Search for the chiral magnetic effect in heavy ion collisions

Special Section on International Workshop on Nuclear Dynamics in Heavy-Ion Reactions (IWND2018)

Search for the chiral magnetic effect in heavy ion collisions

Fu-Qiang Wang
Jie Zhao
Nuclear Science and TechniquesVol.29, No.12Article number 179Published in print 01 Dec 2018Available online 23 Nov 2018
7001

Quark interactions with topological gluon fields in quantum chromodynamics can yield local 𝒫 and 𝒞𝒫 violations that could explain the matter–antimatter asymmetry in our universe. Effects of 𝒫 and 𝒞𝒫 violations can lead to charge separation under a strong magnetic field, a phenomenon called the chiral magnetic effect (CME). Early measurements of the CME-induced charge separation in heavy ion collisions are dominated by physics backgrounds. This report discusses the recent innovative efforts in eliminating those backgrounds, namely, by event-shape engineering, invariant-mass dependence, and reaction and participant plane comparison. The background-free CME measurements using these novel methods are presented.

Heavy ion collisionsChiral magnetic effectAzimuthal correlatorFlow backgroundInvariant massReaction planeParticipant plane

1 Introduction and physics motivation

Our universe started from the Big Bang singularity with equal matter and antimatter [1]; however, it is dominated by matter today. This matter–antimatter asymmetry is caused by the charge conjugation parity ( 𝒞𝒫 ) violation, and a slight difference in the physical laws governing matter and antimatter [2, 3]. 𝒞𝒫 is violated in the weak interaction; however, the magnitude of the violation is too small to explain the present matter–antimatter asymmetry [4]. 𝒞𝒫 violation in the strong interaction of the early universe is required to explain this. 𝒞𝒫 violation is not prohibited in the strong interaction, but it has not been observed experimentally. This is called the strong 𝒞𝒫 problem [5], one of the remaining problems in physics. The problem can be solved if the 𝒞𝒫 symmetry is violated in local metastable domains of topological gluon fields with nonzero topological charges (winding numbers) because of vacuum fluctuations in quantum chromodynamics (QCD) [6-9]. The topological charge, QW, is proportional to the integral of the scalar product of the gluon (color) electric and magnetic fields over the local domain. Interactions with those topological gluon fields change the helicities of the quarks, thereby causing an imbalance between the left- and right-handed quarks, QW=NL-NR≠0, or a local parity (𝒫) violation [9-11]. Such an imbalance can exhibit experimental consequences if submerged in a sufficiently strong magnetic field (B). The quark spins will be locked, either parallel or anti-parallel to the magnetic field direction, depending on the quark charge. This would result in an experimentally observable charge separation in the final-state hadrons that are products of quark hadronization. This charge separation phenomenon is called the chiral magnetic effect (CME) [12]; see the illustration in Fig. 1.

Figure 1:
(Color online) Illustration of the CME. The red arrows denote the direction of momentum, the blue arrows the spin of the quarks. (1) Initially, there were as many left-handed as right-handed quarks. Owing to the strong magnetic field, the up and down quarks are in the lowest Landau level and can only move along the magnetic field. (2) The quarks interact with a topological gluon field with nonzero Qw, converting left-handed quarks into right-handed ones (in this case Qw < 0) by reversing the momentum direction. (3) The right-handed up quarks will move upward, and the right-handed down quarks will move downward, resulting in a charge separation. From Ref. [11].
pic

Relativistic heavy ion collisions provide an ideal environment for the realization of the CME, as illustrated in Fig. 1. The magnetic field produced by the fast moving spectator protons in the early times of Au + Au collisions at BNL’s Relativistic Heavy Ion Collider (RHIC) is of the order of B ~1015 Tesla; eB ~ 0.3mπ2, where is the pion mass [10-13]. The Laudau energy eB/mq∼1 GeV is much larger than the typical transverse momenta of the quarks (or the system temperature); therefore, they are locked in the lowest Laudau level. Here, mq∼ of a few MeV are the light quark masses under the approximate chiral symmetry, which is broken spontaneously under normal conditions but is likely restored in relativistic heavy ion collisions where the relevant degrees of freedom are current quarks and gluons [14-18]. The time variation of the magnetic field is less well understood; it can quickly die off with time when the relativistic nuclei pass by quickly, or it could sustain for a relatively long time in a conducting quark–gluon plasma produced in those collisions [19, 20]. Therefore, even if the physics of the CME is correct, its observation is not guaranteed. Meanwhile, an observation of the CME-induced charge separation would confirm several fundamental properties of QCD and solve the long-standing strong 𝒞𝒫 problem. It is therefore of paramount importance.

2 Early measurements and the background issue

Intensive efforts have been invested to search for the CME in heavy ion collisions at BNL’s Relativistic Heavy Ion Collider (RHIC) and CERN’s Large Hadron Collider (LHC) [21-23]. Among various observables [24-28], a typically used observable to measure the CME-induced charge separation in heavy ion collisions is the three-point correlator [29],

γcos(α+β2ψ), (1)

where α and β are the azimuthal angles of two particles, and ψ is the angle of the reaction plane (RP, span by the beam and impact parameter directions of the colliding nuclei, see Fig. 2 for an illustration). Charge separation along the magnetic field, which is perpendicular to ψ on average, would yield different values of γ for particle pairs of the same-sign (SS) and opposite-sign (OS) charges: γSS=-1 and γOS=+1, respectively; the values have opposite signs but equal magnitude.

