19
Ultracentral Collisions of Small and Deformed Systems at RHIC: UU, d Au, 9 Be Au, 9 Be 9 Be, 3 He 3 He, and 3 He Au Collisions J. Noronha-Hostler, 1, * N. Paladino, 1, S. Rao, 1, Matthew D. Sievert, 1, § and Douglas E. Wertepny 2, 1 Department of Physics and Astronomy, Rutgers University, Piscataway, NJ USA 08854 2 Department of Physics, Ben-Gurion University of the Negev, Beer-Sheva 84105, Israel (Dated: June 3, 2019) In this paper, we study a range of collision systems involving deformed ions and compare the elliptic and triangular flow harmonics produced in a hydrodynamics scenario versus a color glass condensate (CGC) scenario. For the hydrodynamics scenario, we generate initial conditions using TRENTO and work within a linear response approximation to obtain the final flow harmonics. For the CGC scenario, we use the explicit calculation of two-gluon correlations taken in the high-pT “(semi)dilute-(semi)dilute” regime to express the flow harmonics in terms of the density profile of the collision. We consider ultracentral collisions of deformed ions as a testbed for these comparisons because the difference between tip-on-tip and side-on-side collisions modifies the multiplicity depen- dence in both scenarios, even at zero impact parameter. We find significant qualitative differences in the multiplicity dependence obtained in the initial conditions+hydrodynamics scenario and the CGC scenario, allowing these collisions of deformed ions to be used as a powerful discriminator between models. We also find that sub-nucleonic fluctuations have a systematic effect on the elliptic and triangular flow harmonics which are most discriminating in 0 - 1% ultracentral symmetric col- lisions of small deformed ions and in 0 - 10% d 197 Au collisions. The collision systems we consider are 238 U 238 U, d 197 Au, 9 Be 197 Au, 9 Be 9 Be, 3 He 3 He, and 3 He 197 Au. I. INTRODUCTION Since the first confirmation of the existence of the Quark Gluon Plasma (QGP) in the early 2000’s, many new and unexpected properties have been discovered. Contrary to expectations, the QGP behaves as a nearly perfect fluid with nearly vanishing shear viscosity to en- tropy density ratio that is in the range of estimates from AdS/CFT [1–6]. Heavy-ion collision physicists originally thought that the largest possible ions would be needed to recreate this deconfined state of matter. However, more recent experiments have surprisingly discovered signals of the fluid-like nature of a QGP droplet in small systems comparable to the size of a proton (seen in pPb, pAu, dAu, 3 He Au, and possibly even pp collisions) [7–27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with experimental data [28–36], alternative pictures that do not rely on a tiny QGP appearing in small systems have emerged [37–40]. Probably the strongest contender is the color glass con- densate (CGC) effective theory, in which the proliferation of small-x gluons can produce many-body correlations and flow-like signals similar to relativistic hydrodynam- ics. A number of experiments have been conducted (and future ones proposed [41–45]) to disentangle these two scenarios; however, as of yet a strong smoking gun signal is still lacking. * Email: [email protected] Email: [email protected] Email: [email protected] § Email: [email protected] Email: [email protected] During this time, there has also been significant devel- opment in understanding of the influence of nuclear struc- ture on relativistic heavy-ion collisions [4, 46–50]. Ultra- central collisions can be used as a probe of nuclear defor- mations, as recently shown for both 129 Xe 129 Xe [4, 51– 53] and 238 U 238 U collisions [46–50]. In hydrodynamics, a quadrupole deformation of the nucleus enhances the elliptic flow signals in ultracentral collisions where the nuclei are almost entirely overlapping. Additionally, if one selects on the 0 - 1% most central collisions using the ZDC at STAR (i.e. events with the fewest number of spectator nucleons), a sensitivity is seen to tip-to-tip vs. side-to-side collision geometries [46]. Tip-to-tip collisions produce the smallest elliptic flow but the largest multi- plicity, while in contrast side-to-side collisions produce a larger elliptic flow but a smaller multiplicity. This anti- correlation between the magnitude of elliptic flow and the multiplicity arises naturally from the eccentricities, which is then seen in the final flow data due to nearly perfect linear response in ultracentral collisions. We will refer to this as the initial conditions+hydrodynamics scenario. In this scenario, particles are emitted from the freeze-out hypersurface independently, resulting in the mutual cor- relation of bulk particles through the dependence of the single-particle distribution on the event geometry. In contrast to the hydrodynamic picture where the re- sponse is driven by the shape of the collision geometry, gluon correlations in a color glass condensate picture are driven instead by the multiplicity via a dependence on the saturation scales of the colliding ions. Thus, in a color glass condensate scenario, the collision of two deformed ions would result in entirely the opposite behavior: a pos- itive correlation of elliptic flow with multiplicity. Tip-on- tip collisions would produce the greatest two-particle cor- relations owing to the largest saturation scales and mul- arXiv:1905.13323v1 [hep-ph] 30 May 2019

dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

Ultracentral Collisions of Small and Deformed Systems at RHIC:UU, d Au, 9Be Au, 9Be 9Be, 3He 3He, and 3He Au Collisions

J. Noronha-Hostler,1, ∗ N. Paladino,1, † S. Rao,1, ‡ Matthew D. Sievert,1, § and Douglas E. Wertepny2, ¶

1Department of Physics and Astronomy, Rutgers University, Piscataway, NJ USA 088542Department of Physics, Ben-Gurion University of the Negev, Beer-Sheva 84105, Israel

(Dated: June 3, 2019)

In this paper, we study a range of collision systems involving deformed ions and compare theelliptic and triangular flow harmonics produced in a hydrodynamics scenario versus a color glasscondensate (CGC) scenario. For the hydrodynamics scenario, we generate initial conditions usingTRENTO and work within a linear response approximation to obtain the final flow harmonics. Forthe CGC scenario, we use the explicit calculation of two-gluon correlations taken in the high-pT“(semi)dilute-(semi)dilute” regime to express the flow harmonics in terms of the density profile ofthe collision. We consider ultracentral collisions of deformed ions as a testbed for these comparisonsbecause the difference between tip-on-tip and side-on-side collisions modifies the multiplicity depen-dence in both scenarios, even at zero impact parameter. We find significant qualitative differencesin the multiplicity dependence obtained in the initial conditions+hydrodynamics scenario and theCGC scenario, allowing these collisions of deformed ions to be used as a powerful discriminatorbetween models. We also find that sub-nucleonic fluctuations have a systematic effect on the ellipticand triangular flow harmonics which are most discriminating in 0 − 1% ultracentral symmetric col-lisions of small deformed ions and in 0 − 10% d 197Au collisions. The collision systems we considerare 238U 238U, d 197Au, 9Be 197Au, 9Be 9Be, 3He 3He, and 3He 197Au.

I. INTRODUCTION

Since the first confirmation of the existence of theQuark Gluon Plasma (QGP) in the early 2000’s, manynew and unexpected properties have been discovered.Contrary to expectations, the QGP behaves as a nearlyperfect fluid with nearly vanishing shear viscosity to en-tropy density ratio that is in the range of estimates fromAdS/CFT [1–6]. Heavy-ion collision physicists originallythought that the largest possible ions would be needed torecreate this deconfined state of matter. However, morerecent experiments have surprisingly discovered signals ofthe fluid-like nature of a QGP droplet in small systemscomparable to the size of a proton (seen in pPb, pAu,dAu, 3He Au, and possibly even pp collisions) [7–27].

While the predictions of relativistic hydrodynamicshave been reasonably consistent with experimental data[28–36], alternative pictures that do not rely on a tinyQGP appearing in small systems have emerged [37–40].Probably the strongest contender is the color glass con-densate (CGC) effective theory, in which the proliferationof small-x gluons can produce many-body correlationsand flow-like signals similar to relativistic hydrodynam-ics. A number of experiments have been conducted (andfuture ones proposed [41–45]) to disentangle these twoscenarios; however, as of yet a strong smoking gun signalis still lacking.

∗ Email: [email protected]† Email: [email protected]‡ Email: [email protected]§ Email: [email protected]¶ Email: [email protected]

During this time, there has also been significant devel-opment in understanding of the influence of nuclear struc-ture on relativistic heavy-ion collisions [4, 46–50]. Ultra-central collisions can be used as a probe of nuclear defor-mations, as recently shown for both 129Xe 129Xe [4, 51–53] and 238U 238U collisions [46–50]. In hydrodynamics,a quadrupole deformation of the nucleus enhances theelliptic flow signals in ultracentral collisions where thenuclei are almost entirely overlapping. Additionally, ifone selects on the 0 − 1% most central collisions usingthe ZDC at STAR (i.e. events with the fewest number ofspectator nucleons), a sensitivity is seen to tip-to-tip vs.side-to-side collision geometries [46]. Tip-to-tip collisionsproduce the smallest elliptic flow but the largest multi-plicity, while in contrast side-to-side collisions produce alarger elliptic flow but a smaller multiplicity. This anti-correlation between the magnitude of elliptic flow and themultiplicity arises naturally from the eccentricities, whichis then seen in the final flow data due to nearly perfectlinear response in ultracentral collisions. We will referto this as the initial conditions+hydrodynamics scenario.In this scenario, particles are emitted from the freeze-outhypersurface independently, resulting in the mutual cor-relation of bulk particles through the dependence of thesingle-particle distribution on the event geometry.

In contrast to the hydrodynamic picture where the re-sponse is driven by the shape of the collision geometry,gluon correlations in a color glass condensate picture aredriven instead by the multiplicity via a dependence on thesaturation scales of the colliding ions. Thus, in a colorglass condensate scenario, the collision of two deformedions would result in entirely the opposite behavior: a pos-itive correlation of elliptic flow with multiplicity. Tip-on-tip collisions would produce the greatest two-particle cor-relations owing to the largest saturation scales and mul-

arX

iv:1

905.

1332

3v1

[he

p-ph

] 3

0 M

ay 2

019

Page 2: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

2

tiplicity, while side-on-side collisions would produce theleast for the same reasons [54]. The ability to distinguishbetween these qualitatively opposite dependences on themultiplicity makes ultracentral collisions of deformed ionsa powerful tool to study the origin of correlations. Incontrast, for truly round systems, such as ultracentralp Au or Au Au, the initial conditions+hydrodynamics re-sponse leads to a flattened dependence on multiplicity inthe absence of a geometry bias.

One should be clear that, aside from the above sce-nario where the final-state correlations are derived en-tirely from CGC dynamics, it is also possible to basethe initial conditions on a CGC picture where the ob-tained energy-momentum tensor is fed into relativistichydrodynamics. The resulting correlations will still fitinto the initial condition+hydrodynamics scenario be-cause the hydrodynamic response will still reflect the ge-ometry of the initial energy-momentum tensor via inde-pendent particle emission at freezeout.

Generally, the temperatures and densities achieved incentral symmetric collisions of large ions such as uraniumare expected to produce a QGP, so they are not naturalcandidates for explanations based on CGC correlationsalone. Thus, in this paper we study smaller ion-on-ioncollisions producing smaller multiplicities, such that ei-ther scenario could be viable. Here we explore both sym-metric collisions such as 9Be 9Be or 3He 3He and asym-metric collisions such as d 197Au and 3He Au. In thesesmall but deformed systems, we see that the predictionof opposite correlations between elliptic flow and mul-tiplicity for two scenarios persists. We thus argue thatdeformed ion-ion collisions may be a valuable testbed todiscriminate between correlations arising from the colorglass condensate vs. initial condition+hydrodynamicsscenarios.

