external magnetic fields. Since the P and V are not mixed with the S and A, we shall focus on the mixing between S and V. As mentioned in the beginning of this section, a mixing effect arises only in the longitudinal component of the vector field, V||, and the other two transverse components persist as the energy eigenstates. This can be easily confirmed by identifying the non-vanishing components in Eq. (25). With a vanishing spatial momentum qμ=(ω, 0, 0, 0) and the vertex (25), the coupled equations of motion are found to be
Following from the equations of motion (28), we obtain the physical mass eigenvalues as
where
with eigenvectors given by
The mass eigenvalue (30) clearly shows the level repulsion in the presence of the mixing effect.
In case of mixing between ηc and J/ψ, the coupling constant, gPV, can be read off from the measured radiative decay width, J/ψ→ηc + γ [29, 30]. From the vertex (25), the invariant amplitude of this decay process is written as
where
We have also integrated over the phase-space volume in the two-body final state. The center-of-mass momentum in the final state is given by
Figure 11 shows the mass shifts in the presence of the mixing effect. We find the level repulsion between the ηc and longitudinal J/ψ mass spectra. While the mass of ηc decreases as eB increases, the mass of the longitudinal J/ψ (denoted by J/ψ∥) increases. This result clearly provides the simple picture of the effect of weak magnetic fields.
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F011.jpg)
The level repulsion occurs most strongly between a pair of adjacent states. Therefore, the other excited states give rise to only subdominant contributions to the mass shifts of ηc and the longitudinal J/ψ in a weak magnetic field. However, as the mass of the longitudinal J/ψ keeps increasing with an increasing magnetic field, it will approaches the other excited states. In such a strong magnetic field, many states will get involved in the mixing effects, which will result in an intriguing mixing patterns of the spectra as discussed in the next section.
Moreover, the mixing effects induced by the minimal coupling (25) to a magnetic field will not be valid in the strong-field regime. Since the strong magnetic field probes the inner structure of hadrons and modifies the QCD vacuum, we will need to look into the dynamics of the fundamental degrees of freedom. To investigate these effects, one can invoke on the Cornell potential models and QCD sum rules with the help of lattice QCD simulations. We will proceed to exploring these frameworks on top of the simplest framework for the mixing effect discussed in this section.
4.1.2 Cornell potential model in magnetic fields
To study the nonrelativistic bound states of the heavy quark and antiquark on the basis of the fundamental degrees of freedom, one can use the Cornell potential model [210, 211] with the help of the lattice QCD simulations. The basic framework for the 1S states of ηc, J/ψ, ηb, and Υ, was first elaborated by Alford and Strickland [27]. The authors included the interaction with the background magnetic field through the magnetic moments of constituents and the kinetic terms, as well as the mutual interactions by the linear and Coulomb potentials in vacuum. In this framework, the interactions through the magnetic moments give rise to the mixing between the spin singlet and triplet states discussed in the previous section. Based on this model, a systematic spectroscopy for excited states was recently performed by Suzuki and Yoshida owing to an improvement of computational method [38, 39]. Effects of strong magnetic fields on nonperturbative QCD vacuum were investigated by Bonati et al. who computed the modification of the quark-antiquark potential by lattice QCD simulation [32, 34] and plugged it into the bound-state problem [33]. These studies suggest an opportunity to explore the nature of QCD confinement in the presence of a strong magnetic field. In the following, we briefly summarize the framework for the potential model in a magnetic field and the numerical results for the mass spectra.
The Hamiltonian for the two-body bound state in a magnetic field is given by [25, 27, 38, 39, 33]
where
with the gauge potential, A(r), for the external magnetic field, B, the Pauli matrix, σ, and the heavy-quark mass, mQ. The indices i=1,2 denote the quark and antiquark, respectively, and thus q(1)=-q(2)(≡q) for quarkonia. The mutual interaction is described by the Cornell potential,
where μQ=q/(2mQ) and
Consistent with the discussion in the previous section, we find that, while the transverse modes |1,±1〉 are eigenstates of the two-body Hamiltonian, there is a mixing between |0, 0〉 and |1, 0〉 states which need to be diagonalized to the energy eigenstates.
A specific issue of the dynamics in magnetic fields is the non-conservation of the transverse momentum. The gauge potential for an external magnetic field will depend on the transverse coordinate as
where the labels denote the quark and antiquark as before. This momentum is related to the center-of-mass kinetic momentum as 〈Pkinetic〉=K-qB×〈r〉 with q≡q(1)=-q(2) and the relative coordinate r=r1-r2. The same issue of the momentum nonconservation is seen in diagrammatic calculations in the presence of a magnetic field. In such calculations, the gauge- and translational-symmetry breaking parts are encoded in the so-called Schwinger phase appearing from the gauge-dependent wave function of fermions in a magnetic field [212]. Effects of the Schwinger phase on the Bethe-Salpeter equation for light mesons were discussed in Ref. [62].
The Schrödinger equation for the two-body system (34) can be solved numerically for a fixed value of the conserved momentum, K. The parameters for the Cornell potential are so tuned to reproduce the vacuum spectrum in Ref. [27, 38, 39], and are imported from the lattice simulation in a finite magnetic field in Ref. [33]. In those papers, the numerical codes were carefully tested to reproduce the analytic solution for a harmonic potential.
Figure 12 shows the charmonium spectra from Ref. [27] with the potential in vacuum. At the vanishing momentum 〈Pkinetic〉=0 GeV, one sees that the masses of the ηc and longitudinal J/ψ decreases and increases, respectively, as expected from the mixing effect. On the other hand, the modification in the transverse J/ψ is weak without the mixing effect. The masses of the moving charmonia, which were obtained by subtracting the center-of-mass kinetic energy, tend to increase with an increasing in its momentum. The bottomoium spectra shown in Ref. [27] have qualitatively the same behavior as in charmonium spectra, with suppression in magnitude due to the smaller magnetic moments of the bottom quark and antiquark.
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F012.jpg)
This study was extended to the excited states of quarkonia, which was made possible by using the cylindrical Gaussian expansion method proposed in Ref. [38]. Since this method does not resort to the imaginary time evolution τ → ∞ to extract the spectra [27], it would be more suitable for investigating the excited states. The result of the systematic spectroscopy is shown in Fig. 13. In a weak magnetic field, the level repulsions between the nearest mixing partners give rise to the dominant modification of the mass spectra. As the magnetic field becomes stronger, there appear the level-crossing points where the wave functions are strongly mixed. The density plots in the small windows show the transition of the wave functions. One can see the squeezing of the wave function along the magnetic field in the strong-field regime as expected from the shrink of cyclotron orbits. Including the contributions of the excited states, the ηc state is, as it is the lowest-lying state, pushed down additively by all of the mixing partners, and its mass monotonically decreases. This systematic computation was very recently extended to the spectra of bottomonia and open heavy flavors [39].
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F013.jpg)
So far, we have not considered the modification of the Cornell potential by the magnetic field. Effects of a magnetic field on the potential was investigated in Ref. [32] by lattice QCD simulation. The authors extracted the strengths of the potential in two cases where the separation between the quark and antiquark is in parallel and transverse to the magnetic field. The left panel in Fig. 14 shows the results from the lattice simulation. Compared to the potential in vacuum, the strength of the potential is enhanced and suppressed in the transverse (XY) and longitudinal (Z) directions, respectively. This behavior may be understood from the fact that the screening due to the vacuum polarization of light-quark pairs is coupled to only the temporal and parallel (z) components of the gluon field. The full azimuthal-angle and temperature dependences were recently studied by the lattice QCD simulation [34], and those effects were also discussed by the strong couping method [35, 36, 213].