Figure 2:
(Color online) Illustration of a noncentral heavy ion collision, where the overlap participant region is an ellipse (on the transverse plane) with anisotropic expansion (indicated by radial arrows), and a strong magnetic field pointing upward generated by the spectator protons. The reaction plane is defined by the impact parameter direction and the beam direction.
pic

Indeed, signals that are qualitatively consistent with the CME expectations have been observed [25, 30 -34]. Figure 3 shows the example results [32] from STAR in 200 GeV Au + Au collisions. Experimentally, the angle ψ is often reconstructed from the azimuthal distribution of the final-state particle density by the fact that the particle density is the largest along the short axis of the collision overlap geometry (see Fig. 2 )[35]. This is typically attributed to the hydrodynamic expansion of the high-density collision region, generating an elliptical flow (v2) [36, 37]. In particular, it is the second harmonic (typically labeled as ψ2 as in Fig. 3 )of the particle density azimuthal distribution, corrected for finite multiplicity resolution [35]. Because of the fluctuations of the nucleon positions in the colliding nuclei, the reconstructed ψ2 corresponds to the participant plane (PP). It unnecessarily coincides with the RP; however, it fluctuates about the RP event-by-event [38]. Meanwhile, the RP can be determined more accurately by the spectator neutrons measured by a zero-degree calorimeter (ZDC) [39] (typically labeled as ψ1, as in Fig. 3 )because of a slight side kick they received from the collision [40]. More discussions are presented in Section 3.3.

Figure 3:
(Color online) The azimuthal correlator γ measured with the first-order event plane ψ1 from the ZDCs and the second-order event plane from the time projection chamber (TPC) as functions of centrality in Au + Au collisions at sNN=200 GeV from STAR. The Y4 and Y7 represent the results from the 2004 and 2007 RHIC runs, respectively. Shaded areas for the results measured with ψ2 represent the systematic uncertainty of the event plane determination. Systematic uncertainties for the results with respect to ψ1 are negligible compared to the statistical ones shown. From Ref. [32].
pic

However, background correlations unrelated to the CME exist [41-48]. For example, transverse momentum conservation induces correlations among particles that are enhanced back to back pairs [42-46]. Because more pairs are emitted in the RP direction, the net effect of this background is negative, thus dragging the CME-induced γSS and γOS, originally symmetric about zero (as illustrated in Fig. 4a), both down in the negative direction (Fig. 4b). This background is, fortunately, independent of particle charges, thus affecting the SS and OS pairs equally and cancels in the difference,

Figure 4:
(Color online) Left: expected CME signals (Eq. 1) for opposite-sign (OS) and same-sign (SS) particle pairs; they have opposite signs but equal magnitude. Center: effect of momentum conservation is negative and equal for OS and SS. Right: effect of local charge conservation (e.g., neutral resonance decays) is positive and only applies to OS.
pic
ΔγγOSγSS. (2)

Experimental investigations have thus focused on the Δγ observable [21-23]; the CME would yield Δγ > 0.

Unfortunately, mundane physics that differ between SS and OS pairs exist. One such physics is resonance/cluster decays [41-46], that affect OS pairs more significantly than SS pairs (as illustrated in Fig. 4c). This background is positive and arises from the coupling of elliptical anisotropy v2 of resonances/clusters, and the angular correlations between their decay daughters (nonflow) [41, 42, 45]. Use ρ→πpπm as an example (Fig. 5). The effect on γOS from the decay of a ρ in the RP direction is identical to a back-to-back pair from the CME in the B direction perpendicular to the RP [49]. Because more ρ resonances exist in the RP direction than the perpendicular direction because of the finite v2 of the ρ, the overall effect on γOS is positive.

Figure 5:
(Color online) Illustration of the decay π+π- pair from a ρ moving in the RP direction exhibiting the same effect on the γ observable (Eq. 1) as a CME π+π- pair perpendicular to the RP.
pic

There are more sources of particle correlations except that from ρ decays, such as other resonances and jet correlations. We can generally refer to those as cluster correlations [41]. Mathematically, the background can be estimated by

ΔγbkgdNclust.Nπ2cos(α+β2ϕclust.)v2,clust., (3)

where Nclust and are the numbers of clusters and single-charge pions (Nπ+Nπ-), respectively, and v2,clust.≡〈cos2(ϕclust.-ψ)〉 is the v2 of the clusters [22, 23, 29, 49]. A simple estimate, again using the ρ resonance as an example, indicates that the background magnitude is Δγbkgd201002×0.65×0.1104 for mid-central Au + Au collisions, comparable to the experimental data in Fig. 3.

3 Innovative methods and new results

Undoubtedly, the early Δγ measurements [25, 30-34] are dominated by backgrounds. Many proposals and attempts have been realized to reduce or eliminate these backgrounds [24, 25, 49-53]. Examining Eq. 3, it is easy to identify methods to remove backgrounds. One is to measure the Δγ observable, where the elliptical anisotropy is zero. This has already been exploited in various data analyses [24, 52, 53] and is not a new method. The other is to measure where resonance contributions are small, or can be identified and removed [54, 55]. This has not been explored until recently [56-58]. The following subsections (Sect. 3.1 and Sect. 3.2) will discuss these two methods with the emphasis on the second one. The third innovative method [59-61], which will be discussed in Sect. 3.3, is not as obvious, but may present the best and most robust method to search for the CME [58].