Previous UU hydrodynamic predictions were made in[48–50] some of which we will discuss and compare withbelow. Additionally, photons [55] and eccentricities [56]in terms of multiparticle cumulants have also been stud-ied.

The rest of this paper is organized as follows. InSec. II we set up the calculation, detailing the sam-pling of initial conditions in Sec. II A, the CGC picturein Sec. II B, and the initial conditions+hydrodynamicspicture in Sec. II C. In Sec. III we present our resultsfor a variety of collision systems, including 238U 238U inSec. III A, d 197Au in Sec. III B, 9Be 197Au and 9Be 9Bein Sec. III C, and 3He 3He and 3He 197Au in Sec. III D. Fi-nally we summarize our conclusions in Sec. IV. We alsodetail the particular centrality binning method we useto select the 0 − 1% most ultracentral collisions in Ap-pendix A, the shape parameters used to describe 238U inAppendix B, and an analysis of the multiplicity depen-dence in the CGC picture in Appendix C.

II. CALCULATION OF AZIMUTHALANISOTROPIES

A. Sampling of Multiplicity and Nucleon Positions

All collisions here are generated by an adapted ver-sion of TRENTO 2.0 [48]. Within TRENTO the totalmultiplicity is calculated using

S = c

∫d2x⊥ f

(TA(~x⊥), TB(~x⊥)

)(1)

where TA, TB are the nuclear profile functions of collid-ing ions A and B and c is a phenomenological scalingconstant that can depend on collision species and center-of-mass energy, which can be fixed by the total parti-cle yields. For this paper, the precise value of c will beirrelevant since we will only consider quantities whereit cancels out. The effective reduced thickness functionf(A,B) can be chosen phenomenologically or determinedfrom theoretical considerations; we consider the two sce-narios

f(A,B) =√AB (p = 0) (2a)

f(A,B) = AB, (2b)

where the√AB configuration is one of the default “p =

0” options in TRENTO, and we have hard-coded a newlinear scaling to reflect the predicted scaling of the ini-tial energy density in the color glass condensate [57]. Forthis paper we vary the size of the nucleon Gaussian widthfrom σ = 0.3 − 0.51 fm (from [58] we expect a prefer-ence for σ = 0.3 fm in small systems) and we vary themultiplicity scaling factor k from k = 0.5 − 2 for sys-tems that have not yet been run yet. In systems thathave already been run we can fix k by direct compar-isons to the measured multiplicity distributions. For de-tails of the TRENTO parameters such as σ, k and c, see[6, 48, 58, 59].

The changes that we have made to TRENTO are asfollows:

• Inclusion of an option for linear TATB scaling (2b)

• Addition of the ions 9Be, 3He, and new parame-terizations of 238U .

• Inclusion in the output of the moments Iα of thenuclear profile functions for α = 1−3 (See Eq. (12)

• Addition of the subroutine “vUSPhydro” to pro-vide compatibility with Lagrangian hydrodynamiccodes.

We note that in the case of 3He [60] we have made a.hdf configuration file that can be run within TRENTOas well.

The v2.0 release of TRENTO contains an option forincluding sub-nucleonic fluctuations in the density pro-files with variable parameters. In various collision sys-tems considered in this paper, we will study the effect

Page 3: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

3

of turning on these sub-nucleonic fluctuations using theout-of-the-box parameters which were previously tunedto p 208Pb collisions [67] 1 . Those parameters are asfollows: p = 0, k = 0.19, nc = 6, w = 0.855 fm, rcp =0.81 fm, v = 0.43 fm, dmin = 0.81 fm.

B. Two-Gluon Correlations from the CGC

Mirroring the notation of Ref. [61], we can decomposethe two-particle multiplicity in a given event as

dN2

d2p1dy1 d2p2dy2=

dN1

d2p1dy1

dN1

d2p2dy2+ δ2(p1, p2) (3)

Here the first term, identified as “collective flow,” arisesfrom the consequences of coupling the single-particle dis-tribution dN1

d2p dy to characteristic directions in the colli-

sion, such as the orientation of the collision geometry orthe direction of the color electric field in a color domain[62, 63]. The second term δ2 describes the “non-flow”terms arising from genuine correlations between the par-ticles, such QCD scattering. The two-particle cumulantis defined as

(vn{2})2 ≡⟨ein(φ1−φ2)

⟩=

⟨∫p1p2

ein(φ1−φ2) dN2

d2p1 dy1 d2p2 dy2

⟩〈Npairs〉

. (4)

It is important to note that, while the average defined inEq. (4) is often written as an average of the ratio:⟨

1

Npairs

∫p1p2

ein(φ1−φ2) dN2

d2p1 dy1 d2p2 dy2

⟩,

in practice it is actually computed as in Eq. (4). The dif-ference between these quantities is embodied in the corre-lated fluctuations of multiplicity in tandem with the az-imuthal anisotropies. For small enough multiplicity bins,these differences in the two-particle sector may be numer-ically small, but they are conceptually different quantitiesand can affect the systematics of higher-order cumulants.

In particular, note that the definition of what is consid-ered “an event” in (3) is arbitrary and potentially model-dependent. In some CGC calculations, for instance, arandom sampling of color sources ρ is performed suchthat the “event by event fluctuations” include the fluctu-ations of these color fields. In others, the sampling is per-formed only at the level of the fluctuating event geometry,with the color fields already having been averaged to ob-tain color multipole operators, such as Wilson line dipolesand quadrupoles. These various (model-dependent) dis-tinctions can matter when computing the average of the

1 Note that some values quoted here have changed slightly fromRef. [67], as per private communications from the authors.

ratio, but for the experimentally relevant quantity (4),the definition of what constitutes “an event” is arbitraryand does not affect the result for vn{2}.

Expanding (4) using the parameterization (3) gives

(vn{2})2=

⟨∣∣∣∫p einφ dN1

d2p dy

∣∣∣2⟩〈Npairs〉

+

⟨∫p1p2

ein(φ1−φ2) δ2(p1, p2)⟩

〈Npairs〉, (5)

where the first term arises from “flow” and the secondarises from “non-flow.” For the present purposes, weconsider an “event” in the case of two-gluon productionin the CGC to be one randomly sampled collision geom-etry. Then in one event, the single-gluon distribution isisotropic

dN1

d2p dy=

1

2πpT

dN1

dpT dy(6)

such that the first (“flow”) term of (5) vanishes. Thenvn{2} is obtained entirely due to the genuine two-particlecorrelations δ2.

As derived in Ref. [64], when the transverse momentaare much larger than the saturation scale p2

T � Q2s, the

CGC correlations between two gluons emitted in heavy-light ion collisions can be written as:

δ2(p1, p2)L.O.=

(∫d2x⊥T

2A(~x⊥)T 2

B(~x⊥)

)f(p1, p2). (7)

This statement is true at lowest order in the (semi)dilute-dense power counting, and f is a known function of thetransverse momenta of the gluons which is manifestlysymmetric under ~p2⊥ → −~p2⊥, corresponding to ∆φ →∆φ + π. As such, (7) does not contribute to vn{2} forodd values of n.

Similarly, at the next order in the (semi)dilute-denseCGC power counting, the two-gluon correlations ex-pressed in Eq. (77) of Ref. [65] have a comparable form,

δ2(p1, p1)N.L.O.

=

(∫d2x⊥T

3A(~x⊥)T 3

B(~x⊥)

)g(p1, p2),

(8)

but at this order the function g(p1, p2) breaks the ∆φ→∆φ+ π symmetry present at lowest order. Thus Eq. (8)provides the first nonvanishing contribution to the cumu-lants cn{2} and cn{4} for odd values of n. 2

2 We note that the expressions (7) and (8) are derived underthe assumption that transverse gradients of the nuclear pro-file functions are small compared to perturbative length scales� 1/ΛQCD. This assumption is clearly valid for a smoothnuclear profile function, but for a Monte Carlo sampling withnucleon-scale fluctuations, its validity is far less clear.

Page 4: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

4

Thus, in a given event, the two-particle azimuthalmodulation for even harmonics 2n is given by(v(2n){2}

)2=

1

〈Npairs〉

⟨∫d2x⊥T

2A(~x⊥)T 2

B(~x⊥)

⟩×[∫

p1p2

ei(2n)(φ1−φ2) f(p1, p2)

]

≡ 1

〈Npairs〉

⟨∫d2x⊥T

2A(~x⊥)T 2

B(~x⊥)

⟩× f(2n),

(9)

where all the event-by-event fluctuations are contained inthe geometry profiles TA,B(~x⊥) and the azimuthal mod-ulation f(2n) is independent of the event. Note that alldependence on which even harmonic (2n) is chosen isalso contained entirely within the constant f(2n). Like-wise, for the odd harmonics (2n− 1), the event-by-eventazimuthal modulation takes the form(v(2n−1){2}

)2=

1

〈Npairs〉

⟨∫d2x⊥T

3A(~x⊥)T 3

B(~x⊥)

⟩×[∫

p1p2

ei(2n−1)(φ1−φ2) g(p1, p2)

]

≡ 1

〈Npairs〉

⟨∫d2x⊥T

3A(~x⊥)T 3

B(~x⊥)

⟩× g(2n−1).

(10)

While a complete determination of f(2n) and g(2n−1) arenecessary for the absolute normalization or the pT depen-dence of the cumulants vn{2}, they cancel out completelyfor the pT -integrated quantities considered here when ex-pressed in terms of ratios. We can express the even andodd harmonics in the same notation by defining

νn ≡ (nMod 2) + 2 =

{2 if n = even

3 if n = odd. (11)

Then the relevant quantities are the integrals of the nu-clear densities TA/B in a given event, raised to somepower α:

Iα ≡∫d2x⊥T

αA(~x⊥)TαB(~x⊥). (12)

Now we would like to use these quantities to com-pare the cumulants vn{2} for ultracentral collisions, e.g.0 − 1% centrality determined from the ZDC, versus asmaller re-binning of those same events e.g. in 20 sub-bins sorted by multiplicity. We will express the ratio ofthe cumulants in an individual sub-bin i to the full ul-tracentral selection cut as vin{2}/vn{2}, which we canexpress as

vin{2}vn{2}

=

√√√√√√⟨N2

tot

⟩0−1%⟨

N2tot

⟩i

⟨Iνn⟩i⟨

Iνn⟩

0−1%

, (13)

where we have approximated Npairs = Ntot(Ntot − 1) ≈N2

tot. The multiplicity Ntot is determined in TRENTOfrom Eq. (1) from one of the parameterizations (2) of thereduced thickness function. For both cases the multiplic-ity is proportional to one of the weighted integrals (12)as

Ntot ∝ Ir, (14)

with the exponent r = 1/2 corresponding to the the phe-nomenological “p = 0” setting (2a) and r = 1 correspond-ing to the new hard-coded configuration (2b) accordingto the scaling of the initial energy density in the CGC.With these expressions for the multiplicity, the cumu-lants (13) are now entirely controlled by moments of thenuclear profiles (12):

vin{2}vn{2}

=

√√√√√√⟨

(Ir)2⟩

0−1%⟨(Ir)2

⟩i

⟨Iνn⟩i⟨

Iνn⟩

0−1%

. (15)

We then generate a large number of collision geometriesfor a given choice of ion species, select the ultracentral 0−1% bin and a set of sub-bins i, and use (15) to determinethe multiplicity dependence of the cumulants in the CGCframework.