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F014.jpg)
The right panel in Fig. 14 shows the contour plot of the potential which was obtained by interpolating the values in the transverse (XY) and longitudinal (Z) directions with respect to the azimuthal-angle dependence. In Ref. [33], the quarkonium spectra were investigated by including this anisotropic potential as well as the mixing effects. Figure 15 shows the bottomonium spectra with and without the inclusion of the anisotropic potential. The authors argued that effects of the anisotropic potential can be seen in the bottomonium spectra more strongly than in the charmonium spectra, because the magnetic moment of the bottom quark is suppressed by the larger value of the quark mass and the smaller electric charge. Without the anisotropic potential, the spectra barely change due to the suppression of the mixing effects while the effects of the anisotropic potential act to increase the bottomonium masses in all three of the channels. The effect of the anisotropic potential is also much stronger than the momentum dependence of the spectra.
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F015.jpg)
On the basis of the potential model, several phenomenological consequences have also been discussed. Prior to moving to the next section, we briefly summarize these studies.
• In the presence of the mixing effect, the wave function of the mass eigenstate is the mixture of the two components so that the mass eigenstate shares the decay modes with both of the mixing partners. This means, for example, that the physical ηc and ηb are allowed to decay into the dilepton while the dilepton decay channels of the physical longitudinal J/ψ and Υ are suppressed [27]. As mentioned in Ref. [27], a similar modification of the positronium decay mode is known.
• The Lorentz forces acting on a moving quark and antiquark pair are oriented in the opposite direction. Therefore, the force will act to break up the bound state implying a possible dissociation mechanism in a strong magnetic field [23, 27]. This effect may be also understood as an effect of an electric field appearing in the rest frame of a moving quarkonium, which is connected to the original magnetic field by the Lorentz boost. The electric field will stretch the separation between the quark and antiquark pair and induces a tunneling through the binding potential [23].
• Mass shifts of D mesons were also examined with use of the potential model [25]. In turn, these mass shifts induce the shift of the
• A time dependence of the magnetic field was introduced in the potential model [37]. Based on the time-dependent model, the authors investigated the anisotropic production rate of high- pT charmonia. The feed-down effect from the excited states was also examined.
4.1.3 QCD sum rules in magnetic fields
Effects of the magnetic field on the meson spectra were also studied by using QCD sum rule (QCDSR) for open heavy flavor mesons [26, 31] and quarkonia [29, 30]. While we will focus here on the QCD sum rule for charmonia, extension to bottomonia is straightforward and extension to other systems will be discussed in the last of this section.
In the QCD sum rules [214-217], the spectra of bound states are extracted from the time-ordered current correlator defined by
where the superscript J specifies a channel. The dispersion relation relates the hadronic spectral function in the time-like region (q2=-Q2≥0) to the correlator in the deep Euclidean region (Q2 → ∞) as
One may evaluate the correlator on the left-hand side on the basis of the Operator Product Expansion (OPE) in the deep Euclidean region. Here, we shall focus on the charmonia created by the currents:
Up to the dimension-4 operators, the OPE can be written down as
The Wilson coefficients, C0, CG2, and CB2, are responsible for the short-distance processes, so that one may perturbatively calculate the diagrams for the hard quark loops with insertions of the soft external gluon/magnetic fields (see Refs. [29, 30] for explicit expressions). This can be done by using standard methods in the Fock-Schwinger gauge, which is also as known as the fixed-point gauge [216, 218].
On the other side of the dispersion relation (43), we assume an Ansatz for the functional form of the spectral function ρ(s)=π-1Im∏J(s) to extract the lowest-lying pole. In conventional analyses, the lowest-lying pole is simply separated from the continuum as
The parameters sth and f0 are the threshold of the continuum and the coupling strength between the current and the lowest-lying pole, respectively.
We, however, need to carefully reexamine the Ansatz (46) in the presence of a magnetic field, because effects of a magnetic field should be consistently taken into account on both sides of the dispersion relation (43) [26, 29-31]. The effect in the hadronic degrees of freedom is nothing but the mixing effect discussed in the previous sections. In the language of the spectral function, the mixing effect can be seen as the emergence of a new pole for the mixing partner, which is the longitudinal J/ψ in the pseudoscalar channel and the ηc in the vector channel. A schematic picture in Fig. 16 shows the emergence of the longitudinal J/ψ in the pseudoscalar channel, which induces the repulsive shifts of the poles and the reduction of the residue of ηc. It is important to include these effects in order to properly maintain the normalization of the spectral function.
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F016.jpg)
In the pseudoscalar channel, the second-order term in eB can be obtained by first converting the current to the psedoscalar meson with the strength f0 and then using the results for the mixing effect shown in Eqs. (30) and (31). Therefore, the correlator can be written in the hadronic degrees of freedom as [29]
Inserting Eqs. (30) and (31) into Eq. (47) and expanding up to the second order in eB, we find
Interpretation of the terms in Eq. (48) are as follows. The first term corresponds to production of the on-shell longitudinal J/ψ from the pseudoscalar current via an off-shell ηc. The second term with a negative sign is needed to preserve the normalization because the coupling of ηc to the current must be reduced to balance the emergence of the coupling to J/ψ. This is confirmed in Eq. (47), where these two terms come from overlaps between the properly normalized unperturbated and perturbated states obtained as
To make the dispersion integral (43) dominated by the pole contribution and to improve the convergence of the OPE, we use the Borel transform defined by
By applying the Borel transform to both sides of Eq. (43), the contribution of the delta function for the lowest-lying pole is picked up to result in a mass formula:
where
Figure 17 shows the mass spectra from the QCDSR together with those from the hadronic effective model shown in Fig. 11. First, the results from these two methods are consistent with each other, indicating the dominance of the mixing effect in the shown regime. They are also consistent with the results from the potential model discussed in the previous section [27, 38, 39, 33]. Also, the modification in the transverse components of J/ψ (green line) is much smaller than the others due to the absence of the mixing partner. On top of these observations, one finds deviations between the results from the two methods. Although they are small deviations in the current regime of the magnetic field, they suggest the precursor of other effects, which are not explained by the mixing effect. It would be natural to expect the emergence of such effects in stronger magnetic fields. Possible origin of these deviations were discussed in Ref. [30].
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F017.jpg)
One caveat is that the modification of the gluon condensate in a magnetic field has not been included in the evaluation of the OPE (45). Since the lattice QCD simulation has provided the values of the gluon condensate in finite magnetic field and temperature [201], it will be interesting to investigate effects of these modifications by using the QCDSR.
Also, applications of the QCDSR to light mesons and open heavy flavor mesons are an interesting topic because these mesons contain a light quark component, which reflects the modification of the quark condensates. In Refs. [26, 31], the OPE was performed by taking into account not only the modification of
4.2 Transport of heavy quarks in a hot medium and magnetic field
As mentioned in the beginning of this section, the heavy quarks are dominantly created by the initial hard scatterings among the partons from the colliding nuclei (see Fig. 4.2). Therefore, the heavy quarks typically have the power law pT spectrum according to the perturbative QCD. This spectrum is however gradually modified by the multiple interactions with the thermal quarks and gluons in the QGP, and the spectrum approaches a thermal distribution, which is realized at a large enough time. However, the QGP has a finite lifetime, and to what extent the thermalization is achieved depends on the transport property of the QGP. Therefore, the heavy quarks will probe the transport property of the QGP.
In the RHIC and LHC experiments, the nuclear modification factor, RAA, and the anisotropic spectrum, v2, of the open heavy flavors have been measured (see a review [202] and references therein). Both of them are thought to be sensitive to the heavy-quark thermalization and are closely related to each other [227]. However, a consistent theoretical modeling, which simultaneously reproduces RAA and v2, is still under investigation. Effects of the magnetic field could be new ingredients to resolve the discrepancy between the theoretical and experimental results. In this section, we summarize a recent study on the heavy quark dynamics in a hot medium and magnetic field on the basis of Langevin equation and the perturbative computation of the heavy-quark diffusion constant.