3.1 Make the anisotropy vanish

This method comprises two variations. The central idea of the first variation is that "round" events with zero elliptical anisotropy (v2obs) exists owing to event-by-event statistical and dynamical fluctuations. This was exploited by STAR [24], where a charge asymmetry observable (similar to Δγ) was analyzed as a function of the observed event-by-event v2obs; both the Δγ and v2obs are calculated from the same group of particles (particles of interest, or POI). This is shown in Fig. 6. A clear linear dependence was observed, as expected from the background. The background-suppressed CME signal can be extracted from the intercept at v2obs=0. With the limited statistics from Run-4 data (taken in the year 2004 by STAR), the extracted intercept is consistent with the zero in the 200 GeV Au + Au collisions [24]. Analysis of higher statistics data from later runs indicates that the intercept is finite [62]. However, the intercept at v2obs=0 may still contain residual backgrounds because the backgrounds are owing to the non-vanishing v2,clust of the correlation sources (resonances/clusters) (see Eq. 3), and not the v2obs measured by the final-state charged hadrons. It was shown by a toy-model resonance simulation [49] that the resonance v2 is not zero when the event-by-event v2obs of the final-state hadrons is required to be zero. Even if one could ensure that the event-by-event v2 of one resonance type is zero, it is nearly impossible to ensure the event-by-event v2’s of all the background sources to be zero. Furthermore, the background in Eq. 3 is proportional to v2,clust only when cos(α+β-2ϕclust.) and cos2(ϕclust.-ψ) are factorized. This may not be the case because both depend on the transverse momentum (pT) of the cluster [49].

Figure 6:
(Color online) Charge multiplicity asymmetry correlation (Δ) as a function of the event-by-event anisotropy v2obs in 20–40% Au + Au collisions at sNN=200 GeV from Run-4 by STAR. Errors are statistical. From Ref. [24].
pic

The second variation of the method is to analyze the Δγ observable of the POI as a function of the flow vector magnitude (q2) [63] that is calculated not using the POI but the particles from different phase spaces [52, 53]. The q2 is closely related to v2, and this method is known as "event-shape engineering (ESE)" [63]. ALICE [52] divided their data in each collision centrality according to q2 and found the Δγ to be approximately proportional to the v2 of the POI; this is consistent with the background contributions. This is shown in Fig. 7. One could fit the data with the linear function in v2 and extract the possible CME signal by the fit intercept. However, within each relatively wide centrality bin, the magnetic field most probably varies. Thus, ALICE modeled B(v2) to extract the CME signal. The extracted signal is found to be smaller than 20% of the early, inclusive Δγ measurement at the 95% conference level [52].

Figure 7:
(Color online) The charged-particle density-scaled azimuthal correlator, Δγ·dNch/dη, as a function of v2 for q2 shape-selected events for various centrality classes in Pb + Pb collisions by ALICE. Error bars (shaded boxes) represent the statistical (systematic) uncertainties. [52].
pic

The attractive aspect of this second variation of the method is that it can maintain the magnetic field, and vary the event-by-event v2 [24, 64, 65]. CMS attempts to achieve this using narrow centrality bins [53] such that the extracted CME signal is less model dependent. The CMS results indicate that the CME signal is less than 7% of the inclusive Δγ at the 95% confidence level [53].

Because the ESE control knob q2 and the POI are from different phase spaces, a given q2 cut-bin samples a v2 distribution of the POI. The extrapolated zero average v2 of the POI likely corresponds to the zero average v2 of all particle species, including the CME background sources of resonances and clusters. This is clearly advantageous over the first variation of the method using the event-by-event v2obs. The disadvantage is that an extrapolation to v2=0 is required as the ESE q2 sampling does not cover the most important v2=0 region. A dependence of the backgrounds on v2 that is not strictly linear would introduce imprecision in the extracted CME signal.

3.2 Identify backgrounds by invariant mass

The particle pair invariant mass (minv) is a typical method to identify resonances. Until recently [54], minv had not been utilized to investigate the CME Δγ signal.

The upper panel in Fig. 8 shows the relative OS excess over the SS pion pairs in Run-11 Au + Au collisions from STAR [56 -58]. The pions are identified by the TPC and time-of-flight (TOF) detector within pseudorapidity and pT ranges of || < 1 and 0.2 < pT < 1.8 GeV/c, respectively. The resonance peaks of KS and ρ are clearly shown. The large increase toward the low-minv kinematic limit is owing to the acceptance edge effect, where the OS and SS pair acceptance differences of the detector are amplified [58, 66]. The lower panel shows the Δγ measurement as a function of minv. A clear peak at the KS mass is observed; a peak at the ρ mass is observable. Most π+π- pair resonances are below minv < 1.5 GeV/c2; at higher minv, the resonance contribution can be neglected. The easiest method to remove resonance contributions from Δγ is, therefore, to restrict the measurements to the large-minv region. Figure 9 shows the measured Δγ at minv > 1.5 GeV/c2, compared to the inclusive Δγ. The large-mass Δγ is significantly lower by a factor of 20, compared to the inclusive Δγ, and is consistent with zero within a 2σ standard deviation.

Figure 8:
(Color online) minv dependences of the relative OS excess over SS pairs (upper) and Δγ (lower) in 20–50% central Au + Au collisions at sNN=200 GeV from Run-11 by STAR. Errors are statistical. [58].
pic
Figure 9:
(Color online) Average Δγ at large pair mass minv > 1.5 GeV/c2, compared to the inclusive Δγ, as a function of centrality in 20–50% central Au + Au collisions at sNN=200 GeV from Run-11 by STAR. Errors are statistical. [58].
pic

The CME is a low-pT phenomenon and may not be appreciable at high minv, although theoretical calculations suggest that the CME survives at minv > 1.5 GeV/c [67]. To extract the CME signal in the low-minv region, one needs the minv dependence of the background contribution. STAR used the ESE technique [58], dividing events from each narrow centrality bin into two classes according to the event-by-event q2 [63]. Because the magnetic fields are approximately equal while the backgrounds differ, the Δγ(minv) difference between the two classes is a good estimate of the background shape. Figure 10 shows the ΔγA and ΔγB from such two q2 classes in the upper panel, and the difference ΔγAγB together with the inclusive Δγ of all events in the lower panel [58].