C. Collective Flow in Hydrodynamics

Hydrodynamics is, at its core, an initial value problem,with a key input (and a significant source of uncertainty)being the initial conditions for the hydrodynamic evolu-tion. Most models employ some sort of wounded nucleonpicture, including the event-by-event fluctuations of nu-cleon positions. In its most simplistic form, this corre-sponds to a Glauber model where the nucleon positionsare sampled from a Wood-Saxon density distribution

ρ = ρ0

[1 + exp

(r −R(θ)

a

)]−1

(16)

where in the case of a deformed nucleus, the radius is nota constant but varies with azimuthal angle:

R = R0

1 + β2Y20(θ)︸ ︷︷ ︸Quadrupole

+ β4Y40(θ)︸ ︷︷ ︸Hexadecapole

+ · · ·

(17)

More sophisticated models include some kind of pre-equilibrium dynamics, such as the classical Yang-Millsevolution of color fields employed in the IP-Glasma [5]code, and still others like TRENTO [6, 48] take an agnos-tic approach instead use phenomenological parameteriza-tions combining the nuclear thickness functions. For theproton, deuteron, and 3He a Woods-Saxon distribution isnot appropriate; we instead sample the nucleon positions

Page 5: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

5

a [fm] R0 [fm] β2 β4 Origin0.5 6.61 0.236 0.098 (this work)0.55 6.86 0.28 0.093 [50]0.6 6.81 0.236 0.098 [56, 70]0.6 6.81 0.28 0.093 [48]

TABLE I: Deformed Wood-Saxon parameters forUranium initial conditions tested in this paper.

TRENTO [48] comes with three default settings U(listed above) and U2 and U3, which are not shown.

using the Hulthen wave function [66] for the deuteronand the nucleonic configurations from Ref. [60] for 3He.

The above prescriptions describe the procedure forsampling the positions of nucleons in the colliding nu-clei, which are traditionally the degrees of freedom usedto determine the deposition of energy or entropy in theinitial conditions of hydrodynamics. More microscopictreatments of the initial conditions, however, can also in-clude fluctuations of nucleonic substructure which mimicin some fashion the partonic content of the nucleon. Inan updated version of TRENTO [67], for instance, nu-cleonic substructure is implemented and calibrated top Pb and Pb Pb collisions at LHC energies. As a wayto test the sensitivity to these subnucleonic fluctuations,we compare the results obtained by using standard nu-cleon degrees of freedom with an out-of-the-box applica-tion of TRENTO’s subnucleonic fluctuations to RHIC at√sNN = 200 GeV.We emphasize one caveat to this comparison: one may

expect differences both in the multiplicity fluctuationsand in the number of constituents at the lower beamenergies. A detailed recalibration to RHIC kinematicswould require the dedication of significant computationalresources which is beyond the scope of this paper. More-over, since we are proposing collisions of new ion species,in some cases there is no data currently available againstwhich such a recalibration could be performed. Finally,we note that previous papers have explored the ques-tion of sub-nucleonic fluctuations from a different an-gle by systematically smoothing out smaller scale struc-tures, finding almost no effect in the final flow observables[68, 69].

The eccentricities εn are defined with respect to thecenter of mass of the initial entropy deposition such that

En ≡ −〈∫rneinφs(r, φ)rdrdφ〉〈∫rns(r, φ)rdrdφ〉

≡ εneinΦn (18)

with E1 ≡ 0 by definition and the minus sign by conven-tion. In order to study the deformation of Uranium wecompare a number of different choices of the parametersa,R0, β2, β4 in Eqs. (16-17). Our choice of parametersare often based upon previous work, which is cited ac-cordingly in Table I. To the best of our knowledge, theparameterization used in Refs. [56, 70] is the most up-to-date according to the low-energy nuclear structure com-munity.

To a good approximation, the single-particleanisotropic flow harmonics in hydrodynamics

Vn =

⟨∫d2p dy einφ dN1

d2p dy

⟩〈Ntot〉

≡ vneinψn (19)

arise primarily from linear response [71–79] from the ini-tial eccentricities (18):

Vn ≈ κnEn (20)

for some complex coefficient κn. Non-linear deviationsfrom the linear proportionality (20) have also been shownto be relevant for the prediction of v2{4}/v2{2} [42, 80].The quality of the approximation (20) is described by thePearson coefficient

Qn ≡〈VnE∗n〉√

〈|Vn|2〉√〈|En|2〉

(21)

with Q2 → 1 when linear response is exact. As shownin Fig. 1, linear response (20) is a nearly perfect ap-proximation to the full hydrodynamic simulation in mostcentral 238U 238U collisions, independent of the geom-etry parameters. This comparison was performed bygenerating multiple different sets of initial conditionswithin TRENTO and then running them through theevent-by-event relativistic viscous hydrodynamic modelv-USPhydro [81, 82], using the same hydrodynamic pa-rameters as in [83] for Au Au collisions.

a=0.5, R=6.61 β2=0.236

a=0.55, R=6.86 β2=0.28

a=0.6, R=6.81 β2=0.236

a=0.6, R=6.81 β2=0.28

STAR Data UU 200GeV

trento+v-USPhydro

0 20 40 60 800.0

0.2

0.4

0.6

0.8

1.0

Centrality %

Q2

FIG. 1: (Color online) Pearson coefficient for UU 193GeV of various deformed Wood Saxon parameterization

of Uranium.

In [42] the Pearson coefficient Q2 was also studied ver-sus system size. It is true that the linear mapping doesnot hold quite as well for smaller Npart. However, it wasstill found that smaller systems have a better linear map-ping than larger systems with the same Npart; that is, acentral small system has a better linear mapping than aperipheral larger system. Since we consider only ultra-central collisions in this paper, it then seems reasonableto assume that linear mapping still holds in ultracentral

Page 6: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

6

small systems. We note that, while deviations from linearresponse can affect the magnitude of vn{2}, they shouldnot strongly affect its multiplicity dependence, and muchof the systematic uncertainties associated with assuminglinear response will cancel in the ratio vin{2}/vn{2}. Theaccuracy of a linear response assumption can in principlebe tested against our full hydrodynamic simulations, but– as with a recalibration of subnucleonic fluctuations atRHIC – will require the dedication of significant compu-tational resources. We leave such an analysis for futurework, when we hope that some of the new collision speciesconsidered here will be approved.

Accordingly, for most of this paper we will assume lin-ear response (20) in order to calculate high statistics re-sults in ultracentral UU collisions at RHIC. Running afull hydrodynamics simulation on an event-by-event ba-sis is not realistic since millions of events are required toperform such an analysis. Then, since we calculate onlyratios such as vin{2}/vn{2}, the coefficient κn cancels outand we can directly use the eccentricities of the initialconditions to make flow predictions in central collisions.

In the flow-only scenario, the final-state two-particledistribution (3) dN1

d2p dy with δ2 = 0. Then the second term

of (5) vanishes, and the two-particle cumulant vn{2} issimply the root-mean-square of the corresponding single-particle flow harmonic:

vn{2} =√〈v2n〉 ≈ |κn|2

√〈ε2n〉. (22)

As emphasized previously, this arises because in hydrody-namics the particle are emitted independently at freeze-out. Correspondingly, the ratio of cumulants in an indi-vidual sub-bin i to the full 0− 1% ultracentral selectioncut is

vin{2}vn{2}

=

√〈ε2n〉i

〈ε2n〉0−1%

(23)

which can be compared with the CGC scaling in (15).

III. RESULTS

In this Section, we will show results for the ratio ofcumulants in sub-bins of ultracentral collisions of vari-ous ion species, comparing the predictions of two-gluoncorrelations from the CGC-only scenario (15) versus theinitial conditions+hydrodynamics scenario (with linearresponse) (23). In Sec. III A we examine the differ-ences for 238U 238U collisions: a symmetric collision ofhighly-deformed ions for which the CGC scaling was firstdemonstrated [54]. In Sec. III B we study d 197Au colli-sions as an example of an asymmetric collision betweena small, deformed ion (deuteron) with a large, round ion(gold). In Sec. III C we study collisions involving an in-termediate but highly-deformed ion: 9Be. We considerboth symmetric 9Be 9Be and asymmetric 9Be 197Au col-lisions. And in Sec. III D we similarly consider the gen-eration of triangular flow v3{2} from collisions involving

STAR data

Trento a=0.55, R=6.86 β2=0.28

CGC

UU 193 GeV, ZDC 0-1%

0.90 0.95 1.00 1.05 1.100.90

0.95

1.00

1.05

1.10

1.15

Mi/<M>

v2i{2}/v

2{2}

FIG. 2: Elliptic flow versus multiplicity for the 0− 1%most central 238U 238U collisions by ZDC at RHIC (see

Appendix A). Note that the axes are scaled by theaverage multiplicity 〈M〉 and elliptic flow vn{2} for theentire 0− 1% centrality class. Here the multiplicity is

determined using the√TATB scaling (the p = 0

configuration of TRENTO (2a)).

an inherently triangular nucleus: 3He. We consider bothsymmetric 3He 3He and asymmetric 3He 197Au collisions.Relevant details about the centrality class selection andWood-Saxon parameters for 238U 238U are contained inAppendices A and B.

A. 238U 238U Collisions

In Ref. [54], 238U 238U collisions were considered,leading to the prediction of opposite correlations ofvn{2} with multiplicity M = dN

dy between the CGC-

mediated two-gluon correlations and the initial condi-tions+hydrodynamics picture. While it is not widelybelieved that CGC mechanisms alone can describe thehadronic correlations in such a large system, it is still in-teresting to examine a highly-deformed ion like uraniumto study the sensitivity of both pictures to deformed nu-clei. This is especially true, given that the relevant datahas already been published by the STAR Collaboration[46] and that 238U 238U is the only collisional system ofdeformed ions with published data to date for this spe-cific measurement.

The results for 238U 238U are shown in Fig. 2, wherewe have subdivided the 0 − 1% centrality class into 20sub-bins i by multiplicity. In contrast to the initial con-dition+hydrodynamic picture, which leads to an anticor-relation of vi2{2} with M i, the CGC picture positivelycorrelates the two. This behavior is qualitatively incon-sistent with the STAR data [46], indicating (unsurpris-ingly) that an initial-state only picture cannot describeultracentral 238U 238U collisions; in contrast, the initialconditions+hydrodynamics picture is able to describe thedata with the indicated model parameters (Appendix B).

It is important to note that the manner in which the

Page 7: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

7

Trento TATB

CGC TATB

nucleons

STAR

UU 200 GeV, ZDC 0-1%

0.8 0.9 1.0 1.1 1.2 1.30.85

0.90

0.95

1.00

1.05

1.10

1.15

1.20

Mi/<M>

v2i{2}/v

2{2}

FIG. 3: Same as Fig. 2 but with linear TATB scaling(2b) for the multiplicity.

multiplicity dependence is calculated is subject to sig-nificant theoretical uncertainty, due in part to the non-perturbative effects associated with hadronization. Toexamine this theoretical uncertainty, we compare Fig. 2,which uses a multiplicity determined by the phenomeno-logical p = 0 configuration of TRENTO (2a) such thatM ∝

∫ √TATB , with Fig. 3, which uses the linear scaling

M ∝∫TATB from the CGC (2b). One notable effect is

the enhancement of multiplicity fluctuations with a TATBscaling, as reflected in the greater range of the horizon-tal axis. The initial conditions+hydrodynamics pictureis largely unaffected here by the change in multiplicityscaling and is consistent with the STAR data in bothFigs. 2 and 3. The CGC picture, on the other hand, seesthe magnitude of its positive slope reduced by switchingto TATB multiplicity scaling.