The dynamics of the low-energy heavy particles is modeled as a Brownian motion caused by random kicks by the thermal particles which is described by the Langevin equations [227]:
Since the external magnetic field provides a preferred spatial direction, we have a set of two equations for the heavy quark motions, one parallel and one perpendicular to the magnetic field that is oriented in the z-direction. The random forces are assumed to be white noises;
and these coefficients, κ∥ and κ⊥, are related to the drag coefficients, ≤∥ and ≤⊥, through the fluctuation-dissipation theorem as
Here, we would like to mention a recent numerical study [42] where the authors investigated the Lorentz force exerting on the heavy quark with the inclusion of the Lorentz force in the Langevin equation (51).
In addition to effects of the Lorentz force exerting on the heavy quark as a direct effect of the magnetic field, the diffusion and drag coefficients will be changed because the light quarks in the hot medium are strongly affected by the magnetic field [40, 41]. At the leading order in gs, the anisotropic momentum diffusion coefficients, κ∥ and κ⊥, can be defined by
where q is the amount of the momentum transfer from the thermal particles to the heavy quark, and the static limit (q0 → 0) is assumed in the above definitions. The momentum transfer rate
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F018.jpg)
In a strong magnetic field, the fermion wave function is strongly squeezed along the magnetic field, corresponding to the small radius of the cyclotron orbit. Indeed, from the Landau level quantization and the Zeeman effect, the dispersion relation of fermions in the lowest Landau level (LLL) becomes the (1+1) dimensional, i.e.,
(1) First, the Debye screening mass is obtained from the gluon self-energy computed by the standard method in finite temperature field theory in the (1+1) dimension. The gluon self-energy is completely factorized into the transverse and longitudinal parts as
where the factor of
Here, the longitudinal momentum is defined by
(2) The change in the dispersion relation of fermion scatterers is important for the kinematics of the Coulomb scattering process [40, 233]. We shall consider a massless fermion scatterer in the lowest Landau level (see Fig. 19). For the massless quarks to be on-shell both in the initial and final states, the energy transfer has to be
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F019.jpg)
Therefore, strong magnetic fields give rise to an anisotropy of the momentum diffusion constant.
Nonvanishing contributions to the longitudinal momentum transfer come from either the finite quark-mass correction or the contribution of the gluon scatterers. In Ref. [40], the contribution of the mass correction was obtained as
while the gluon contribution is obtained by substituting the Debye mass
We have assumed a hierarchy αseB≪T2≪eB. Complete expressions, including the prefactors, are found in Ref. [40]. It is instructive to compare the mass correction (58) with the gluon contribution (59). The ratio is written as
While the first two factors are small in our working regime, the last factor can be large. Therefore, the massive contribution
Now that we obtained the leading contributions in Eqs. (57) and (59), the anisotropy in the momentum diffusion constant and the drag force is estimated to be
This large anisotropy is induced by the enhancement of the density of states in the quark contribution
Based on the above findings, a toy model for the dynamical modeling was discussed in Ref. [40] by using the Fokker-Planck equation [227]. Since the friction exerted on the heavy quarks is stronger in the direction transverse to the magnetic field (see Fig. 20), the heavy-quark momentum in the final state will acquire an anisotropy. In the small momentum region, the heavy quarks will be dragged by the flow more strongly in the transverse direction while they will feel a larger resistance in the transverse direction as the momentum increases. Therefore, there is a turnover in the anisotropy of the heavy quark spectrum (see the right panel in Fig. 20).
-201702/1001-8042-28-02-012/alternativeImage/1001-8042-28-02-012-F020.jpg)
It will be interesting to implement the anisotropic diffusion and drag coefficient in the numerical simulations. As the magnetic field in the heavy-ion collisions is time dependent, it is also necessary to investigate the weak-field limit as the first step in considering the time dependence. These issues are left as open questions. Finally, the drag force in the presence of the anomalous currents was also investigated in Ref. [233, 235, 236].
5 Summary
In summary, we have discussed the recent progresses in understanding the novel quantum phenomena in heavy-ion collisions. We first briefly reviewed some special features of the magnetic fields generated in heavy-ion collisions from both numerical and analytical approaches in Sec. 2. The magnetic fields in RHIC Au + Au and in LHC Pb + Pb collisions can be comparable to or even larger than, the QCD confinement scale,
The interplay between the magnetic fields and the non-trivial topological configuration of QCD leads to a number of anomalous transports, for example, the chiral magnetic effect (CME), chiral separation effect (CSE), and chiral magnetic wave (CMW). These effects, once experimentally confirmed, provide "soft probes" of the topological sector of QCD in heavy-ion collisions. We discussed the basic mechanisms that lead to the occurrence of CME, CSE, and CMW in Sec. 3.1 and Sec. 3.3. The experimental search of CME and CMW in heavy-ion collisions has been performed for years, and the data already shows very encouraging features that are consistent with the expectations of CME and CMW. However, background effects, which are not related to CME and CMW, exist and make the interpretation of the experimental measurements ambiguous. More efforts from both experimental and theoretical sides are certainly needed before we can make any conclusive claims. We discussed the current experimental status of the search of CME and CMW in Sec. 3.2 and Sec. 3.4.
In the latter half of this short review, we discussed effects of the magnetic fields on the heavy quark sector, which are expected to be observed by hard probes in the heavy-ion collisions. In Sec. 4.1, we first discussed the quarkonium spectra in magnetic fields, which has close connection to the recent lattice QCD simulations on the properties of the QCD vacuum, especially the confinement potential. Extension to the finite temperature and/or density is left for the future study. Finally, in Sec. 4.2, we discussed a recent study on the heavy-quark transport in a hot medium and magnetic field on the basis of the Langevin equation. The momentum diffusion constant was computed by using the joint resummation for Hard Thermal Loop and strong magnetic field. While the computation has been performed for the lowest Landau level in the strong-field limit so far, extension to other regimes, e.g., weak field, and inclusion of the higher Landau levels are relevant for the phenomenology in the heavy-ion physics.
The topics covered by the present review are still fast developing, a number of important issues should be further investigated in future. Here we briefly discuss a few of them. 1) Although there are already thorough numerical simulations of the magnetic fields in the initial stage of the heavy-ion collisions, the precise time evolution of the magnetic fields is still unknown, see discussions in Sec. 2. The knowledge of the precise time evolution of the magnetic fields is very important for quantitative understanding of a number of magnetic-field induced phenomena, like the CME, CMW, etc. Thus detailed simulations of the time dependence of the magnetic field by using, for example, the parton cascade models or by using magnetohydrodynamics are very desirable. 2) The experimental measurements of the CME and CMW contain significant background contributions, see Sec. 3; to extract the CME or CMW signals we have to understand these background contributions at the quantitative level which still remains an unsolved problem and will be one of the most important directions to pursue in this field. 3) Heavy flavors will serve as an alternative probe to explore the dynamics in the strong magnetic field and to constrain the spacetime profile of the magnetic field. Therefore, it will be important to develop the numerical modelling of the heavy-flavor dynamics with the inclusion of the effects of the magnetic field. It is also interesting to investigate other transport coefficients. For example, the electrical conductivity will be an important input for the magnetohydrodynamics, which in turn will be relevant for the estimate of the magnetic-field profile. The techniques and ingredients developed in the computation of the heavy-quark diffusion coefficients have been applied in a few works for the jet quenching parameter [237] and the electrical conductivity [238, 239] in the strong-field limit. The other regime, regarding the strength of the magnetic field, still needs to be explored for the phenomenology of the heavy-ion collisions.