Figure 10:
(Color online) minv dependences of the Δγ in large and small q2 events (upper), and the Δγ difference between large and small q2 events together with the inclusive Δγ (lower) in 20–50% central Au + Au collisions at sNN=200 GeV from Run-16 by STAR. Errors are statistical. From Ref. [58].
pic

With the background shape given by ΔγAγB, the CME can be extracted from a two-component fit to the following form: Δγ=κΔγAγBγCME. The left panel in Fig. 11 shows Δγ as a function of ΔγAγB, where each data point corresponds to one minv bin in the lower panel of Fig 10 [58]. Only the minv > 0.4 GeV/c2 data points are included in Fig. 11 because the Δγ from the lower minv region is affected by the edge effects of the STAR TPC acceptance [58, 66]. As shown, a positive linear correlation exists between Δγ and ΔγAγB. However, because the same data were used in the measurements of Δγ and ΔγAγB, their statistical errors are correlated. To accommodate the statistical errors, one can simply fit the independent measurements of ΔγA against ΔγB, namely by

Figure 11:
(Color online) Δγ versus ΔγAγB (left), and ΔγA versus ΔγB (right) in 20–50% central Au + Au collisions at sNN=200 GeV from Run-16 by STAR. Each data point in the left (right) panel corresponds to one minv bin in the lower (upper) panel of Fig. 10; only the minv > 0.4 GeV/c2 data points are included. Errors are statistical. [58].
pic
ΔγA=bΔγB+(1b)ΔγCME, (4)

where b and ΔγCME are the fitting parameters. The right panel of Fig. 11 shows ΔγA against ΔγB, and the fit by Eq. 4 superimposed as the straight line [58]. The straight line superimposed on the left panel of Fig. 11 is the same fit to Eq. 4, converted properly. The parameter b reflects the relative background contribution in the large-q2 (large-v2) event class to that in the small-q2 (small-v2) event class; further, because the background increases with v2, the value of b is larger than unity. The CME signal ΔγCME obtained from the fit is consistent with zero.

In this fit model, unlike the simple ESE method described in Sect. 3.1, the background is not required to be strictly proportional to v2. Provided that the backgrounds are different for different q2 event classes, one can extract the background shape as a function of minv. The slope fit parameter in Eq. 4 indicates how good the linearity of the background is against v2 The fit model, however, assumes that the CME signal is independent of minv. The good fit quality shown in Fig. 10 indicates that this is a good assumption within the current statistical precision of the data.

3.3 Compare participant plane and reaction plane

The magnetic field is primarily produced by spectator protons; therefore, its direction is determined by the spectator plane. It is found that the spectator plane nearly coincides with the RP in heavy ion collisions except for highly central collisions [60]. The elliptic flow v2 is generated by the participants, and is therefore determined by the PP [38]. The PP and RP are correlated, but, owing to fluctuations [38], they are not identical. See the illustration in the left panel of Fig. 12. The CME-induced charge separation, driven by the magnetic field [11], will be the strongest along the direction perpendicular to the spectator plane, and will be weaker along the direction perpendicular to the PP. The reduction factor is determined by the opening angle between the two planes and equals to

Figure 12:
(Color online) Left: sketch of a heavy ion collision projected onto the transverse plane (perpendicular to the beam direction). ψRP is the RP (impact parameter, b) direction, ψPP the PP direction (of interacting nucleons, denoted by the solid circles), and ψB the magnetic field direction (primarily from spectator protons, denoted by the open circles together with spectator neutrons). Right: illustration of the "CME-background filter." In a single collision, a CME signal "along" the RP and a background "along" the PP are present. The RP and PP are not the same but differ by an opening angle factor, a =〈cos2(ψPP-ψRP)〉. With the "filter" RP, the background is reduced by a factor of a and the CME remains the same, whereas with the "filter" PP, the background remains the same and the CME is reduced by the same factor a.
pic
a=cos2(ψPPψRP). (5)

The v2-induced background, meanwhile, will be the largest in the Δγ measurement with respect to the PP, and will be reduced by the same factor a in the Δγ measurement with respect to the RP. In other words, the Δγ measurements with respect to the PP and RP contain different amounts of v2 backgrounds and CME signals. Thus, the two Δγ measurements can disentangle the background and the CME signal. This is illustrated in the right panel of Fig. 12; the PP and RP serve as two different filters for the v2 background and CME signal.

In the experiment, the spectator plane (or closely the RP) can be measured by the ZDCs because of the slight side kick to the spectators in the collision [40]. The PP is, as usual, measured by final-state particles, for example, by the TPC in the STAR experiment [68]. The Δγ measurements with respect to the RP and PP can readily present the CME signal fraction in the inclusive Δγ measurement [60]. It is noteworthy that the ZDCs measure only the spectator neutrons [69]; therefore, the measured first-order harmonic plane fluctuates about the RP. Similarly, the final-state particle measurement of the second-harmonic plane is affected by effects other than the elliptic flow [70] and fluctuates about the PP. However, our method does not require a precise determination of the RP and PP [60]. Provided that two experimentally assessable planes exist, onto which the projections of the magnetic field and the elliptic flow are the opposite, our method is robust and is not affected by the uncertainties in assessing the true RP and PP. The plane projection relationship is given by Eq. (5), where the ψPP and ψRP, in an experimental data analysis context should be regarded as the experimentally measured harmonic planes. Figure 13 shows the CME signal in terms of its fraction in the inclusive Δγ measurement as a function of centrality in 200 GeV Au + Au collisions [58]. The two sets of data points correspond to two different acceptance cuts in the analysis. The results are primarily consistent with zero.