While the disfavoring of an initial-state-only model fora heavy-ion collision like 238U 238U was largely a foregoneconclusion, the ability to discriminate between qualita-tively different models demonstrated in Fig. 2 serves asa baseline proof of concept which we will next extend tosmaller systems where the expected outcome is not soclear.

B. d 197Au Collisions

Next we move to a smaller, asymmetric system:deuteron-gold (d Au) collisions at

√sNN = 200 GeV.

The deuteron is a deformed system, possessing a dom-inant elliptical geometry; however the proton-neutronseparation can be rather wide, colliding as two well-separated nucleons or even with one missing the goldion entirely. Very central d Au collisions can also corre-spond to a “tip-on” deuteron orientation, with the twonucleons aligned close to the beam axis such that thetransverse profile is very round. The result, shown inthe solid curves in Fig. 4, is that the geometry-driven re-sponse of the initial conditions+hydrodynamics picture isnearly flat across multiplicity, whereas the CGC picture

Trento

CGC

nucleons substructure

dAu 200 GeV, ZDC 0-1%

0.5 1.0 1.5 2.0

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

Mi/<M>

v 2i {2}/v2{2}

FIG. 4: Elliptic flow versus multiplicity for the 0− 1%most central d Au collisions by ZDC at 200 GeV at

RHIC, using the p = 0 multiplicity scaling (2a)proportional to

√TATB . Solid curves indicate nucleon

degrees of freedom; dashed curves include n = 6sub-nucleonic constituents.

still yields elliptic flow which increases with the multi-plicity.

The emerging qualitative picture is that hydrodynamicpicture and the CGC picture couple oppositely to de-formed geometries; however, the quantitative magnitudeof this effect can depend on nonperturbative details of themodels. These include not only the choice of multiplicityschemes like M ∝

∫ √TATB versus M ∝

∫TATB as com-

pared in Figs. 2 and 3, but also the choice of whether toconsider possible nucleon substructure in the event-by-event fluctuations of the nuclear profile functions. Tostudy this latter effect, we use the constituent TRENTOmodel with n = 6 constituents per nucleon with param-eters that were previously tuned to p Pb collisions [67].The substructure effect is reflected in the dashed curvesin Fig. 4. We see a modest ∼ O (5%) enhancement inv2{2} in the initial conditions+hydrodynamics picture,reflecting a small enhancement in the overall ellipticityof the deuteron when substructure is included. In con-trast, turning on substructure leads to a slight flatteningof the multplicity dependence in the CGC picture. Asdiscussed in Appendix C, while the gluon correlationsleading to v2{2} in the CGC picture are not directly af-fected by the ellipticity of the collision geometry, they areaffected by the smoothness or lumpiness of the distribu-tion. Going from a deuteron characterized by nucleondegrees of freedom to one including sub-nucleonic fluc-tuations results in a somewhat lumpier density distribu-tion, which generally flattens the multiplicity dependenceof the CGC effect.

Another substantial effect of including nucleon sub-structure is on the distribution of multiplicity fluctua-tions, as shown in Fig. 5. Nucleon degrees of freedomwith Gaussian width σ = 0.3 fm describe the STAR dataperfectly, while the larger value σ = 0.51 fm slightly un-derpredicts the high-multiplicity tail. This is consistent

Page 8: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

8

Trento TaTb

Trento TaTb

Trento substructure TaTb

Trento substructure TaTb

Trento TaTb σ=0.51 fm

STAR dAu 200 GeV

0 2 4 6 8

10-11

10-9

10-7

10-5

0.001

0.100

Nch/<Nch>

<N

ch>

P(N

ch)

FIG. 5: Multiplicity distributions in TRENTO fordifferent parameter choices, compared to STAR data[85]. For nucleons, unless otherwise noted we use a

default width σ = 0.3 fm.

Trento σ=0.3 fm

Trento σ=0.51 fm

dAu 200 GeV, ZDC 0-1%

0.6 0.8 1.0 1.2 1.4 1.6 1.80.96

0.98

1.00

1.02

1.04

Mi/<M>

v2i{2}/v

2{2}

FIG. 6: Dependence on the nucleon width parameter σin TRENTO for the 0− 1% most central d Au collisionsby ZDC at 200 GeV. Here we use

√TATB multiplicity

scaling.

with the conclusions of Ref. [58] that a slight preferencemay exist for the smaller value σ = 0.3 fm in smallsystems. As seen in Fig. 6, this change in the nucleonwidth parameter also leads to a small O (2%) change inthe multiplicity dependence of v2{2} in the initial con-ditions+hydrodynamics picture. The multiplicity fluc-tuations including n = 6 sub-nucleonic constituents for√TATB scaling also capture the trends of the experimen-

tal multiplicity distribution in Fig. 5 reasonably well, butoverpredict the high-multiplicity tail. This is not surpris-ing, because the particular substructure parameters usedhere were tuned for p 208Pb collisions at the LHC.

The multiplicity distribution for nucleons with linearTATB scaling fails to describe the experimental data inFig. 5, suggesting that a significant retuning of the modelparameters may be needed for this new scaling choice. Itis also quite surprising that the multiplicity distributionfor linear TATB scaling including nucleon substructure

Trento k=1.6

Trento k=2

dAu 200 GeV, ZDC 0-1%

0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.40.90

0.95

1.00

1.05

1.10

Mi/<M>

v2i{2}/v

2{2}

FIG. 7: Dependence on the multiplicity fluctuationparameter k in TRENTO for the 0− 1% most central

d Au collisions by ZDC at 200 GeV. Here we use√TATB multiplicity scaling. (See also Fig. 5).

nucleons Ta Tb

substructure Ta Tb

nucleons TaTb

CGC dAu 200 GeV, ZDC 0-1%

0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.5

1.0

1.5

2.0

Mi/<M>

v2i{2}/v

2{2}

FIG. 8: Sensitivity of the CGC picture to TATB vs.√TATB multiplicity scaling and to nucleonic

substructure in d 197Au collisions at RHIC.

agrees so well with the data, given that this is a newconfiguration of TRENTO and no retuning has been per-formed. We suspect this is simply an accident of thevarious competing changes and consider a thorough ex-ploration of the parameter space to be necessary to fullyunderstand the content of the model. We leave such ananalysis for future work, and we will primarily focus onthe phenomenologically successful

√TATB scaling from

here on, unless otherwise noted. We do note, however,that the most important quantity controlling the multi-plicity fluctuations is the TRENTO parameter k, whichwill need to be re-fit for d 197Au collisions at RHIC.Thankfully, however, the flow harmonics of interest tous here seem to be relatively insensitive to corrections tothe fluctuation parameter k as shown in Fig. 7. Thus thedeviations in the multiplicity distribution seen in Fig. 5may not strongly affect the resulting flow harmonics – atleast for nucleon degrees of freedom.

The resulting sensitivity to the choice of multiplicity

Page 9: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

9

n=2

n=3

dAu 200 GeV

CGC only TaTb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

M/<M>0-1%

vn{2}/vn0-

1%{2}

n=2

n=3

dAu 200 GeV

CGC only Ta T

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

M/<M>0-1%

vn{2}/vn0-

1%{2}

FIG. 9: Elliptic and triangular flow versus multiplicity across the entire centrality range for d Au collisions at RHICfor the CGC mechanism (15). Here we compare the linear TATB multiplicity scaling (left panel) with

√TATB

scaling (right panel). Note that the axes are scaled by the flow harmonics and multiplicity in the 0− 1% mostcentral collisions, although results are shown for all centralities.

Trento

σ=0.3 fm 0-1%

σ=0.3 fm 0-10%

substructure 0-1%

substructure 0-10%

dAu 200 GeV, ZDC 0-10%

0.5 1.0 1.5 2.00.90

0.95

1.00

1.05

1.10

1.15

Mi/<M>

v2i{2}/v

2{2}

FIG. 10: Comparison of elliptic flow in the central0− 10% and ultracentral 0− 1% bins for d 197Au

collisions at RHIC in the hydrodynamics picture, forboth nucleon and sub-nucleonic degrees of freedom.

scaling and to the presence of nucleonic substructure isshown in Fig. 8 for the CGC picture. As seen previouslyin Fig. 4, there is a small flattening of the multiplicity de-pendence caused by the presence of nucleonic substruc-ture for

√TATB multiplicity scaling. Far more signifi-

cant is the dependence on the mechanism of multiplicitygeneration. The smooth, deformed profile achieved withnucleon degrees of freedom and soft

√TATB multiplic-

ity scaling yields the greatest ∼ O (50%) impact on themultiplicity dependence, and sharpening the multiplic-ity scaling to TATB flattens the multiplicity dependencetremendously. Both of these effects can be interpretedas due to increasing the lumpiness of the density profiles,leading to a flattening of the multiplicity dependence. Wehave also checked that this trend continues for the caseof both linear TATB scaling and nucleonic substructure.

It is also interesting to observe the impact of the de-

formed deuteron geometry by comparing the 0−1% ultra-central collisions seen in Fig. 4 against the same quantityover a wider centrality range. A comparison of the ul-tracentral 0−1% bin versus wider centrality selections isshown in Figs. 9 and 10 for the CGC and initial condi-tions+hydrodynamics pictures, respectively.

For the CGC picture, we examine the full (mini-mum bias) centrality range in Fig. 9 for both TATBand

√TATB scaling. The multiplicity dependence seen

in the left panel of Fig. 9 for TATB scaling appears toreproduce beautifully the predictions v2{2} ∝ M0 andv3{2} ∝ M1/2 from Ref. [84]. On the other hand, thecorresponding multiplicity dependence that would arisefrom the p = 0 scaling

√TATB shown in the right panel

of Fig. 9 appears to have been significantly damped.We study the multiplicity dependence of various colli-sion systems and the impact of nuclear deformation inAppendix C.

For the initial conditions+hydrodynamics picture inFig. 10, the slopes and trends are practically identicalfor the two centrality cuts, both for nucleon degrees offreedom and for nucleonic substructure. But in the caseof sub-nucleonic fluctuations, the width of the multiplic-ity fluctuations in the 0 − 10% bin has increased, lead-ing to an increase in the overall magnitude of the sub-structure effect from O (5%) in the ultracentral bin toO (10%) in the central bin. This suggests that selectioncuts with wider bins that make it possible to capturewider multiplicity fluctuations with high statistics are amore promising approach to distinguishing the effects ofsub-nucleonic fluctuations.

C. 9Be 197Au and 9Be 9Be Collisions

A useful system intermediate to d 197Au and 238U 238Ucollisions is 9Be 197Au, since it collides the small, highly-

Page 10: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

10

Trento

CGC

nucleons substructure

Be Be 200 GeV, ZDC 0-1%

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

0.6

0.8

1.0

1.2

1.4

Mi/<M>

v 2i {2}/v2{2}

FIG. 11: Elliptic flow versus multiplicity for the 0− 1%most central 9Be 9Be collisions by ZDC at 200 GeV atRHIC, using

√TATB multiplicity scaling. Solid curves

indicate nucleon degrees of freedom; dashed curvesinclude n = 6 sub-nucleonic constituents.

Be Be

Be Au

nucleons substructure

CGC 0-1% 200 GeV

0.6 0.8 1.0 1.2 1.4 1.6 1.8

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Mi/<M>

v2i{2}/v

2{2}

FIG. 12: Sensitivity of the CGC picture to nucleonicsubstructure in ultracentral collisions of both 9Be 9Be

and 9Be 197Au at RHIC. Here we use√TATB

multiplicity scaling.

deformed beryllium ion with the large spherical gold ion.This highly asymmetric collision is to be contrasted witha symmetric collision of small deformed ions with eachother, such as 9Be 9Be. Collisions involving 9Be werepreviously proposed in Ref. [86] as part of a programto study collisions of light polarized nuclei, although thepolarization studies are not our main focus in this paper.For 9Be we use the parameters R = 1.4 fm, a = 0.75 fm,β2 = 0.64, β4 = 0.27 from Ref. [87]; in comparison, thequadrupole deformation β2 of 238U is only about 0.28.