Magnetic splitting of quasimolecular electronic states in strong fields
.Phys. Rev. Lett. 36, 517 (1976).doi: 10.1103/PhysRevLett.36.517Properties of a pion condensate in a magnetic field
. Sov. Phys. JETP 51, 13-22 (1980).Quark - anti-quark condensates in strong magnetic fields
. Mod. Phys. Lett. A 7, 973-982 (1992). doi: 10.1142/S0217732392000860Exact Dirac propagator in a magnetic ’sheet’
. Phys. Lett. A 164, 28 (1992). doi: 10.1016/0375-9601(92)90900-7The Effects of topological charge change in heavy ion collisions: ’Event by event P and CP violation’
. Nucl. Phys.A 803, 227-253 (2008).doi: 10.1016/j.nuclphysa.2008.02.298Estimate of the magnetic field strength in heavy-ion collisions
. Int. J. Mod. Phys. A 24, 5925-5932 (2009).doi: 10.1142/S0217751X09047570(Electro-)Magnetic field evolution in relativistic heavy-ion collisions
.Phys. Rev. C 83, 054911 (2011).doi: 10.1103/PhysRevC.83.054911Event-by-event fluctuations of magnetic and electric fields in heavy ion collisions
. Phys. Lett.B 710, 171-174 (2012).doi: 10.1016/j.physletb.2012.02.065Magnetic effects in heavy-ion collisions at intermediate energies
. Phys. Rev. C 84, 064605 (2011).doi: 10.1103/PhysRevC.84.064605Event-by-event generation of electromagnetic fields in heavy-ion collisions
. Phys. Rev. C 85, 044907 (2012).doi: 10.1103/PhysRevC.85.044907Azimuthally fluctuating magnetic field and its impacts on observables in heavy-ion collisions
. Phys. Lett.B 718, 1529-1535 (2013).doi: 10.1016/j.physletb.2012.12.030Charge-dependent azimuthal correlations from AuAu to UU collisions
. Nucl. Phys. A 939, 85-100 (2015).doi: 10.1016/j.nuclphysa.2015.03.012A systematic study of magnetic field in Relativistic Heavy-ion Collisions in the RHIC and LHC energy regions
. Adv. High Energy Phys. 2014, 193039 (2014).doi: 10.1155/2014/193039Spatial distributions of magnetic field in the RHIC and LHC energy regions
. Chin. Phys. C 39, 104105 (2015).doi: 10.1088/1674-1137/39/10/104105Electromagnetic fields with electric and chiral magnetic conductivities in heavy ion collisions
. (2016).arXiv: 1602.02223Classical electromagnetic fields from quantum sources in Heavy-Ion collisions
. (2016).arXiv:1604.04572The chiral magnetic effect
. Phys. Rev. D 78, 074033 (2008).doi: 10.1103/PhysRevD.78.074033Quantum anomalies in dense matter
. Phys. Rev. D 70,074018 (2004).doi: 10.1103/PhysRevD.70.074018Anomalous axion interactions and topological currents in dense matter
. Phys. Rev. D 72,045011 (2005).doi: 10.1103/PhysRevD.72.045011Chiral magnetic wave
. Phys. Rev. D 83,085007 (2011).doi: 10.1103/PhysRevD.83.085007Normal ground state of dense relativistic matter in a magnetic field
. Phys. Rev. D 83, 085003 (2011).doi: 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).doi: 10.1103/PhysRevLett.107.052303Quarkonium dissociation in quark-gluon plasma via ionization in magnetic field
. Phys. Rev. C 84, 044908 (2011). doi: 10.1103/PhysRevC.84.044908J/ψ Production by Magnetic Excitation of ηc
. J. Phys. G 39, 015007 (2012). doi: 10.1088/0954-3899/39/1/015007Heavy quarkonium production in a strong magnetic field
. Phys. Rev. D 88, 034009 (2013). doi: 10.1103/PhysRevD.88.034009Modification of the B meson mass in a magnetic field from QCD sum rules
. Phys. Rev. D 89, 074027 (2014). doi: 10.1103/PhysRevD.89.074027Charmonia and bottomonia in a magnetic field
. Phys. Rev. D 88, 105017 (2013). doi: 10.1103/PhysRevD.88.105017Spin Asymmetry of J/psi in Peripheral Pb+Pb Collisions at LHC
. (2014). arXiv: 1403.4317QCD sum rules for magnetically induced mixing between ηc and J/ψ
. Phys. Rev. Lett. 113, 172301 (2014). doi: 10.1103/PhysRevLett.113.172301Charmonium spectroscopy in strong magnetic fields by QCD sum rules: S-wave ground states
. Phys. Rev. D 91, 045025 (2015). doi: 10.1103/PhysRevD.91.045025D mesons in a magnetic field
. Phys. Rev. D 93, 054026 (2016). doi: 10.1103/PhysRevD.93.054026Anisotropy of the quark-antiquark potential in a magnetic field
. Phys. Rev. D 89, 114502 (2014). doi: 10.1103/PhysRevD.89.114502Heavy quarkonia in strong magnetic fields
. Phys. Rev. D 92, 054014 (2015). doi: 10.1103/PhysRevD.92.054014Magnetic field effects on the static quark potential at zero and finite temperature
. Phys. Rev. D 94, 094007 (2016). doi: 10.1103/PhysRevD.94.094007Anisotropic heavy quark potential in strongly-coupled 𝒩=4 SYM in a magnetic field
. Phys. Rev. D 91, 066001 (2015). doi: 10.1103/PhysRevD.91.066001Melting of charmonium in a magnetic field from an effective AdS/QCD model
. Phys. Rev. D 91, 086002 (2015). doi: 10.1103/PhysRevD.91.086002Magnetic field effect on charmonium production in high energy nuclear collisions
. Phys. Lett. B 751, 215-219 (2015). doi: 10.1016/j.physletb.2015.10.038Cigar-shaped quarkonia under strong magnetic field
. Phys. Rev. D 93, 051502 (2016). doi: 10.1103/PhysRevD.93.051502Heavy meson spectroscopy under strong magnetic field
. Phys. Rev. D 94, 074043 (2016). doi: 10.1103/PhysRevD.94.074043Heavy quark diffusion in strong magnetic fields at weak coupling and implications for elliptic flow
. Phys. Rev. D 93, 074028 (2016). doi: 10.1103/PhysRevD.93.074028Momentum transport in strongly coupled anisotropic plasmas in the presence of strong magnetic fields
. (2016). arXiv: 1605.06061Directed flow of charm quarks induced by the initial magnetic field at LHC
. (2016) arXiv: 1608.02231Dimensional reduction and dynamical chiral symmetry breaking by a magnetic field in (3+1)-dimensions
. Phys. Lett.B 349, 477-483 (1995).doi: 10.1016/0370-2693(95)00232-ACatalysis of dynamical flavor symmetry breaking by a magnetic field in (2+1)-dimensions
. Phys. Rev. Lett. 73, 3499-3502 (1994).doi: 10.1103/PhysRevLett.73.3499Dimensional reduction and catalysis of dynamical symmetry breaking by a magnetic field
. Nucl. Phys. B 462, 249-290 (1996).doi: 10.1016/0550-3213(96)00021-1Magnetic catalysis: A review
. Lect. Notes Phys. 871, 13-49 (2013).doi: 10.1007/978-3-642-37305-3_2Inverse magnetic catalysis in dense holographic matter
. JHEP 03, 033 (2011).doi: 10.1007/JHEP03(2011)033The QCD phase diagram for external magnetic fields
. JHEP 02, 044 (2012). doi: 10.1007/JHEP02(2012)044QCD quark condensate in external magnetic fields
. Phys. Rev.D 86, 071502 (2012).doi: 10.1103/PhysRevD.86.071502Inverse magnetic catalysis and the Polyakov loop