Figure 13:
(Color online) Extracted fraction of potential CME signal as a function of collision centrality in 200 GeV Au + Au collisions by STAR, combining data from Run-11, Run-14, and Run-16. Error bars (horizontal caps) represent statistical (systematic) uncertainties. [58].
pic

4 Discussions and Summary

At the LHC, the CMS and ALICE experiments used the ESE method to measure the CME signal without flow background contamination. The CME signal was found to be less than 7% (CMS) [53] and 20% (ALICE) [52] of the inclusive Δγ measurement at the 95% confidence level.

At the RHIC, two novel methods—invariant mass dependence and comparative PP-RP measurements—have been developed recently. Figure 14 summarizes the current status of the CME signal from STAR in 20–50% central Au + Au collisions at sNN=200 GeV [58] using these two methods. The data indicate that the CME signal is small, of the order of a few percent of the inclusive Δγ measurement, with relatively large errors [58].

Figure 14:
(Color online) The possible CME signal, relative to the inclusive Δγ measurement, extracted from the PP–RP comparative measurements and the invariant mass method, in 20–50% central Au + Au collisions at sNN=200 GeV by STAR, with a total of 2.5 billion minimum-bias events combining Run-11, Run-14, and Run-16 data. Error bars (vertical caps) represent statistical (systematic) uncertainties. From [58].
pic

In summary, the CME resulted from local 𝒫 and 𝒞𝒫 violations caused by topological charge fluctuations in QCD. Relativistic heavy ion collisions provided an ideal environment to search for the CME with the strong color gluon field and electromagnetic field. An observation of the CME would confirm several fundamental properties of QCD and solve the strong 𝒞𝒫 problem responsible for the matter–antimatter asymmetry in today’s universe. Charge-dependent azimuthal correlations with respect to the RP (and PP) were sensitive to the CME, but were contaminated by major physics backgrounds arising from the coupling of resonance/cluster decays and their elliptic flows (v2). Intensive theoretical and experimental efforts have been devoted to eliminating those backgrounds and significant progresses have been made in this regard. We herein discussed three novel methods that could potentially measure the background-free CME: ESE, invariant-mass (minv) dependence, and comparative measurements with respect to the PP and RP. The current estimates on the strength of the possible CME signal are of the order of a few percent of the inclusive Δγ values, and 1–2σ standard deviation from zero. It is clear that the experimental challenges in the CME search are daunting, but the important physics warrants continued efforts.

References
[1] R. A. Alpher, H. Bethe, G. Gamow,

The origin of chemical elements

. Phys. Rev. 73, 803 (1948). doi: 10.1103/PhysRev.73.803
Baidu ScholarGoogle Scholar
[2] A. D. Sakharov,

Violation of CP Invariance, C asymmetry, and baryon asymmetry of the universe

. Pisma Zh. Eksp. Teor. Fiz. 5, 32 (1967)
A. D. Sakharov,

Violation of CP Invariance, C asymmetry, and baryon asymmetry of the universe

. [JETP Lett. 5, 24 (1967)]
A. D. Sakharov,

Violation of CP Invariance, C asymmetry, and baryon asymmetry of the universe

. [Sov. Phys. Usp. 34, no. 5, 392 (1991)]
A. D. Sakharov,

Violation of CP Invariance, C asymmetry, and baryon asymmetry of the universe

. [Usp. Fiz. Nauk 161, no. 5, 61 (1991)]. doi: 10.1070/PU1991v034n05ABEH002497
Baidu ScholarGoogle Scholar
[3] M. Dine, A. Kusenko,

The origin of the matter - antimatter asymmetry

. Rev. Mod. Phys. 76, 1 (2003) doi: 10.1103/RevModPhys.76.1
Baidu ScholarGoogle Scholar
[4] T. Mannel,

Theory and phenomenology of CP violation

. Nucl. Phys. Proc. Suppl. 167, 115 (2007). doi: 10.1016/j.nuclphysbps.2006.12.083
Baidu ScholarGoogle Scholar
[5] R. D. Peccei, H. R. Quinn,

CP conservation in the presence of instantons

. Phys. Rev. Lett. 38, 1440 (1977). doi: 10.1103/PhysRevLett.38.1440
Baidu ScholarGoogle Scholar
[6] T. D. Lee,

A theory of spontaneous T violation

. Phys. Rev. D 8, 1226 (1973). doi: 10.1103/PhysRevD.8.1226
Baidu ScholarGoogle Scholar
[7] T. D. Lee, G. C. Wick,

Vacuum stability and vacuum excitation in a spin 0 field theory

. Phys. Rev. D 9, 2291 (1974). doi: 10.1103/PhysRevD.9.2291
Baidu ScholarGoogle Scholar
[8] P. D. Morley, I. A. Schmidt,

Strong P, CP, T violations in heavy ion collisions

. Z. Phys. C 26, 627 (1985). doi: 10.1007/BF01551807
Baidu ScholarGoogle Scholar
[9] D. Kharzeev, R. Pisarski, M. H. Tytgat,

Possibility of spontaneous parity violation in hot QCD

. Phys.Rev.Lett. 81, 512 (1998). doi: 10.1103/PhysRevLett.81.512
Baidu ScholarGoogle Scholar
[10] D. Kharzeev,