An overview of the elliptic flow results for sym-metric 9Be 9Be collisions in both the initial condi-tions+hydrodynamics and CGC pictures is shown inFig. 11. More detailed results comparing 9Be 9Be and9Be 197Au collisions are shown in Fig. 12 for the CGCpicture and in Fig. 13 for the hydrodynamics picture.The overall qualitative picture seen for 9Be 9Be colli-

Be Be

Be Au

nucleons substructure

Trento 0-1% 200 GeV

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.80.90

0.95

1.00

1.05

1.10

Mi/<M>

v 2i {2}/v2{2}

FIG. 13: Sensitivity of the hydrodynamics picture tonucleonic substructure in ultracentral collisions of both9Be 9Be and 9Be 197Au at RHIC. Here we use

√TATB

multiplicity scaling.

0-1%

0-10%

nucleons substructure

Trento Be Be 200 GeV, ZDC

0.5 1.0 1.5 2.00.90

0.95

1.00

1.05

1.10

Mi/<M>

v 2i {2}/v2{2}

FIG. 14: Comparison of elliptic flow in the central0− 10% and ultracentral 0− 1% bins for 9Be 9Be

collisions at RHIC in the hydrodynamics picture, forboth nucleon and sub-nucleonic degrees of freedom.

Here we use√TATB multiplicity scaling.

sions in Fig. 11 is quite consistent with the results shownin Fig. 4 for d 197Au. For nucleon degrees of freedom,the hydrodynamics picture produces a mostly constantv2{2}, with the addition of nucleonic substructure pro-ducing an O (5− 10%) enhancement at low multiplicityand a small negative slope. In contrast, the CGC mecha-nism increases almost linearly with multiplicity, with theaddition of substructure leading to a slight decrease inthe slope of the multiplicity dependence.

The elliptic flow vs. multiplicity results for the CGCpicture with

√TATB multiplicity scaling are shown in

more detail in Fig. 12, including both 9Be 9Be and9Be 197Au collisions as well as for σ = 0.3 fm nucleonsand n = 6 nucleonic substructure. The results seen hereare consistent with the ones shown in Figs. 4 and 8 ford 197Au: the effect of nucleonic substructure is to slightlyflatten the multiplicity distribution due to the increas-

Page 11: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

11

Trento

CGC

nucleons substructure

Ta Tb dAu 200 GeV, ZDC 0-1%

0.6 0.8 1.0 1.2 1.4 1.6 1.80.0

0.5

1.0

1.5

2.0

Mi/<M>

v 3i {2}/v3{2}

FIG. 15: Triangular flow v3{2} in ultracentral dAucollisions at RHIC for the hydrodynamics and CGC

pictures, with and without n = 6 sub-nucleonicconstituents, using

√TATB multiplicity scaling.

ing lumpiness of the initial density profile. We also notethat the slope of the CGC effect is slightly greater inthe collisions of 9Be than it was for d 197Au: the sameO (60%) effect on v2{2} is achieved for 9Be over a some-what smaller range of multiplicity fluctuations than ind 197Au. Presumably, this is due to the large deforma-tion of 9Be .

Similar results for the initial condi-tions+hydrodynamics picture are shown in moredetail in Fig. 13. As with d 197Au collisions shownin Figs. 4 and 10, for nucleon degrees of freedom themultiplicity dependence is quite flat, both for 9Be 9Beand 9Be 197Au collisions. Turning on nucleonic sub-structure introduces a discernable negative slope to themultiplicity dependence, although the effects are hardto distinguish in the case of 9Be 197Au collisions. But,while the slopes appear to be similar in both 9Be 9Beand 9Be 197Au collisions with sub-nucleonic fluctuations,the wider multiplicitiy fluctuations seen in the smaller9Be 9Be system results in a more prominent effect.

We can further enhance the width of the multiplicityfluctuations in 9Be 9Be collisions by relaxing our ultra-central 0− 1% cut to a 0− 10% centrality cut, as shownin Fig. 14. As expected, this does increase the widthof the multiplicity fluctuations, increasing the magni-tude of the signal of nucleonic substructure. However,the 0 − 10% cut also begins to capture a nontrivial el-lipticity for nucleons; this generates a slope for the nu-cleon case and competes with the enhanced multiplicityfluctuations, making the ability to resolve nucleonic sub-structure nontrivial. It would seem that the ultracentral0− 1% events in 9Be 9Be collisions really select on roundcollision geometries, whereas a less extreme centrality cutreflects the elliptical shape of 9Be.

The emerging picture regarding nucleonic substructureis that, while the impact of substructure is fairly small, itcan be more easily distinguished by enhancing the role of

Trento

CGC

nucleons substructure

3He 3He 200 GeV, ZDC 0-1%

0.0 0.5 1.0 1.5 2.0 2.5

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Mi/<M>

v 3i {2}/v3{2}

FIG. 16: Triangular flow v3{2} in ultracentral 3He 3Hecollision at RHIC for the hydrodynamics and CGC

pictures, with and without n = 6 sub-nucleonicconstituents, using

√TATB multiplicity scaling.

multiplicity fluctuations. This can be achieved by consid-ering ultracentral collisions of small symmetric systemssuch as 9Be 9Be rather than 9Be 197Au. Relaxing the ul-tracentral centrality cuts can also enhance the multiplic-ity fluctuations associated with nucleonic substructure.This leads to a clear enhancement of the discriminatingpower between nucleon and sub-nucleonic degrees of free-dom in Fig. 10 for d 197Au collisions, but for larger de-formed ions like 9Be this multiplicity enhancement com-petes against the innate ellipticity of the nucleon distri-bution. We expand upon this idea in greater detail for3He Au versus 3He 3He collisions in the following Section.

D. Triangular Flow and 3He 3He and 3He 197AuCollisions

Unlike the deformed ions that discussed previously,3He has an inherent triangular shape rather than ellip-tical shape. Thus it is interesting to study the responseof the triangular flow v3{2} in collisions involving 3He,as compared to the other elliptically-deformed systems.As a baseline to compare against the triangular shapeof 3He, in Fig. 15 we plot the triangular flow v3{2} ford 197Au collisions for

√TATB scaling, with and without

nucleonic substructure. As expected from the lack ofa triangular geometry with nucleon degrees of freedom,the initial conditions+hydrodynamics response is almostcompletely flat. Contrary to what might be expected,introducing nucleonic substructure does not appear togenerate an inherent triangularity ε3 from its fluctuatingsubstructure – at least not contributing to the ultracen-tral 0 − 1% centrality bin. The CGC picture, on theother hand, generates a triangular flow directly withoutneeding an underlying triangular geometry ε3, so it pro-duces a positively-sloped curve both with and withoutsubstructure. And, as with previous collision systems,turning on nucleonic substructure somewhat flattens the

Page 12: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

12

slope of the CGC mechanism.

In Fig. 16 we perform the same comparison for thehydrodynamics picture in 3He 3He collisions. Becauseof the innate ε3 of the triangular 3He ion, the initialconditions+hydrodynamics picture possesses a negativeslope even for nucleon degrees of freedom. Turning onnucleonic substructure, however, flattens this behavior,smoothing out the multiplicity dependence associatedwith the triangularity of 3He. The innate ε3 at the levelof nucleons leads to an ∼ O (10%) effect, whereas turningon substructure reduces this to an ∼ O (5%) effect. Inter-estingly, this substructure effect moves the triangularityof 3He in an opposite direction from the ellipticity seenin 9Be and the deuteron. Meanwhile, the CGC pictureshows a steeper slope for v3{2} than for v2{2} in accor-dance with the dependence on a higher power T 3

AT3B of

the nuclear profiles in Eq. (13). As seen in the other col-lision systems previously, turning on nucleonic substruc-ture dilutes the effect and flattens the slope somewhat.

Previously, we saw that the substructure effect in9Be 9Be versus 9Be 197Au collisions in Fig. 13 led to com-parable slopes, with the distinguishing feature being thewider multiplicity fluctuations present in the smaller col-lision system. This difference enhanced an effect whichmay have been unidentifiable in 9Be 197Au collisions tobe an O (5%) effect in 9Be 9Be collisions. Accordingly,we expect that symmetric collisions of small deformedions will be more discriminating (within a hydrodynam-ics picture) to the presence or absence of sub-nucleonicfluctuations. We can further test this interpretation byextending it to the v3{2} sector and comparing 3He 3Heversus 3He 197Au collisions, which are shown in Fig. 17.As anticipated, we find in Fig. 17 that 3He 3He collisionsare more sensitive than 3He 197Au collisions to the pres-ence of sub-nucleonic fluctuations. The reason for thisdiscriminating power for v3{2} in 3He is opposite to thereason for its discriminating power for v2{2} in d 197Auand 9Be – a suppression due to substructure, rather thanan enhancement – but it is a discriminator nonetheless.

Similarly, we saw in Fig. 10 that the increase in multi-plicity fluctuations in going from 0−1% ultracentral col-lisions to 0−10% central collisions for d 197Au magnifiedan O (5%) substructure effect to an O (10%) effect. How-ever, we also saw in Fig. 14 that for 9Be 9Be collisions thisenhancement of the substructure signal due to increasedmultiplicity was partially offset by an increased ellipticflow for nucleon degrees of freedom only. For 3He 3Hecollisions binned with 0 − 1% versus 0 − 10% centralitycuts shown in Fig. 18, the story is similar. Keeping inmind that nucleonic substructure in this case leads toa suppression of v3{2}, we see that for 0 − 1% ultra-central collisions, the separation between nucleonic andsub-nucleonic degrees of freedom is around O (5%). Re-laxing the centrality cuts to 0− 10% central events doessomewhat increase the range of multiplicity fluctuationsfor nucleonic substructure, but it also leads to a simul-taneous reduction in the v3{2} for nucleons. This againreduces the discriminating power between nucleon and

3He 3He3He Au

nucleons substructure

200 GeV, ZDC 0-1%0.0 0.5 1.0 1.5 2.0 2.5

0.90

0.95

1.00

1.05

1.10

1.15

Mi/<M>

v 3i {2}/v3{2}

FIG. 17: Comparison of triangular flow v3{2} inultracentral 3He 3He versus 3He 197Au collisions collision

at RHIC for the hydrodynamics picture, with andwithout n = 6 sub-nucleonic constituents, using

√TATB

multiplicity scaling.

0-1%

0-10%

nucleons substructure

Trento 200 GeV, ZDC0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.90

0.95

1.00

1.05

1.10

1.15

Mi/<M>

v 3i {2}/v3{2}

FIG. 18: Comparison of elliptic flow in the central0− 10% and ultracentral 0− 1% bins for 3He 3He

collisions at RHIC in the hydrodynamics picture, forboth nucleon and sub-nucleonic degrees of freedom.

Here we use√TATB multiplicity scaling.

sub-nucleonic degrees of freedom in the looser 0 − 10%centrality cuts.