. JHEP 04,112 (2013).doi: 10.1007/JHEP04(2013)112Magnetic catalysis in hot and dense quark matter and quantum fluctuations
. Phys. Rev. D 86,076013 (2012).doi: 10.1103/PhysRevD.86.076013Magnetic catalysis versus magnetic inhibition
. Phys. Rev. Lett. 110,031601 (2013).doi: 10.1103/PhysRevLett.110.031601The quark mass gap in a magnetic field
. Phys. Lett.B 720,192-197 (2013).doi: 10.1016/j.physletb.2013.02.024Inverse magnetic catalysis induced by sphalerons
. Phys. Rev. D 88,054009 (2013).doi: 10.1103/PhysRevD.88.054009Spontaneous generation of local CP violation and inverse magnetic catalysis
. Phys. Rev. D 90,074009 (2014).doi: 10.1103/PhysRevD.90.074009Magnetic and inverse magnetic catalysis in the Bose-Einstein condensation of neutral bound pairs
. Phys. Rev. D 92,065011 (2015).doi: 10.1103/PhysRevD.92.065011Inverse magnetic catalysis in the three-flavor NJL model with axial-vector interaction
. Phys. Rev. D 91,074011 (2015).doi: 10.1103/PhysRevD.91.074011Collective modes and Kosterlitz-Thouless transition in a magnetic field in the planar Nambu-Jona-Lasino model
. Phys. Rev. D 90,056005 (2014).doi: 10.1103/PhysRevD.90.056005Quark antiscreening at strong magnetic field and inverse magnetic catalysis
. Phys. Rev. D 91,054006 (2015).doi: 10.1103/PhysRevD.91.054006Delayed magnetic catalysis
. Phys. Lett.B 755, 265-269 (2016).doi: 10.1016/j.physletb.2016.02.017Magnetic catalysis and inverse magnetic catalysis in QCD
. Phys. Rev. D 91,116010 (2015).doi: 10.1103/PhysRevD.91.116010Mesons in strong magnetic fields: (I) General analyses
. Nucl. Phys. A 951, 1-30 (2016).doi: 10.1016/j.nuclphysa.2016.03.016Quark matter in a parallel electric and magnetic field background: Chiral phase transition and equilibration of chiral density
. Phys. Rev. D 93, 094021 (2016).doi: 10.1103/PhysRevD.93.094021Analogy between rotation and density for Dirac fermions in a magnetic field
. Phys. Rev. D 93,104052 (2016).doi: 10.1103/PhysRevD.93.104052Charge redistribution from anomalous magneto-vorticity coupling
. (2016). arXivi: 1607.01513Boundary effects and gapped dispersion in rotating fermionic matter
. (2016). arXivi: 1608.00336Superconductivity of QCD vacuum in strong magnetic field
. Phys. Rev. D 82,085011 (2010).doi: 10.1103/PhysRevD.82.085011Spontaneous electromagnetic superconductivity of vacuum in strong magnetic field: evidence from the Nambu-Jona-Lasinio model
. Phys. Rev. Lett. 106,142003 (2011).doi: 10.1103/PhysRevLett.106.142003Charged vector mesons in a strong magnetic field
. Phys. Rev. D 87,094502 (2013).doi: 10.1103/PhysRevD.87.094502Charged and neutral vector ρ mesons in a magnetic field
. Phys. Rev. D 91,014017 (2015).doi: 10.1103/PhysRevD.91.014017Survival of charged rho condensation at high temperature and density
. Chin. Phys. C 40,023102 (2016).doi: 10.1088/1674-1137/40/2/023102Electromagnetic triangle anomaly and neutral pion condensation in QCD vacuum
. Phys. Lett. B 757,1-5 (2016).doi: 10.1016/j.physletb.2016.03.066Transport processes in a plasma
. in Reviews of Plasma Physics, Vol. 1,Transport processes in a plasma
. Physcial Kinetics, Course of Theoretical Physics, Vol. 10,Anisotropic hydrodynamics, bulk viscosities and R-modes of strange quark stars with strong magnetic fields
. Phys. Rev.D 81,045015 (2010).doi: 10.1103/PhysRevD.81.045015Kubo formulae for relativistic fluids in strong magnetic fields
. Annals Phys. 326,3075-3094 (2011).doi: 10.1016/j.aop.2011.08.001On viscous flow and azimuthal anisotropy of quark-gluon plasma in strong magnetic field
. J. Phys. G 39,025010 (2012).doi: 10.1088/0954-3899/39/2/025010Hydrodynamic transport coefficients for the non-conformal quark-gluon plasma from holography
. JHEP, 02,051 (2015).doi: 10.1007/JHEP02(2015)051Synchrotron radiation by fast fermions in heavy-ion collisions
. Phys. Rev. C 82, 034904 (2010).doi: 10.1103/PhysRevC.82.034904Photon decay in strong magnetic field in heavy-ion collisions
. Phys. Rev. C 83, 017901 (2011).doi: 10.1103/PhysRevC.83.017901Electromagnetic radiation by quark-gluon plasma in a magnetic field
. Phys. Rev. C 87, 024912 (2013).doi: 10.1103/PhysRevC.87.024912Particle production in strong electromagnetic fields in relativistic heavy-ion collisions
. Adv. High Energy Phys. 2013, 490495 (2013).doi: 10.1155/2013/490495Synchrotron radiation of vector bosons at relativistic colliders
. Phys. Rev. D 91, 033004 (2015).doi: 10.1103/PhysRevD.91.033004Role of magnetic field in photon excess in heavy ion collisions
. Phys. Rev. C 91, 014902 (2015).doi: 10.1103/PhysRevC.91.014902Conformal anomaly as a source of soft photons in heavy ion collisions
. Phys. Rev. Lett. 109, 202303 (2012). doi: 10.1103/PhysRevLett.109.202303Wess-Zumino-Witten action and photons from the Chiral Magnetic Effect
. Phys. Rev.D 86 071501(2012).doi: 10.1103/PhysRevD.86.071501Neutral-pion reactions induced by chiral anomaly in strong magnetic fields
.(2013).arXiv: 1305.7224Thermal photon production from gluon fusion induced by magnetic fields in relativistic heavy-ion collisions
.(2016).arXiv: 1604.02713Effect of magnetic field on the photon radiation from quark-gluon plasma in heavy ion collisions
.(2016).arXiv: 1609.04324Strongly interacting matter in magnetic fields
. Lect. Notes Phys. 871, 1-624 (2013). doi: 10.1007/978-3-642-37305-3Phase diagram of QCD in a magnetic field: A review
. Rev. Mod. Phys. 88, 025001 (2016). doi: 10.1103/RevModPhys.88.025001Quantum field theory in a magnetic field: From quantum chromodynamics to graphene and Dirac semimetals
. Phys. Rept. 576, 1-209 (2015). doi: 10.1016/j.physrep.2015.02.003Electromagnetic fields and anomalous transports in heavy-ion collisions — A pedagogical review
. Rept. Prog. Phys. 79, 076302 (2016). doi: 10.1088/0034-4885/79/7/076302The McGill magnetar catalog
. Astrophys. J. Suppl. 212, 6 (2014). doi: 10.1088/0067-0049/212/1/6Magnetars: the physics behind observations
. A review. Rept. Prog. Phys. 78, 116901 (2015). doi: 10.1088/0034-4885/78/11/116901Transport coefficients in high temperature gauge theories. 2. Beyond leading log
. JHEP 05, 051 (2003). doi: 10.1088/1126-6708/2003/05/051The Electrical conductivity and soft photon emissivity of the QCD plasma
. Phys. Lett. B 597, 57-62 (2004). doi: 10.1016/j.physletb.2004.05.079Spectral functions at small energies and the electrical conductivity in hot, quenched lattice QCD