Parity violation in hot QCD: Why it can happen, and how to look for it

. Phys. Lett. B 633, 260 (2006). doi: 10.1016/j.physletb.2005.11.075
Baidu ScholarGoogle Scholar
[11] D. E. Kharzeev, L. D. McLerran, H. J. Warringa,

The effects of topological charge change in heavy ion collisions: ’Event by event P and CP violation’

. Nucl. Phys. A 803, 227 (2008). doi: 10.1016/j.nuclphysa.2008.02.298
Baidu ScholarGoogle Scholar
[12] K. Fukushima, D. E. Kharzeev, H. J. Warringa,

The Chiral Magnetic Effect

. Phys. Rev. D 78, 074033 (2008). doi: 10.1103/PhysRevD.78.074033
Baidu ScholarGoogle Scholar
[13] B. Muller, A. Schafer,

Charge Fluctuations from the Chiral Magnetic Effect in Nuclear Collisions

. Phys. Rev. C 82, 057902 (2010). doi: 10.1103/PhysRevC.82.057902
Baidu ScholarGoogle Scholar
[14] J. Adams et al.[STAR Collaboration],

Experimental and theoretical challenges in the search for the quark gluon plasma: The STAR Collaboration’s critical assessment of the evidence from RHIC collisions

. Nucl. Phys. A 757, 102 (2005). doi: 10.1016/j.nuclphysa.2005.03.085
Baidu ScholarGoogle Scholar
[15] K. Adcox et al.[PHENIX Collaboration],

Formation of dense partonic matter in relativistic nucleus-nucleus collisions at RHIC: Experimental evaluation by the PHENIX collaboration

. Nucl. Phys. A 757, 184 (2005). doi: 10.1016/j.nuclphysa.2005.03.086
Baidu ScholarGoogle Scholar
[16] I. Arsene et al.[BRAHMS Collaboration],

Quark gluon plasma and color glass condensate at RHIC? The Perspective from the BRAHMS experiment

. Nucl. Phys. A 757, 1 (2005). doi: 10.1016/j.nuclphysa.2005.02.130
Baidu ScholarGoogle Scholar
[17] B. B. Back et al.,

The PHOBOS perspective on discoveries at RHIC

. Nucl. Phys. A 757, 28 (2005). doi: 10.1016/j.nuclphysa.2005.03.084
Baidu ScholarGoogle Scholar
[18] B. Muller, J. Schukraft, B. Wyslouch,

First Results from Pb+Pb collisions at the LHC

. Ann. Rev. Nucl. Part. Sci. 62, 361 (2012). doi: 10.1146/annurev-nucl-102711-094910
Baidu ScholarGoogle Scholar
[19] K. Tuchin,

Synchrotron radiation by fast fermions in heavy-ion collisions

. Phys. Rev. C 82, 034904 (2010). doi: 10.1103/PhysRevC.82.034904
Erratum:

Synchrotron radiation by fast fermions in heavy-ion collisions

[Phys. Rev. C 82, 034904 (2010)].
Erratum:

Synchrotron radiation by fast fermions in heavy-ion collisions

Phys. Rev. C 83, 039903 (2011) doi: 10.1103/PhysRevC.83.039903
Baidu ScholarGoogle Scholar
[20] D. She, S. Q. Feng, Y. Zhong, et al.,

Chiral magnetic currents with QGP medium response in heavy ion collisions at RHIC and LHC energies

. Eur. Phys. J. A 54, 48 (2018). doi: 10.1140/epja/i2018-12481-x
Baidu ScholarGoogle Scholar
[21] D. E. Kharzeev, J. Liao, S. A. Voloshin, et al.,

Chiral magnetic and vortical effects in high-energy nuclear collisions—A status report

. Prog. Part. Nucl. Phys. 88, 1 (2016). doi: 10.1016/j.ppnp.2016.01.001
Baidu ScholarGoogle Scholar
[22] J. Zhao,

Search for the chiral magnetic effect in relativistic heavy-ion collisions

. Int. J. Mod. Phys. A 33, 1830010 (2018). doi: 10.1142/S0217751X18300107
Baidu ScholarGoogle Scholar
[23] J. Zhao, Z. Tu, F. Wang,

Status of the Chiral Magnetic Effect Search in Relativistic Heavy-Ion Collisions

. arXiv:1807.05083 [nucl-ex].
Baidu ScholarGoogle Scholar
[24] L. Adamczyk et al.[STAR Collaboration],

Measurement of charge multiplicity asymmetry correlations in high-energy nucleus-nucleus collisions at sNN=200 GeV

. Phys. Rev. C 89, 044908 (2014). doi: 10.1103/PhysRevC.89.044908
Baidu ScholarGoogle Scholar
[25] N. N. Ajitanand, R. A. Lacey, A. Taranenko, et al.,

A New method for the experimental study of topological effects in the quark-gluon plasma

. Phys. Rev. C 83, 011901 (2011).
Baidu ScholarGoogle Scholar
[26] N. Magdy, S. Shi, J. Liao, et al.,

New correlator to detect and characterize the chiral magnetic effect

. Phys. Rev. C 97, 061901 (2018). doi: 10.1103/PhysRevC.97.061901
Baidu ScholarGoogle Scholar
[27] P. Bozek,

Azimuthal angle dependence of the charge imbalance from charge conservation effects

. Phys. Rev. C 97, 034907 (2018). doi: 10.1103/PhysRevC.97.034907
Baidu ScholarGoogle Scholar
[28] Y. Feng, J. Zhao, F. Wang,

Responses of the chiral-magnetic-effect-sensitive sine observable to resonance backgrounds in heavy-ion collisions

. Phys. Rev. C 98, 034904 (2018). doi: 10.1103/PhysRevC.98.034904
Baidu ScholarGoogle Scholar
[29] S. A. Voloshin,