IV. CONCLUSIONS

In this paper we studied ultracentral collisions of de-formed ions, specifically studying the scaling of ellipticflow v2{2} and triangular flow v3{2} versus multiplic-ity. This work was motivated in part by the tensionbetween the prediction of Ref. [54] that gluon correla-tions in the CGC should exhibit a positive correlationof v2{2} with multiplicity in central 238U 238U collisions,and the anticorrelation seen in STAR data. We thereforecompared these opposing correlations of elliptic flow ver-sus multiplicity across a multitude of smaller ultracentral

Page 13: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

13

collisional systems where initial-state CGC effects mayplay a more significant role: 238U 238U, d 197Au, 9Be 9Be,9Be 197Au, 3He 3He, and 3He 197Au. Across all systems,we find that this qualitative distinction holds, with theCGC picture generally leading to a positive correlationbetween elliptic flow and multiplicity and the initial con-ditions+hydrodynamics picture leading to an anticorre-lation. A comparable picture also holds for the triangularflow arising in d 197Au, 3He 3He, and 3He 197Au collisions.

We also study the dependence of these results on anumber of parameters and model choices. These includecomputing the initial conditions from nucleon versus nu-cleonic structure, computing the multiplicity from a givenevent proportional to

√TATB versus TATB , and varying

the nucleon width. The size of the nucleon width makesa small O (2%) difference in the slope for the initial con-ditions+hydrodynamics picture (see Fig. 6), with varia-tions in the multiplicity fluctuation parameter k havinga smaller effect shown in Fig. 7, within error bars.

The inclusion of n = 6 sub-nucleonic constituentswithin TRENTO v2.0 leads to a systematic effect whichis opposite in its enhancement of ellipticity in thedeuteron and in 9Be and its suppression of triangular-ity in 3He. These effects are difficult to distinguish inasymmetric collisions such as 9Be 197Au and 3He 197Au,but can be enhanced by increasing the multiplicity fluc-tuations using small symmetric collisions such as 9Be 9Beor 3He 3He. Using a looser 0−10% centrality cut does sig-nificantly enhance the substructure signal in d 197Au col-lisions, although the discriminating power in this widerbin is reduced for somewhat larger ions. Taken together,these results confirm our interpretation of central andultracentral collisions involving small deformed ions asbeing sensitive to sub-nucleonic fluctuations. This holdstrue both in for elliptic and triangular flow, across a rangeof collision systems.

The CGC picture generally obtains positive correla-tions between the flow harmonics and multiplicity in de-formed systems, with the slope of the correlation be-ing the greatest for small asymmetric collisions with nu-cleon degrees of freedom and

√TATB multiplicity scaling,

which produces fairly smooth density profiles. Increasingthe lumpiness of the density profile, either by turning onnucleon substructure or changing to the sharper TATBmultiplicity scaling, tends to flatten out the distribution.

In this paper we have focused exclusively on measure-ments for various ions at RHIC, based on the successfuluse of the STAR ZDC to select on tip-on-tip versus side-on-side collision geometries of 238U. One further possibil-ity for future study using the unique capabilities of RHICwould include the interplay between polarization and de-formation for light nuclei like 9Be; as is well-known in thespin community, polarizing a hadron can induce a corre-sponding spatial deformation of its internal constituents[88]. The analysis performed here can also in principlebe extended to the LHC, if a comparable selection onultracentral events by spectator binning in the ZDC ispossible; we leave this possibility for future work.

ACKNOWLEDGMENTS

The authors wish to thank K. Dusling, T. Lappi, M.Luzum, M. Mace, S. Moreland, J. Nagle, B. Schenke, S.Schlichting, P. Tribedy, R. Venugopalan, and W. Zajcfor useful discussions. The authors acknowledge sup-port from the US-DOE Nuclear Science Grant No. de-sc0019175, the Alfred P. Sloan Foundation, and the Officeof Advanced Research Computing (OARC) at Rutgers,The State University of New Jersey for providing accessto the Amarel cluster and associated research computingresources that have contributed to the results reportedhere. D.E.W. acknowledges support from the Zucker-man STEM Leadership Program. M.D.S. and D.E.W.wish to thank the Department of Energy’s Institute forNuclear Theory at the University of Washington for itshospitality while facilitating a portion of this work.

Appendix A: Centrality binning method

Here we explore which centrality binning method isclosest to experimental data. The simplest way to un-derstand what STAR’s ZDC measures is it essentially thenumber of nucleons that don’t participate in the collision.This would be roughly equivalent to NZDC = 2A−Npart

and for ultracentral collisions NZDC → 0. From thisstandpoint, we anticipate that binning by Npart wouldthen be the most effective way to select on the 0 − 1%most central events. However, after that initial selec-tion is made the events are binned by multiplicity. Thus,we conclude that one first must bin by Npart to obtainthe 0 − 1% centrality class but then one must bin bymultiplicity, which from an initial conditions standpointcorresponds to total entropy. (Note that here we useM ∼ S0

4 .)

STAR data

Trento Npart bins

Trento S0 bins

UU 193 GeV, ZDC 0-1%

0.85 0.90 0.95 1.00 1.05 1.10 1.150.90

0.95

1.00

1.05

1.10

Mi/<M>

v2i{2}/v

2{2}

FIG. 19: UU collision events in the 0− 1% most centralcollisions (selecting on Npart vs. entropy S0) where theaverage multiplicity is 〈M〉 and the flow is vn{2} for theentire centrality class. 20 subdivisions i in multiplicityare made and their corresponding vin{2} is calculated.

Page 14: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

14

In order to demonstrate this we compare binning the0−1% by either Npart or S0 as shown in Fig. 19. One no-tices immediately that binning by Npart provides a muchwider range in multiplicity compared to binning by S0

such that Npart binning is more consistent with exper-imental data. Certainly, experimental data has even awider range in multiplicity. This may be simply due tostatistics (for this result we ran 3 million events overall centralities, which is then 30,000 in the 0 − 1% bin)but also may be attributed to differences between the-ory and the detector (e.g the detector can only measurecharged particles whereas the multiplicity here includesall hadrons).

In Fig. 19 we only show the results for TRENTO butwe note that the CGC case has equivalent results (justwith the inverse slope).

Appendix B: Deformation Parameters for for 238U

Another natural question is if these symmetric ultra-central collisions can be used to constrain nuclear defor-mations. In order to investigate this, we use two differentparameters sets for 238U which have been used in recentpapers [50, 56]. We do note, however, that in this paperwe use a = 0.55 rather than a = 0.6 for the configurationin Ref. [56] due to comparisons to data across all centrali-ties with TRENTO+v-USPhydro calculations, which willbe shown in an upcoming paper.

In Fig. 20 we compare the two different parameter setsfor 238U and find that there is a slight preference fora = 0.55, R = 6.86, and β2 = 0.28 in these comparisons,although we note that this should be taken in conjunc-tion with other flow observable comparisons, which willbe shown in future work. Overall, we conclude that thedeformation parameters play a small role in this observ-able but other flow observables may be more useful forconstraining Wood-Saxon parameters. For the CGC pic-ture we see no difference whatsoever in Fig. 20 when wevary the Wood-Saxon parameters, since the CGC calcu-lation is sensitive only to the multiplicity and not to thegeometry itself.

Appendix C: Multiplicity Dependence in the CGC

In this Appendix we examine the simple theoreticalexpressions for the multiplicity dependence of the flowharmonics in the CGC picture and the role of nucleardeformations and gradients in modifying them.

The flow harmonics of two-gluon correlations arising inthe CGC are expressed in Eq. (13), which we can rewrite

STAR data

Trento a=0.55, R=6.86 β2=0.28

Trento a=0.5, R=6.61 β2=0.236

UU 193 GeV, ZDC 0-1%

0.90 0.95 1.00 1.05 1.100.90

0.95

1.00

1.05

1.10

1.15

Mi/<M>

v2i{2}/v

2{2}

STAR data

Trento a=0.55, R=6.86 β2=0.28

Trento a=0.5, R=6.61 β2=0.236

UU 193 GeV, ZDC 0-1%

0.90 0.95 1.00 1.05 1.100.90

0.95

1.00

1.05

1.10

1.15

Mi/<M>

v2i{2}/v

2{2}

FIG. 20: 238U 238U collision events in the 0− 1% mostcentral collisions (selecting on Npart) where the averagemultiplicity is 〈M〉 and the flow is vn{2} for the entire

centrality class. 20 subdivisions i in multiplicity aremade and their corresponding vin{2} is calculated.

Comparisons between the initialconditions+hydrodynamics picture (top panel) and

CGC picture (bottom panel) are shown.

as

v2{2} ∼

√√√√∫ d2x⊥ T 2A(~x⊥)T 2

B(~x⊥)[∫d2x⊥

dNd2x

]2 (C1a)

v3{2} ∼

√√√√∫ d2x⊥ T 3A(~x⊥)T 3

B(~x⊥)[∫d2x⊥

dNd2x

]2 , (C1b)

where we have omitted various proportionality factors tofocus on the multiplicity dependence. The translationfrom a dependence on TA, TB to a multiplicity depen-dence varies based on how the multiplicity is calculated.The multiplicity dependence one naturally obtains in theCGC uses a linear TATB scaling (2b), but the most phe-nomenologically successful multiplicity dependence im-plemented in TRENTO uses a softer

√TATB scaling (2a).

Depending on which framework is used, therefore, wehave one of the following expressions for the multiplicity

Page 15: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

15

density arising from the collision profile:

dN

d2x∼ TA(~x⊥)TB(~x⊥), (C2a)

dN

d2x∼√TA(~x⊥)TB(~x⊥). (C2b)

In terms of multiplicity alone, this then gives for the flowharmonics

v2{2}TATB∼

√√√√∫ d2x⊥(dNd2x

)2[∫d2x⊥

dNd2x

]2 ∼ const (C3a)

v3{2}TATB∼

√√√√∫ d2x⊥(dNd2x

)3[∫d2x⊥

dNd2x

]2 ∼√dN

d2x(C3b)

v2{2}√TATB∼

√√√√∫ d2x⊥(dNd2x

)4[∫d2x⊥

dNd2x

]2 ∼ dN

d2x(C3c)

v3{2}√TATB∼

√√√√∫ d2x⊥(dNd2x

)6[∫d2x⊥

dNd2x

]2 ∼(dN

d2x

)2

,

(C3d)

where in each case the last scaling on the right-hand sidefollows if the multiplicity dependence is roughly indepen-dent of the spatial geometry, dN

d2x ≈ const. This approxi-mate scaling can be satisfied if the collision region is nottoo large compared to the spatial gradients of the multi-plicity density in a given framework.

This analysis is essentially the basis of the back-of-theenvelope estimate by Mace et al. [84], concluding thatv2{2} ∝ (Nch)0 and v3{2} ∝ (Nch)1/2. In Ref. [84], theauthors attribute multiplicity fluctuations to the fluctu-ations of the dilute projectile in (semi)dilute/dense colli-sions such as d Au. Thus, although their calculation dif-fers from the (semi)dilute / (semi)dilute calculation pre-sented here, the two agree on the multiplicity of vn{2} de-pendence through the dominance of the (semi)dilute pro-jectile. The back-of-the-envelope estimates in Ref. [84] fo-cus on the role of color-charge fluctuations in generatingmultiplicity fluctuations, without explicitly factoring inthe geometry dependence of these fluctuations (althoughthey do include event by event geometry fluctuations intheir calculation). As seen clearly in Eqs. (C3), this es-timate corresponds to a gradient expansion in which themultiplicity density dN

d2x is slowly varying.To best realize the approximation of a slowly vary-

ing geometry inherent in the scalings (C3), we con-sider a highly idealized approximation to p Pb collisions:a “cylindrical” proton of constant density TA(~x⊥) ∝θ(0.5 fm−|~x⊥|) which strikes a smooth (optical Glauber)Woods-Saxon profile. We sample random impact pa-rameters of the proton on the target Pb ion, leading tothe results shown in Fig. 21. In this scenario of arti-ficially slow geometry dependence, we see that the scal-ings (C3) are realized beautifully. For linear TATB multi-plicity dependence, v2{2} is roughly constant, and v3{2}

has a downward curvature consistent with a square-rootdependence, as anticipated by Ref. [84]. Likewise, theanalogous scaling behavior for the

√TATB multiplicity

dependence is also clearly satisfied: a linear growth ofv2{2}, and an upward curvature for v3{2} consistent witha quadratic multiplicity dependence. It is interesting tonote, however, that for TATB scaling in peripheral col-lisions where the density gradients are not negligible, asmall deviation from v2{2} ∼ const is seen, giving a slightnegative slope to v2{2}.