. Phys. Rev. Lett. 99, 022002 (2007). doi: 10.1103/PhysRevLett.99.022002Thermal dilepton rate and electrical conductivity: An analysis of vector current correlation functions in quenched lattice QCD
. Phys. Rev. D 83, 034504 (2011). doi: 10.1103/PhysRevD.83.034504On the temperature dependence of the electrical conductivity in hot quenched lattice QCD
. Prog. Part. Nucl. Phys. 67, 212-217 (2012). doi: 10.1016/j.ppnp.2011.12.020Vector spectral functions and transport properties in quenched QCD
.Thermal Correlators in the ρ channel of two-flavor QCD
. JHEP 03, 100 (2013). doi: 10.1007/JHEP03(2013)100Electrical conductivity of the quark-gluon plasma across the deconfinement transition
. Phys. Rev. Lett. 111, 172001 (2013). doi: 10.1103/PhysRevLett.111.172001Electrical conductivity and charge diffusion in thermal QCD from the lattice
. JHEP 02, 186 (2015). doi: 10.1007/JHEP02(2015)186Thermal dilepton rates and electrical conductivity of the QGP from the lattice
. Phys. Rev. D 94, 034504 (2016). doi: 10.1103/PhysRevD.94.034504Thermodynamics of strong-interaction matter from Lattice QCD
. Int. J. Mod. Phys. E 24, 1530007 (2015). doi: 10.1142/S0218301315300076Time and space dependence of the electromagnetic field in relativistic heavy-ion collisions
. Phys. Rev. C 88, 024911 (2013). doi: 10.1103/PhysRevC.88.024911Electromagnetic field and the chiral magnetic effect in the quark-gluon plasma
. Phys. Rev. C 91, 064902 (2015). doi: 10.1103/PhysRevC.91.064902Electromagnetic fields in high energy heavy-ion collisions
. Int. J. Mod. Phys. E 23, 1430001 (2014). doi: 10.1142/S021830131430001XInitial value problem for magnetic fields in heavy ion collisions
. Int. J. Mod. Phys. Rev. C 93, 014905 (2016). doi: 10.1103/PhysRevC.93.014905Magnetohydrodynamics, charged currents and directed flow in heavy ion collisions
. Phys. Rev. C 89, 054905 (2014). doi: 10.1103/PhysRevC.89.054905Electromagnetic response of quark?Cgluon plasma in heavy-ion collisions
. Phys. Lett. B 737, 262-266 (2014). doi: 10.1016/j.physletb.2014.08.068Comments about the electromagnetic field in heavy-ion collisions
. Nucl. Phys. A 929, 184-190 (2014). doi: 10.1016/j.nuclphysa.2014.05.008Estimation of electric conductivity of the quark gluon plasma via asymmetric heavy-ion collisions
. Phys. Rev. C 90, 021903 (2014). doi: 10.1103/PhysRevC.90.021903Electric fields and chiral magnetic effect in Cu+Au collisions
. Phys. Lett. B 742, 296-302 (2015). doi: 10.1016/j.physletb.2015.01.050Charge-dependent directed flow in asymmetric nuclear collisions
.Phys. Rev. C 90, 064903 (2014). doi: 10.1103/PhysRevC.90.064903Evidence for creation of strong electromagnetic fields in relativistic heavy-ion collisions
. (2016). arXiv: 1604.06231Separation of flow from the chiral magnetic effect in U + U collisions using spectator asymmetry
. Phys. Rev. C 92, 011902 (2015). doi: 10.1103/PhysRevC.92.011902Charge splitting of directed flow and charge-dependent effects in pion spectra in heavy ion collisions
.Charge-dependent directed flow in Cu+Au collisions at sNN = 200 GeV
. (2016). arXiv: 1608.04100Topologically induced local P and CP violation in QCD x QED
. Annals Phys. 325, 205-218 (2010). doi: 10.1016/j.aop.2009.11.002The chiral magnetic eEffect and anomaly-iInduced transport
. Prog. Part. Nucl. Phys. 75, 133-151 (2014). doi: 10.1016/j.ppnp.2014.01.002Views of the chiral magnetic effect
. Lect. Notes Phys. 871, 241-259 (2013). doi: 10.1007/978-3-642-37305-3_9Topology, magnetic field, and strongly interacting matter
. Ann. Rev. Nucl. Part. Sci. 65, 193-214 (2015). doi: 10.1146/annurev-nucl-102313-025420Anomalous transport effects and possible environmental symmetry 'violation' in heavy-ion collisions
. Pramana 84, 901-926 (2015). doi: 10.1007/s12043-015-0984-xChiral magnetic and vortical effects in high-energy nuclear collisions? A status report
. Prog. Part. Nucl. Phys. 88, 1-28 (2016). doi: 10.1016/j.ppnp.2016.01.001Equilibrium parity violating current in a magnetic field
. Phys. Rev. D 22, 3080-3084 (1980). doi: 10.1103/PhysRevD.22.3080Observation of the chiral magnetic effect in ZrTe5
. Nature Phys. 12, 550-554 (2016). doi: 10.1038/nphys3648Signature of the chiral anomaly in a Dirac semimetal: a current plume steered by a magnetic field
. Science 350, 413 (2015).Negative magnetoresistance without well-defined chirality in the Weyl semimetal TaP
. Nature Commun. 7, 1615 (2016). doi: 10.1038/ncomms11615Negative magnetoresistance without well-defined chirality in the Weyl semimetal TaP
. Phys. Rev. X 5, 031023 (2015). arXiv: 1506.06577Absence of equilibrium chiral magnetic effect
. Phys. Rev. D 93, 105036 (2016). doi: 10.1103/PhysRevD.93.105036Parity violation in hot QCD: Why it can happen, and how to look for it
. Phys. Lett. B 633, 260-264 (2006). doi: 10.1016/j.physletb.2005.11.075The anomalous transport of axial charge: topological vs non-topological fluctuations
. JHEP 09, 030 (2015). doi: 10.1007/JHEP09(2015)030Topology in the Weinberg-Salam theory
. Phys. Rev. D 28, 2019 (1983). doi: 10.1103/PhysRevD.28.2019A saddle point solution in the Weinberg-Salam theory
. Phys. Rev. D 30, 2212 (1984). doi: 10.1103/PhysRevD.30.2212Anomalous chirality fluctuations in the initial stage of heavy ion collisions and parity odd bubbles
. Phys. Lett. B 545, 298-306 (2002). doi: 10.1016/S0370-2693(02)02630-8Some features of the glasma
. Nucl. Phys. A 772, 200-212 (2006). doi: 10.1016/j.nuclphysa.2006.04.001The chiral magnetic effect in heavy-ion collisions from event-by-event anomalous hydrodynamics
. (2014). arXiv: 1412.0311Parity violation in hot QCD: How to detect it
. Phys. Rev. C 70, 057901 (2004). doi: 10.1103/PhysRevC.70.057901Azimuthal charged-particle correlations and possible local strong parity violation
. Phys. Rev. Lett. 103, 251601 (2009). doi: 10.1103/PhysRevLett.103.251601Observation of charge-dependent azimuthal correlations and possible local strong parity violation in heavy ion collisions
. Phys. Rev. C 81, 054908 (2010). doi: 10.1103/PhysRevC.81.054908Fluctuations of charge separation perpendicular to the event plane and local parity violation in sNN=200 GeV Au+Au collisions at the BNL Relativistic Heavy Ion Collider