Parity violation in hot QCD: How to detect it

. Phys. Rev. C 70, 057901 (2004). doi: 10.1103/PhysRevC.70.057901
Baidu ScholarGoogle Scholar
[30] B. I. Abelev et al.[STAR Collaboration],

Observation 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.81.054908
Baidu ScholarGoogle Scholar
[31] B. I. Abelev et al.[STAR Collaboration],

Azimuthal charged-particle correlations and possible local strong parity violation

. Phys. Rev. Lett. 103, 251601 (2009). doi: 10.1103/PhysRevLett.103.251601
Baidu ScholarGoogle Scholar
[32] L. Adamczyk et al.[STAR Collaboration],

Fluctuations 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, no. 6, 064911 (2013). doi: 10.1103/PhysRevC.88.064911
Baidu ScholarGoogle Scholar
[33] L. Adamczyk et al.[STAR Collaboration],

Beam-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.113.052302
Baidu ScholarGoogle Scholar
[34] B. Abelev et al.[ALICE Collaboration],

Charge 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.110.012301
Baidu ScholarGoogle Scholar
[35] A. M. Poskanzer, S. A. Voloshin,

Methods for analyzing anisotropic flow in relativistic nuclear collisions

. Phys. Rev. C 58, 1671 (1998). doi: 10.1103/PhysRevC.58.1671
Baidu ScholarGoogle Scholar
[36] J. Y. Ollitrault,

Anisotropy as a signature of transverse collective flow

. Phys. Rev. D 46, 229 (1992). doi: 10.1103/PhysRevD.46.229
Baidu ScholarGoogle Scholar
[37] U. Heinz and R. Snellings,

Collective flow and viscosity in relativistic heavy-ion collisions

. Ann. Rev. Nucl. Part. Sci. 63, 123 (2013). doi: 10.1146/annurev-nucl-102212-170540
Baidu ScholarGoogle Scholar
[38] B. Alver et al.[PHOBOS Collaboration],

System size, energy, pseudorapidity, and centrality dependence of elliptic flow

. Phys. Rev. Lett. 98, 242302 (2007). doi: 10.1103/PhysRevLett.98.242302
Baidu ScholarGoogle Scholar
[39] C. Adler, A. Denisov, E. Garcia, et al.,

The RHIC zero-degree calorimeters

. Nucl. Instrum. Meth. A 499, 433 (2003). doi: 10.1016/j.nima.2003.08.112
Baidu ScholarGoogle Scholar
[40] W. Reisdorf, H. G. Ritter,

Collective flow in heavy-ion collisions

. Ann. Rev. Nucl. Part. Sci. 47, 663 (1997). doi: 10.1146/annurev.nucl.47.1.663
Baidu ScholarGoogle Scholar
[41] F. Wang,

Effects of cluster particle correlations on local parity violation observables

. Phys. Rev. C 81, 064902 (2010). doi: 10.1103/PhysRevC.81.064902
Baidu ScholarGoogle Scholar
[42] A. Bzdak, V. Koch, J. Liao,

Remarks on possible local parity violation in heavy ion collisions

. Phys. Rev. C 81, 031901 (2010). doi: 10.1103/PhysRevC.81.031901
Baidu ScholarGoogle Scholar
[43] J. Liao, V. Koch, A. Bzdak,

On the charge separation effect in relativistic heavy ion collisions

. Phys. Rev. C 82, 054902 (2010). doi: 10.1103/PhysRevC.82.054902
Baidu ScholarGoogle Scholar
[44] A. Bzdak, V. Koch, J. Liao,

Azimuthal correlations from transverse momentum conservation and possible local parity violation

. Phys. Rev. C 83, 014905 (2011). doi: 10.1103/PhysRevC.83.014905
Baidu ScholarGoogle Scholar
[45] S. Schlichting, S. Pratt,

Charge 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.83.014913
Baidu ScholarGoogle Scholar
[46] S. Pratt, S. Schlichting, S. Gavin,

Effects of momentum conservation and flow on angular correlations at RHIC

. Phys. Rev. C 84, 024909 (2011). doi: 10.1103/PhysRevC.84.024909
Baidu ScholarGoogle Scholar
[47] H. Petersen, T. Renk, S. A. Bass,

Medium-modified jets and initial state fluctuations as sources of charge correlations measured at RHIC

. Phys. Rev. C 83, 014916 (2011). doi: 10.1103/PhysRevC.83.014916
Baidu ScholarGoogle Scholar
[48] V. D. Toneev, V. P. Konchakovski, V. Voronyuk, et al.,

Event-by-event background in estimates of the chiral magnetic effect

. Phys. Rev. C 86, 064907 (2012). doi: 10.1103/PhysRevC.86.064907
Baidu ScholarGoogle Scholar
[49] F. Wang, J. Zhao,

Challenges in flow background removal in search for the chiral magnetic effect

. Phys. Rev. C 95, no. 5, 051901 (2017). doi: 10.1103/PhysRevC.95.051901
Baidu ScholarGoogle Scholar
[50] A. Bzdak,

Suppression of elliptic flow induced correlations in an observable of possible local parity violation

. Phys. Rev. C 85, 044919 (2012). doi: 10.1103/PhysRevC.85.044919
Baidu ScholarGoogle Scholar
[51] F. Wen, J. Bryon, L. Wen, et al.,

Event-shape-engineering study of charge separation in heavy-ion collisions

. Chin. Phys. C 42, no. 1, 014001 (2018). doi: 10.1088/1674-1137/42/1/014001
Baidu ScholarGoogle Scholar
[52] S. Acharya et al.[ALICE Collaboration],