Relaxing these restrictive assumptions about smoothcollision geometry in p 208Pb leads to a smearing of thescalings (C3), but leaves the overall trends intact. InFig. 22 the constant-density proton is randomly incidenton a nuclear profile for the Pb ion which has been gener-ated through Monte Carlo sampling of the nucleon posi-tions. This effect leads to a slight smearing of the scalings(C3) in most cases, but causes a pronounced deviationof v2{2} for TATB scaling in peripheral collisions. Forthe most part, though, because the collision region setby the constant-density proton, the gradient-based devi-ations from the scalings (C3) are small.

These deviations become more severe, however, whenwe also allow the proton to have a nontrivial Gaussianprofile of its own, as shown in Fig. 23. Now the structureof the proton samples a wider region because of the tailof its Gaussian distribution, and over this larger regionthe gradient effects become more significant. This leadsnot only to a broadening of the underlying scalings (C3),but also to a much wider region for which v2{2} for TATBscaling has a negative slope.

In Fig. 24 we perform a simple simulation of208Pb 208Pb collisions which uses Monte Carlo samplingof both nuclei and determines the participant geometryby hard-sphere collision detection of the nucleons. Thisis still at a level which is far simpler than TRENTO,which uses a more sophisticated scheme for collision de-tection and includes additional features such as event-by-event multiplicity fluctuations for each participant nu-cleon. Here the deviations from the expected scalings(C3) are maximized due to the large scale of the inter-action region compared to the gradients of the collisionprofile, which is itself a complicated combination of thetwo nuclear densities. Qualitatively, the large densitygradients lead to a significant flattening of the observ-ables, including v3{2} for TATB scaling and both v2{2}and v3{2} for

√TATB scaling. It is notable, however, that

v2{2} for TATB scaling now exhibits a negative slope overalmost the entire multiplicity range due to the gradients.We thus conclude that a large collision region over a verylumpy collision profile generally results in a flattening ofthe flow harmonics in a CGC picture, which in the ex-treme can completely damp out the baseline scalings inEqs. (C3).

The multiplicity scalings seen here for smooth realiza-tions of p 208Pb collisions, and the deviations from themwith more realistic collision geometries and for simple208Pb 208Pb collisions, show the baseline behavior one

Page 16: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

16

n=2 ∼const

n=3 ∼ MiTATB

p Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

n=2 ∼Mi

n=3 ∼Mi2

TATB

p Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

FIG. 21: Elliptic flow (n = 2, black) and triangular flow (n = 3, red) versus multiplicity across the entire centralityrange for an idealized model of p Pb collisions. As described in the text, a proton of uniform density randomlystrikes a smooth optical Glauber profile of the nucleus. We compare the linear TATB (left panel) versus

√TATB

(right panel) multiplicity scalings. Note that the axes are scaled by the most central 0− 10% bin in this case due tolower statistics.

n=2 ∼const

n=3 ∼ Mi

TATB

p Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.5

1.0

1.5

2.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

n=2 ∼Mi

n=3 ∼Mi2

TATB

p Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

FIG. 22: Same as Fig. 21, but for a constant-density proton on a lumpy nuclear profile generated by Monte Carlosampling of nucleon positions in the Pb ion.

can expect from the CGC correlations in the absence ofnuclear deformations. In contrast, for the same observ-ables a different scaling was shown in Ref. [54] for colli-sions of deformed ions such as 238U 238U. For a smoothoptical Glauber profile with ellipsoidal geometry

ρ(~r) = ρ0 exp

[−x

2 + y2

R2− λ2 z

2

R2

](C4)

the ratio of tip-on-tip to side-on-side collision geometriesyields

[v2{2}]tip[v2{2}]side

=1

λ≈ 1.26 (C5)

where TATB scaling has been assumed and λ ≈ 0.79 isreasonable for parameterizations of 238U.

Thus the multiplicity scalings (C3) arising from asmooth round geometry seen in Fig. 21 are modified bynuclear deformation, leading instead to v2{2} which in-creases with multiplicity instead of remaining constant.This prediction is realized in Fig. 25 for smooth opticalGlauber collisions of 238U 238U. Here we see that, be-cause of the deformation, all of the curves now show anincreasing and roughly linear trend with the multiplicity.

Most importantly, v2{2} for TATB scaling has deviatedfrom its constancy to show a positive slope for the firsttime due to the deformation.

Finally, in Fig. 26 we similarly show the modificationof the scalings (C3) due to nuclear deformation for theasymmetric collision of a small deformed ion on a largerround one. Here we use smooth optical Glauber profilesfor 9Be 197Au collisions as a way to explore an intermedi-ate collision system. Most of the trends seen for smooth238U 238U collisions are already present in Fig. 26 for9Be 197Au: flow harmonics which increase linearly withmultiplicity in most cases. The exception is that we stillsee a small negative slope of v2{2} with TATB scaling,similar to what was seen for the collision of undeformedsystems previously. We thus suspect that this is becausewe are not yet seeing the full impact of the deformationin a truly ultracentral regime.

This analysis demonstrates that the multiplicity scal-

ing v2{2} ∝ N0ch and v3{2} ∝ N

1/2ch for multiplicities

computed proportional to TATB in the CGC are well-realized in p 208Pb collisions, even when a lumpy nuclearprofile on the scale of the proton is present. These scal-ings are violated, however, when the collision region en-

Page 17: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

17

n=2 ∼const

n=3 ∼ Mi

TATB

p Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

1.2

Mi/M0-10%

vni{2}/vn0-10

%{2}

n=2 ∼Mi

n=3 ∼Mi2

TATB

p Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

FIG. 23: Same as Fig. 21, but for a proton with a Gaussian profile on a lumpy nuclear profile generated by MonteCarlo sampling of nucleon positions in the Pb ion.

n=2 ∼const

n=3 ∼ Mi

TATB Pb Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.5

1.0

1.5

2.0

2.5

3.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

n=2 ∼Mi

n=3 ∼Mi2

TATB

Pb Pb

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

1.2

Mi/M0-10%

vni{2}/vn0-10

%{2}

FIG. 24: A simpler model of 208Pb 208Pb collisions than used in TRENTO illustrating the generic flattening of themultiplicity dependence seen in collisions of large, round ions.

n=2 ∼Mi

n=3 ∼Mi2

TATB

U U

0.85 0.90 0.95 1.00

0.85

0.90

0.95

1.00

Mi/M0-10%

vni{2}/vn0-10

%{2}

n=2 ∼Mi

n=3 ∼Mi2

TATB

U U

0.93 0.94 0.95 0.96 0.97 0.98 0.99 1.000.75

0.80

0.85

0.90

0.95

1.00

Mi/M0-10%

vni{2}/vn0-10

%{2}

FIG. 25: Smooth optical Glauber 238U 238U collisions realizing the different scaling (C5) for deformed nuclei.

n=2 ∼Mi

n=3 ∼Mi2

TATB

Be Au

0.6 0.7 0.8 0.9 1.00.80

0.85

0.90

0.95

1.00

1.05

1.10

Mi/M0-10%

vni{2}/vn0-10

%{2}

n=2 ∼Mi

n=3 ∼Mi2

TATB

Be Au

0.80 0.85 0.90 0.95 1.000.5

0.6

0.7

0.8

0.9

1.0

Mi/M0-10%

vni{2}/vn0-10

%{2}

FIG. 26: Smooth optical 9Be 197Au collisions exploring the role of nuclear deformations in smaller, asymmetriccollisions.

Page 18: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

18

compasses regions having significant gradients, such asin heavy-ion collisions like 208Pb 208Pb and presumablyas well for smaller systems with sub-nucleonic structure.And importantly, even for smooth density profiles, thesescalings can change significantly due to the presence ofdeformed nuclei, leading in particular to a v2{2} whichgrows with multiplicity in ultracentral collisions, as first

predicted in Ref. [54].

REFERENCES

[1] H. Niemi, K. J. Eskola, R. Paatelainen, and K. Tuomi-nen, Phys. Rev. C93, 014912 (2016), arXiv:1511.04296[hep-ph].

[2] J. Noronha-Hostler, M. Luzum, and J.-Y. Ollitrault,Phys. Rev. C93, 034912 (2016), arXiv:1511.06289 [nucl-th].

[3] K. J. Eskola, H. Niemi, R. Paatelainen, and K. Tuomi-nen, Phys. Rev. C97, 034911 (2018), arXiv:1711.09803[hep-ph].

[4] G. Giacalone, J. Noronha-Hostler, M. Luzum, andJ.-Y. Ollitrault, Phys. Rev. C97, 034904 (2018),arXiv:1711.08499 [nucl-th].

[5] C. Gale, S. Jeon, B. Schenke, P. Tribedy, andR. Venugopalan, Phys. Rev. Lett. 110, 012302 (2013),arXiv:1209.6330 [nucl-th].

[6] J. E. Bernhard, J. S. Moreland, S. A. Bass, J. Liu,and U. Heinz, Phys. Rev. C94, 024907 (2016),arXiv:1605.03954 [nucl-th].

[7] S. Chatrchyan et al. (CMS), Phys. Lett. B724, 213(2013), arXiv:1305.0609 [nucl-ex].

[8] M. Aaboud et al. (ATLAS), Eur. Phys. J. C77, 428(2017), arXiv:1705.04176 [hep-ex].

[9] M. Aaboud et al. (ATLAS), Phys. Rev. C97, 024904(2018), arXiv:1708.03559 [hep-ex].

[10] G. Aad et al. (ATLAS), Phys. Lett. B725, 60 (2013),arXiv:1303.2084 [hep-ex].

[11] A. M. Sirunyan et al. (CMS), Phys. Rev. Lett. 121,082301 (2018), arXiv:1804.09767 [hep-ex].

[12] V. Khachatryan et al. (CMS), Phys. Lett. B742, 200(2015), arXiv:1409.3392 [nucl-ex].

[13] V. Khachatryan et al. (CMS), Phys. Rev. Lett. 115,012301 (2015), arXiv:1502.05382 [nucl-ex].

[14] V. Khachatryan et al. (CMS), Phys. Rev. C92, 034911(2015), arXiv:1503.01692 [nucl-ex].

[15] A. M. Sirunyan et al. (CMS), Phys. Rev. Lett. 120,092301 (2018), arXiv:1709.09189 [nucl-ex].

[16] B. B. Abelev et al. (ALICE), Phys. Lett. B726, 164(2013), arXiv:1307.3237 [nucl-ex].

[17] B. B. Abelev et al. (ALICE), Phys. Rev. C90, 054901(2014), arXiv:1406.2474 [nucl-ex].