. Phys. Rev. C 88, 064911 (2013). doi: 10.1103/PhysRevC.88.064911A New method for the experimental study of topological effects in the quark-gluon plasma
. Phys. Rev. C 83, 011901 (2011). doi: 10.1103/PhysRevC.83.011901Charge separation relative to the reaction plane in Pb-Pb collisions at sNN=2.76 TeV
. Phys. Rev. Lett. 110, 012301 (2013). doi: 10.1103/PhysRevLett.110.012301Search for chiral magnetic effects in high-energy nuclear collisions
.Search for chiral magnetic effects in high-energy nuclear collisions
Nucl. Phys. A 904-905, 248c-255c (2013). doi: 10.1016/j.nuclphysa.2013.01.069Beam-energy dependence of charge separation along the magnetic field in Au+Au collisions at RHIC
. Phys. Rev. Lett. 113, 052302 (2014). doi: 10.1103/PhysRevLett.113.052302Alternative contributions to the angular correlations observed at RHIC associated with parity fluctuations
. (2010). arXiv: 1002.1758Effects of momentum conservation and flow on angular correlations at RHIC
. Phys. Rev. C 84, 024909 (2011). doi: 10.1103/PhysRevC.84.024909Azimuthal correlations from transverse momentum conservation and possible local parity violation
. Phys. Rev. C 83, 014905 (2011). doi: 10.1103/PhysRevC.83.014905Effects of charge conservation and flow on fluctuations of parity-odd observables ar RHIC
. (2010) arXiv: 1005.5341Charge conservation at energies available at the BNL Relativistic Heavy Ion Collider and contributions to local parity violation observables
. Phys. Rev. C 83, 014913 (2011). doi: 10.1103/PhysRevC.83.014913Procedure for removing flow backgrounds from the charge-separation observable perpendicular to the reaction plane in heavy-ion collisions
. (2016). arXivi: 1608.03205Challenges in flow background removal in search for the chiral magnetic effect
. (2016). arXivi: 1608.06610Testing the chiral magnetic effect with central U+U collisions
. Phys. Rev. Lett. 105, 172301 (2010). doi: 10.1103/PhysRevLett.105.172301Azimuthal anisotropy in U+U and Au+Au collisions at RHIC
. Phys. Rev. Lett. 115, 222301 (2015). doi: 10.1103/PhysRevLett.115.222301harge asymmetry dependency of π/K anisotropic flow in U+U sNN= 193 GeV and Au+Au sNN= 200 GeV collisions at STAR
. J. Phys. Conf. Ser. 509, 012033 (2014). doi: 10.1088/1742-6596/509/1/012033Test the chiral magnetic effect with isobaric collisions
. (2016) arXiv: 1607.04697Chiral magnetic effect task force report
. (2016) arXiv: 1608.00982Observation of charge-dependent azimuthal correlations in pPb collisions and its implication for the search for the chiral magnetic effect
. (2016). arXivi: 1610.00263Response of strongly-interacting matter to magnetic field: Some exact results
. Phys. Rev. D 73, 045006 (2006). doi: 10.1103/PhysRevD.73.045006Magnetic properties of dense holographic QCD
. Phys. Rev. D 79, 105024 (2009). doi: 10.1103/PhysRevD.79.105024Chiral asymmetry of the Fermi surface in dense relativistic matter in a magnetic field
. Phys. Rev. C 80, 032801 (2009). doi: 10.1103/PhysRevC.80.032801Chiral magnetic spiral
. Phys. Rev. Lett. 104, 232301 (2010). doi: 10.1103/PhysRevLett.104.232301Chiral asymmetry and axial anomaly in magnetized relativistic matter
. Phys. Lett. B 695, 354-358(2011). doi: 10.1016/j.physletb.2010.11.022Anomalous currents in dense matter under a magneticield
. Phys. Lett. B 699, 305-308(2011). doi: 10.1016/j.physletb.2011.04.010Gravitational anomaly and transport
. Phys. Rev. Lett. 107, 021601 (2011). doi: 10.1103/PhysRevLett.107.021601Radiative corrections to chiral separation effect in QED
. Phys. Rev. D 88, 025025 (2013). doi: 10.1103/PhysRevD.88.025025Generalized Bloch theorem and chiral transport phenomena
. Phys. Rev. D 92, 085011 (2015). doi: 10.1103/PhysRevD.92.085011Simulating chiral magnetic and separation effects with spin-orbit coupled atomic gases
. Sci. Rep. 6, 20601 (2016). doi: 10.1038/srep20601Collective modes of chiral kinetic theory in a magnetic field
. Phys. Rev. D 91, 125014 (2015). doi: 10.1103/PhysRevD.91.125014From the chiral magnetic wave to the charge dependence of elliptic flow
. (2012). arXiv: 1208.2537Chiral magnetic wave in an expanding QCD fluid
. Phys. Rev. C 91, 024902 (2015). doi: 10.1103/PhysRevC.91.024902Realistic implementation of chiral magnetic wave in heavy ion collisions
. Phys. Rev. C 89, 044909 (2014). doi: 10.1103/PhysRevC.89.044909Charge asymmetry dependency of π+/π- elliptic flow in Au + Au collisions at sNN =200 GeV
. J. Phys. Conf. Ser. 389, 012035 (2012). doi: 10.1088/1742-6596/389/1/012035Observation of charge asymmetry dependence of pion elliptic flow and the possible chiral magnetic wave in heavy-ion collisions
. Phys. Rev. Lett. 114, 252302 (2015). doi: 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. C 93, 044903 (2016). doi: 10.1103/PhysRevC.93.044903Charged elliptic flow at zero charge asymmetry
. Phys. Rev. C 88, 014908 (2013). doi: 10.1103/PhysRevC.88.014908Constituent quark scaling violation due to baryon number transport
. Phys. Rev. C 84, 044914 (2011). doi: 10.1103/PhysRevC.84.044914Effects of hadronic potentials on elliptic flows in relativistic heavy ion collisions
. Phys. Rev. C 85, 041901 (2012). doi: 10.1103/PhysRevC.85.041901Partonic mean-field effects on matter and antimatter elliptic flows
. Nucl. Phys. A 928, 234-246 (2014). doi: 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. B 726, 239-243 (2013). doi: 10.1016/j.physletb.2013.08.003Elliptic flow difference of charged pions in heavy-ion collisions
. Nucl. Phys. A 947, 155-160 (2016). doi: 10.1016/j.nuclphysa.2015.12.009Axial current generation from electric field: Chiral electric separation effect
. Phys. Rev. Lett. 110, 232302 (2013). doi: 10.1103/PhysRevLett.110.232302Chiral electric separation effect in the quark-gluon plasma
. Phys. Rev. D 91, 045001 (2015). doi: 10.1103/PhysRevD.91.045001Holographic chiral electric separation effect
. Phys. Rev. D 89, 085024 (2014). doi: 10.1103/PhysRevD.89.085024Chiral hall effect and chiral electric waves
. Phys. Rev. D 91, 025011 (2015). doi: 10.1103/PhysRevD.91.025011Possible observables for the chiral electric separation effect in Cu?+?Au collisions
. Phys. Rev. C 91, 054901 (2015). doi: 10.1103/PhysRevC.91.054901Charge separation induced by P-odd bubbles in QCD matter
. Nucl. Phys. A 797, 67-79 (2007). doi: 10.1016/j.nuclphysa.2007.10.001Fluid dynamics of R-charged black holes