Constraining the magnitude of the Chiral Magnetic Effect with Event Shape Engineering in Pb-Pb collisions at sNN = 2.76 TeV

. Phys. Lett. B 777, 151 (2018). doi: 10.1016/j.physletb.2017.12.021
Baidu ScholarGoogle Scholar
[53] A. M. Sirunyan et al.[CMS Collaboration],

Constraints on the chiral magnetic effect using charge-dependent azimuthal correlations in pPb and PbPb collisions at the CERN Large Hadron Collider

. Phys. Rev. C 97, 044912 (2018). doi: 10.1103/PhysRevC.97.044912
Baidu ScholarGoogle Scholar
[54] J. Zhao, H. Li, F. Wang,

Isolating the chiral magnetic effect from backgrounds by pair invariant mass

. arXiv:1705.05410 [nucl-ex].
Baidu ScholarGoogle Scholar
[55] H. Li, J. Zhao, F. Wang,

A novel invariant mass method to isolate resonance backgrounds from the chiral magnetic effect

. arXiv:1808.03210 [nucl-ex].
Baidu ScholarGoogle Scholar
[56] J. Zhao [STAR Collaboration],

Chiral magnetic effect search in p+Au, d+Au and Au+Au collisions at RHIC

. EPJ Web Conf. 172, 01005 (2018). doi: 10.1051/epjconf/201817201005
Baidu ScholarGoogle Scholar
[57] J. Zhao[STAR Collaboration],

Chiral magnetic effect search in p(d)+Au, Au+Au collisions at RHIC

. Int. J. Mod. Phys. Conf. Ser. 46, 1860010 (2018). doi: 10.1142/S2010194518600108
Baidu ScholarGoogle Scholar
[58] J. Zhao [STAR Collaboration],

Measurements of the chiral magnetic effect with background isolation in 200 GeV Au+Au collisions at STAR

. arXiv:1807.09925 [nucl-ex].
Baidu ScholarGoogle Scholar
[59] H. J. Xu, X. Wang, H. Li, et al.,

Importance of isobar density distributions on the chiral magnetic effect search

. Phys. Rev. Lett. 121, 022301 (2018). doi: 10.1103/PhysRevLett.121.022301
Baidu ScholarGoogle Scholar
[60] H. j. Xu, J. Zhao, X. Wang, et al.,

Varying the chiral magnetic effect relative to flow in a single nucleus-nucleus collision

. Chin. Phys. C 42, 084103 (2018). doi: 10.1088/1674-1137/42/8/084103
Baidu ScholarGoogle Scholar
[61] H. j. Xu, J. Zhao, X. Wang, et al.,

Re-examining the premise of isobaric collisions and a novel method to measure the chiral magnetic effect

. arXiv:1808.00133 [nucl-th].
Baidu ScholarGoogle Scholar
[62] B. Tu,

Charge Asymmetry Correlations to Search for the Chiral Magnetic Effect from Beam Energy Scan by STAR. Kobe, Japan, Sep. 27 - Oct. 3, 2015

https://drupal.star.bnl.gov/STAR/presentations/qm2015/biao-tu NoStop
Baidu ScholarGoogle Scholar
[63] J. Schukraft, A. Timmins, S. A. Voloshin,

Ultra-relativistic nuclear collisions: event shape engineering

. Phys. Lett. B 719, 394 (2013). doi: 10.1016/j.physletb.2013.01.045
Baidu ScholarGoogle Scholar
[64] S. A. Voloshin,

Testing the chiral magnetic effect with central U+U collisions

. Phys. Rev. Lett. 105, 172301 (2010). doi: 10.1103/PhysRevLett.105.172301
Baidu ScholarGoogle Scholar
[65] S. Chatterjee and P. Tribedy,

Separation of flow from the chiral magnetic effect in U + U collisions using spectator asymmetry

. Phys. Rev. C 92, no. 1, 011902 (2015). doi: 10.1103/PhysRevC.92.011902
Baidu ScholarGoogle Scholar
[66] L. Adamczyk et al.[STAR Collaboration],

Measurements of dielectron production in Au+Au collisions at sNN = 200 GeV from the STAR experiment

. Phys. Rev. C 92, no. 2, 024912 (2015). doi: 10.1103/PhysRevC.92.024912
Baidu ScholarGoogle Scholar
[67] S. Shi, Y. Jiang, E. Lilleskov, et al.,

Anomalous chiral transport in heavy ion collisions from anomalous-viscous fluid dynamics

. Annals Phys. 394, 50 (2018). doi: 10.1016/j.aop.2018.04.026
Baidu ScholarGoogle Scholar
[68] M. Anderson, J. Berkovitz, W. Betts, et al.,

The Star time projection chamber: A Unique tool for studying high multiplicity events at RHIC

. Nucl. Instrum. Meth. A 499, 659 (2003). doi: 10.1016/S0168-9002(02)01964-2
Baidu ScholarGoogle Scholar
[69] C. Adler, A. Denisov, E. Garcia, et al.,

The RHIC zero degree calorimeter

. Nucl. Instrum. Meth. A 470, 488 (2001). doi: 10.1016/S0168-9002(01)00627-1
Baidu ScholarGoogle Scholar
[70] N. M. Abdelwahab et al.[STAR Collaboration],

Isolation of flow and nonflow correlations by two- and four-particle cumulant measurements of azimuthal harmonics in sNN= 200 GeV Au+Au collisions

. Phys. Lett. B 745, 40 (2015). doi: 10.1016/j.physletb.2015.04.033
Baidu ScholarGoogle Scholar