[18] A. Adare et al. (PHENIX), Phys. Rev. Lett. 111, 212301(2013), arXiv:1303.1794 [nucl-ex].

[19] A. Adare et al. (PHENIX), Phys. Rev. Lett. 114, 192301(2015), arXiv:1404.7461 [nucl-ex].

[20] C. Aidala et al. (PHENIX), (2018), arXiv:1805.02973[nucl-ex].

[21] A. Adare et al. (PHENIX), (2018), arXiv:1807.11928[nucl-ex].

[22] A. Adare et al. (PHENIX), Phys. Rev. Lett. 115, 142301(2015), arXiv:1507.06273 [nucl-ex].

[23] C. Aidala et al., Phys. Rev. C95, 034910 (2017),arXiv:1609.02894 [nucl-ex].

[24] A. Adare et al. (PHENIX), Phys. Rev. C97, 064904(2018), arXiv:1710.09736 [nucl-ex].

[25] A. Adare et al. (PHENIX), Phys. Rev. C98, 014912(2018), arXiv:1711.09003 [hep-ex].

[26] C. Aidala et al. (PHENIX), Phys. Rev. C96, 064905(2017), arXiv:1708.06983 [nucl-ex].

[27] C. Aidala et al. (PHENIX), Phys. Rev. Lett. 120, 062302(2018), arXiv:1707.06108 [nucl-ex].

[28] P. Bozek, Phys. Rev. C85, 014911 (2012),arXiv:1112.0915 [hep-ph].

[29] P. Bozek and W. Broniowski, Phys. Lett. B718, 1557(2013), arXiv:1211.0845 [nucl-th].

[30] P. Bozek, W. Broniowski, and G. Torrieri, Phys. Rev.Lett. 111, 172303 (2013), arXiv:1307.5060 [nucl-th].

[31] P. Bozek and W. Broniowski, Phys. Rev. C88, 014903(2013), arXiv:1304.3044 [nucl-th].

[32] I. Kozlov, M. Luzum, G. Denicol, S. Jeon, and C. Gale,(2014), arXiv:1405.3976 [nucl-th].

[33] Y. Zhou, X. Zhu, P. Li, and H. Song, Phys. Rev. C91,064908 (2015), arXiv:1503.06986 [nucl-th].

[34] W. Zhao, Y. Zhou, H. Xu, W. Deng, and H. Song, Phys.Lett. B780, 495 (2018), arXiv:1801.00271 [nucl-th].

[35] H. Mntysaari, B. Schenke, C. Shen, and P. Tribedy,Phys. Lett. B772, 681 (2017), arXiv:1705.03177 [nucl-th].

[36] R. D. Weller and P. Romatschke, Phys. Lett. B774, 351(2017), arXiv:1701.07145 [nucl-th].

[37] M. Greif, C. Greiner, B. Schenke, S. Schlichting, andZ. Xu, Phys. Rev. D96, 091504 (2017), arXiv:1708.02076[hep-ph].

[38] B. Schenke, S. Schlichting, P. Tribedy, andR. Venugopalan, Phys. Rev. Lett. 117, 162301 (2016),arXiv:1607.02496 [hep-ph].

[39] H. Mntysaari and B. Schenke, Phys. Rev. Lett. 117,052301 (2016), arXiv:1603.04349 [hep-ph].

[40] J. L. Albacete, H. Petersen, and A. Soto-Ontoso, Phys.Lett. B778, 128 (2018), arXiv:1707.05592 [hep-ph].

[41] Z. Citron et al. (2018) arXiv:1812.06772 [hep-ph].[42] M. D. Sievert and J. Noronha-Hostler, (2019),

arXiv:1901.01319 [nucl-th].[43] S. H. Lim, J. Carlson, C. Loizides, D. Lonardoni, J. E.

Lynn, J. L. Nagle, J. D. Orjuela Koop, and J. Ouellette,(2018), arXiv:1812.08096 [nucl-th].

[44] S. Huang, Z. Chen, J. Jia, and W. Li, (2019),arXiv:1904.10415 [nucl-ex].

[45] A. V. Giannini, F. Grassi, and M. Luzum, (2019),arXiv:1904.11488 [nucl-th].

[46] L. Adamczyk et al. (STAR), Phys. Rev. Lett. 115,222301 (2015), arXiv:1505.07812 [nucl-ex].

[47] H. Wang and P. Sorensen (STAR), Proceedings, 6thInternational Conference on Hard and ElectromagneticProbes of High-Energy Nuclear Collisions (Hard Probes

Page 19: dAu BeAu He, and HeAu Collisions · dAu, 3He Au, and possibly even pp collisions) [7{27]. While the predictions of relativistic hydrodynamics have been reasonably consistent with

19

2013): Cape Town, South Africa, November 4-8, 2013,Nucl. Phys. A932, 169 (2014), arXiv:1406.7522 [nucl-ex].

[48] J. S. Moreland, J. E. Bernhard, and S. A. Bass, Phys.Rev. C92, 011901 (2015), arXiv:1412.4708 [nucl-th].

[49] A. Goldschmidt, Z. Qiu, C. Shen, and U. Heinz, Phys.Rev. C92, 044903 (2015), arXiv:1507.03910 [nucl-th].

[50] B. Schenke, C. Shen, and P. Tribedy, (2019),arXiv:1901.04378 [nucl-th].

[51] C. Collaboration (CMS), (2018).[52] S. Acharya et al. (ALICE), Phys. Lett. B784, 82 (2018),

arXiv:1805.01832 [nucl-ex].[53] T. A. collaboration (ATLAS), CERN (CERN, Geneva,

2018).[54] Y. V. Kovchegov and D. E. Wertepny, Nucl. Phys. A925,

254 (2014), arXiv:1310.6701 [hep-ph].[55] P. Dasgupta, R. Chatterjee, and D. K. Srivastava, Phys.

Rev. C95, 064907 (2017), arXiv:1609.01949 [nucl-th].[56] G. Giacalone, Phys. Rev. C99, 024910 (2019),

arXiv:1811.03959 [nucl-th].[57] T. Lappi, Phys. Lett. B643, 11 (2006), arXiv:hep-

ph/0606207 [hep-ph].[58] G. Giacalone, J. Noronha-Hostler, and J.-Y. Ollitrault,

Phys. Rev. C95, 054910 (2017), arXiv:1702.01730 [nucl-th].

[59] J. E. Bernhard, P. W. Marcy, C. E. Coleman-Smith,S. Huzurbazar, R. L. Wolpert, and S. A. Bass, Phys.Rev. C91, 054910 (2015), arXiv:1502.00339 [nucl-th].

[60] J. Carlson and R. Schiavilla, Rev. Mod. Phys. 70, 743(1998).

[61] M. Luzum and H. Petersen, J. Phys. G41, 063102 (2014),arXiv:1312.5503 [nucl-th].

[62] A. Dumitru, L. McLerran, and V. Skokov, Phys. Lett.B743, 134 (2015), arXiv:1410.4844 [hep-ph].

[63] A. Dumitru, A. V. Giannini, and V. Skokov, (2015),arXiv:1503.03897 [hep-ph].

[64] Y. V. Kovchegov and D. E. Wertepny, Nucl. Phys. A906,50 (2013), arXiv:1212.1195 [hep-ph].

[65] Y. V. Kovchegov and V. V. Skokov, Phys. Rev. D97,094021 (2018), arXiv:1802.08166 [hep-ph].

[66] L. Hulthen and M. Sagawara, Structure of Atomic Nuclei,Handbuch der Physik , 39:14 (1957).

[67] J. S. Moreland, J. E. Bernhard, and S. A. Bass, (2018),arXiv:1808.02106 [nucl-th].

[68] J. Noronha-Hostler, J. Noronha, and M. Gyulassy, Phys.Rev. C93, 024909 (2016), arXiv:1508.02455 [nucl-th].

[69] F. G. Gardim, F. Grassi, P. Ishida, M. Luzum, P. S. Ma-galhes, and J. Noronha-Hostler, Phys. Rev. C97, 064919(2018), arXiv:1712.03912 [nucl-th].

[70] P. Moller, A. J. Sierk, T. Ichikawa, and H. Sagawa,Atom. Data Nucl. Data Tabl. 109-110, 1 (2016),

arXiv:1508.06294 [nucl-th].[71] T. Hirano, P. Huovinen, and Y. Nara, Phys. Rev. C84,

011901 (2011), arXiv:1012.3955 [nucl-th].[72] D. Teaney and L. Yan, Phys. Rev. C83, 064904 (2011),

arXiv:1010.1876 [nucl-th].[73] Z. Qiu, C. Shen, and U. Heinz, Phys. Lett. B707, 151

(2012), arXiv:1110.3033 [nucl-th].[74] Z. Qiu and U. W. Heinz, Phys. Rev. C84, 024911 (2011),

arXiv:1104.0650 [nucl-th].[75] F. G. Gardim, F. Grassi, M. Luzum, and J.-Y. Ollitrault,

Phys. Rev. C85, 024908 (2012), arXiv:1111.6538 [nucl-th].

[76] H. Niemi, G. S. Denicol, H. Holopainen, and P. Huovi-nen, Phys. Rev. C87, 054901 (2013), arXiv:1212.1008[nucl-th].

[77] D. Teaney and L. Yan, Phys. Rev. C86, 044908 (2012),arXiv:1206.1905 [nucl-th].

[78] F. G. Gardim, J. Noronha-Hostler, M. Luzum,and F. Grassi, Phys. Rev. C91, 034902 (2015),arXiv:1411.2574 [nucl-th].

[79] B. Betz, M. Gyulassy, M. Luzum, J. Noronha,J. Noronha-Hostler, I. Portillo, and C. Ratti, Phys. Rev.C95, 044901 (2017), arXiv:1609.05171 [nucl-th].

[80] J. Noronha-Hostler, L. Yan, F. G. Gardim, andJ.-Y. Ollitrault, Phys. Rev. C93, 014909 (2016),arXiv:1511.03896 [nucl-th].

[81] J. Noronha-Hostler, G. S. Denicol, J. Noronha, R. P. G.Andrade, and F. Grassi, Phys. Rev. C88, 044916 (2013),arXiv:1305.1981 [nucl-th].

[82] J. Noronha-Hostler, J. Noronha, and F. Grassi, Phys.Rev. C90, 034907 (2014), arXiv:1406.3333 [nucl-th].

[83] P. Alba, V. Mantovani Sarti, J. Noronha, J. Noronha-Hostler, P. Parotto, I. Portillo Vazquez, and C. Ratti,Phys. Rev. C98, 034909 (2018), arXiv:1711.05207 [nucl-th].

[84] M. Mace, V. V. Skokov, P. Tribedy, and R. Venugopalan,(2018), arXiv:1807.00825 [hep-ph].

[85] B. I. Abelev et al. (STAR), Phys. Rev. C79, 034909(2009), arXiv:0808.2041 [nucl-ex].

[86] P. Bozek and W. Broniowski, Phys. Rev. Lett. 121,202301 (2018), arXiv:1808.09840 [nucl-th].

[87] S. M. Lukyanov, A. S. Denikin, E. I. Voskoboynik, S. V.Khlebnikov, V. A. Maslov, Yu. E. Penionzhkevich, Yu. G.Sobolev, W. H. Trzaska, G. P. Tyurin, and M. N.Harakeh, J. Phys. G41, 035102 (2014), arXiv:1310.2965[nucl-ex].

[88] M. Burkardt, Int. J. Mod. Phys. A18, 173 (2003),arXiv:hep-ph/0207047 [hep-ph].