. JHEP 01, 055 (2009). doi: 10.1088/1126-6708/2009/01/055Hydrodynamics from charged black branes
. JHEP 01, 094 (2011). doi: 10.1007/JHEP01(2011)094Hydrodynamics with triangle anomalies
. Phys. Rev. Lett. 103, 191601 (2009). doi: 10.1103/PhysRevLett.103.191601Chiral vortical wave and induced flavor charge transport in a rotating quark-gluon plasma
. Phys. Rev. D 92, 071501 (2015). doi: 10.1103/PhysRevD.92.07150Flow vorticity in peripheral high energy heavy ion collisions
. Phys. Rev. C 87, 034906 (2013). doi: 10.1103/PhysRevC.87.034906A study of vorticity formation in high energy nuclear collisions
. Eur. Phys. J. C 75, 406 (2015). doi: 10.1140/epjc/s10052-015-3624-1Rotating quark-gluon plasma in relativistic heavy ion collisions
. (2016). arXiv: 1602.06580Vorticity in heavy-ion collisions
. Phys. Rev. C 93, 064907 (2016). doi: 10.1103/PhysRevC.93.06490Testing the chiral magnetic and chiral vortical effects in heavy ion collisions
. Phys. Rev. Lett. 106, 062301 (2011). doi: 10.1103/PhysRevLett.106.062301Λ(KS0)?Ch± and Λ-p azimuthal correlations with respect to event plane and search for chiral magnetic and vortical effects
. Nucl. Phys. A 931, 746-751 (2014). doi: 10.1016/j.nuclphysa.2014.08.108Chiral properties of strong interactions in a magnetic background
.Phys. Rev.D 83, 114028 (2011).doi: 10.1103/PhysRevD.83.114028Magnetic field-induced gluonic (inverse) catalysis and pressure (an)isotropy in QCD
. JHEP 04,130 (2013).doi: 10.1007/JHEP04(2013)130Heavy-flavour and quarkonium production in the LHC era: from proton?proton to heavy-ion collisions
. Eur. Phys. J. C 76, 107 (2016). doi: 10.1140/epjc/s10052-015-3819-5Asymptotic freedom in strong magnetic fields
. Phys. Rev. Lett. 110, 162002 (2013). doi: 10.1103/PhysRevLett.110.162002Meson spectrum in strong magnetic fields
. Phys. Rev. D 87, 094029 (2013). doi: 10.1103/PhysRevD.87.094029QCD spectroscopy and quark mass renormalisation in external magnetic fields with Wilson fermions
. PoS, 265 (2016). arXiv: 1510.03899 doi: 10.1103/PhysRevD.87.094029Magnetic polarizabilities of light mesons in SU(3) lattice gauge theory
. Nucl. Phys. B 898, 627-643 (2015). doi: 10.1016/j.nuclphysb.2015.07.023Hadron masses in strong magnetic fields
. Phys. Rev. D 92, 014038 (2015). doi: 10.1103/PhysRevD.92.014038Model-independent study of electric dipole transitions in quarkonium
. Phys. Rev. D 85, 094005 (2012). doi: 10.1103/PhysRevD.85.094005Improved determination of heavy quarkonium magnetic dipole transitions in potential nonrelativistic QCD
. Phys. Rev. D 87, 074024 (2013). doi: 10.1103/PhysRevD.87.074024Charmonium: The model
. Phys. Rev. D 17, 3090 (1978). doi: 10.1103/PhysRevD.17.3090Charmonium: Comparison with experiment
. Phys. Rev. D 21, 203 (1980). doi: 10.1103/PhysRevD.21.203On gauge invariance and vacuum polarization
. Phys. Rev. 82, 664-679 (1951). doi: 10.1103/PhysRev.82.664The charmonium dissociation in an "ganomalous wind"
. JHEP 01, 052 (2016). doi: 10.1007/JHEP01(2016)052QCD and resonance physics. theoretical foundations
. Nucl. Phys. B 147, 385-447 (1979). doi: 10.1016/0550-3213(79)90022-1QCD and resonance physics: Applications
. Nucl. Phys. B 147, 448-518 (1979). doi: 10.1016/0550-3213(79)90023-3Hadron properties from QCD sum rules
. Phys. Rept. 127, 1 (1985). doi: 10.1016/0370-1573(85)90065-1Snapshots of hadrons or the story of how the vacuum which are produced, live and die in the QCD vacuum
. Prog. Theor. Phys. Suppl. 131, 1-71 (1998). doi: 10.1143/PTPS.131.1Calculations in external fields in quantum chromodynamics
. Technical review. Fortsch. Phys. 32, 585 (1984).Magnetically induced QCD kondo effect
. (2015). arXiv: 1509.06966QCD Kondo effect: quark matter with heavy-flavor impurities
. Phys. Rev. D 92, 065003 (2015). doi: 10.1103/PhysRevD.92.065003Kondo phase diagram of quark matter
. (2016). arXiv: 1604.07208An effective field theory of QCD at high density
. Phys. Lett. B 473, 118-125 (2000). doi: 10.1016/S0370-2693(99)01472-0Aspects of high density effective theory in QCD
. Nucl. Phys. B 582, 451-476 (2000). doi: 10.1016/S0550-3213(00)00330-8Hard loops, soft loops, and high density effective field theory
. Nucl. Phys. A 728, 251-271 (2003). doi: 10.1016/j.nuclphysa.2003.08.028RG analysis of magnetic catalysis in dynamical symmetry breaking
. Phys. Rev. D 54, 7879-7883 (1996). doi: 10.1103/PhysRevD.54.7879Magnetic catalysis in quantum electrodynamics
. Phys. Rev. D 57, 3759-3762 (1998). doi: 10.1103/PhysRevD.57.3759How much do heavy quarks thermalize in a heavy ion collision? Phys
. Rev. C 71, 064904 (2005). doi: 10.1103/PhysRevC.71.064904Symmetry behavior at finite temperature
. Phys. Rev. D 9, 3320-3341 (1974). doi: 10.1103/PhysRevD.9.3320On the axial anomaly at finite temperature in the Schwinger model
. Z. Phys. C 52, 339-342 (1991). doi: 10.1007/BF01560452Magnetic-field induced screening effect and collective excitations
. Phys. Rev. D 83, 11150 (2011). doi: 10.1103/PhysRevD.83.111501Vacuum birefringence in strong magnetic fields: (I) Photon polarization tensor with all the Landau levels
. Annals Phys. 330, 23-54 (2013). doi: 10.1016/j.aop.2012.11.010Vacuum birefringence in strong magnetic fields: (II)Complex refractive index from the lowest Landau level
. Annals Phys. 334, 58-82 (2013). doi: 10.1016/j.aop.2013.03.016Drag suppression in anomalous chiral media
. Phys. Rev. D 93, 125026 (2016). doi: 10.1103/PhysRevD.93.125026Heavy quark diffusion in perturbative QCD at next-to-leading order
. Phys. Rev. Lett. 100, 052301 (2008). doi: 10.1103/PhysRevLett.100.052301Chiral drag force
. JHEP 10, 018 (2015). doi: 10.1007/JHEP10(2015)018No-drag frame for anomalous chiral fluid
. Phys. Rev. Lett. 116, 122302 (2016). doi: 10.1103/PhysRevLett.116.122302Jet quenching parameter of the quark-gluon plasma in a strong magnetic field: Perturbative QCD and AdS/CFT correspondence
.Phys. Rev.D 94 085016, (2016).doi: 10.1103/PhysRevD.94.085016Longitudinal conductivity in strong magnetic field in perturbative QCD: Complete Leading Order
. 2016, arXiv:1610.06839Electrical conductivity of Quark-Gluon plasma in strong magnetic fields
. 2016, arXiv:1610.06818Quarkonium hadron interactions in QCD
. Phys. Rev. C 65, 067901 (2002). doi: 10.1103/PhysRevC.65.067901Quarkonium-hadron interactions in perturbative QCD
. Phys. Rev. D 72, 034002 (2005). doi: 10.1103/PhysRevD.72.034002