Role of clustered nuclear geometry in particle production through
p–C and p–O collisions at the Large Hadron Collider

Aswathy Menon K R Aswathy.Menon@cern.ch    Suraj Prasad Suraj.Prasad@cern.ch    Neelkamal Mallick Neelkamal.Mallick@cern.ch    Raghunath Sahoo Raghunath.Sahoo@cern.ch Department of Physics, Indian Institute of Technology Indore, Simrol, Indore 453552, India
(July 4, 2024)
Abstract

Long-range multi-particle correlations in heavy-ion collisions have shown conclusive evidence of the hydrodynamic behavior of strongly interacting matter, and are associated with the final-state azimuthal momentum anisotropy. In small collision systems, azimuthal anisotropy can be influenced by the hadronization mechanism and residual jet-like correlations. Thus, one of the motives of the planned p–O and O–O collisions at the LHC and RHIC is to understand the origin of small system collectivity. As the anisotropic flow coefficients (vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) are sensitive to the initial-state effects including nuclear shape, deformation, and charge density profiles, studies involving 12C and 16O nuclei are transpiring due to the presence of exotic α𝛼\alphaitalic_α (4He) clusters in such nuclei. In this study, for the first time, we investigate the effects of nuclear α𝛼\alphaitalic_α–clusters on the azimuthal anisotropy of the final-state hadrons in p–C and p–O collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV within a multi-phase transport model framework. We report the transverse momentum (pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT) and pseudorapidity (η𝜂\etaitalic_η) spectra, participant eccentricity (ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangularity (ϵ3subscriptitalic-ϵ3\epsilon_{3}italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT), and estimate the elliptic flow (v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangular flow (v3subscript𝑣3v_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) of the final-state hadrons using the two-particle cumulant method. These results are compared with a model-independent Sum of Gaussians (SOG) type nuclear density profile for 12C and 16O nuclei.

I Introduction

The world’s most powerful particle accelerators, the Relativistic Heavy-Ion Collider (RHIC) at BNL, USA, and the Large Hadron Collider (LHC) at CERN, Switzerland, offer unparalleled experimental capabilities to perform high-energy hadronic and nuclear collisions to study the behavior of nuclear matter under extreme conditions of temperature and baryon density. Such ultra-relativistic heavy-ion collisions produce a state of locally thermalized and deconfined partonic medium, known as the quark-gluon plasma (QGP). While the existence of QGP has long been established in heavy-ion collisions, its presence in small collision systems is a matter of intense research, recently. Thanks to the recent measurements of heavy-ion-like features in high multiplicity pp collisions such as strangeness enhancement [1], ridge-like structure [3, 2], and radial flow effects [4, 5, 6], the possible existence of QGP droplets in small collision systems needs to be re-examined. As p–C and p–O collisions fill the multiplicity gap between pp to p–Pb and peripheral Pb–Pb collisions, it provides a perfect system size for studying several heavy-ion-like effects in small collision systems at the RHIC and LHC energies [7, 8].

Refer to caption
Figure 1: (Color online) Pictorial representation of the arrangement of α𝛼\alphaitalic_α–particles inside 12C (left) and 16O (right) nucleus.

In non-central heavy-ion collisions, the appearance of strong final-state azimuthal momentum anisotropy is associated with the hydrodynamic collective expansion of the medium [9, 10]. This collectivity is usually quantified as the Fourier coefficients (vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) of the final-state azimuthal momentum distribution (dN/dϕ𝑑𝑁𝑑italic-ϕdN/d\phiitalic_d italic_N / italic_d italic_ϕ[11, 12]. Azimuthal anisotropy mainly arises from the initial nuclear geometry and fluctuations in the energy and entropy density, which further gets embedded in the system through its equation-of-state and transport coefficients [9, 10]. Hydrodynamic model predictions when confronted with the experimental findings have shown conclusive evidence of the collective expansion of the thermalized medium and strongly hint at the formation of QGP in heavy-ion collisions [13]. However, such collective behavior is usually not anticipated in small collision systems, thus, the applicability of hydrodynamics, and hence, thermalization in the early stages of the collision is still debatable in small collision systems.

The anisotropic flow coefficients, such as elliptic flow (v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangular flow (v3subscript𝑣3v_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT), are found to be sensitive to the initial nuclear distribution, nuclear shape deformation, and fluctuations in the nuclear overlap region [16, 14, 15]. In Xe–Xe collisions at the LHC, the presence of a deformed nuclear structure gives rise to a higher value of elliptic flow in the most central collisions compared to Pb–Pb collisions [17, 18, 20, 19]. In addition, a larger density fluctuation due to a smaller number of participating nucleons in Xe–Xe collisions gives rise to a larger value of triangular flow in contrast to Pb–Pb collisions [17, 18]. A similar study using AMPT in U–U collisions at the RHIC energy shows that the values of anisotropic flow coefficients change with a change in the deformation parameter of the nuclear distribution [14]. However, the impact of this deformation is larger for elliptic flow in contrast to the triangular flow [14, 15, 21]. In addition, recent observations of ridge-like structures in p–Pb collisions have brought significant attention to asymmetric collisions, which are customarily considered to provide a baseline measurement for the cold nuclear effects [22, 23, 24, 25, 6], where the presence of strong long-range multi-particle correlations are not expected. Further, the measured values of anisotropic flow coefficients in p–Pb collisions are comparable to that of peripheral Pb–Pb collisions [27, 26, 28]. Although, in p–Pb collisions, the ordering of the magnitude of flow coefficients, i.e., v2>v3>v4subscript𝑣2subscript𝑣3subscript𝑣4v_{2}>v_{3}>v_{4}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT > italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT, is not as strong as in Pb–Pb collisions, the bare presence of signatures of collectivity in p–Pb collisions makes the study of asymmetric collisions a vital contribution of the heavy-ion physics community.

Nuclei having 4n4𝑛4n4 italic_n number of nucleons, are theorized to have α𝛼\alphaitalic_α–clusters (4He-nucleus), which include 8Be, 12C, and 16O nuclei. In 12C, the α𝛼\alphaitalic_α–particles arrange themselves at the corners of an equilateral triangle. Similarly, nucleons inside 16O are expected to configure themselves in four α𝛼\alphaitalic_α–clusters forming a regular tetrahedral geometry [29, 30, 31, 32, 33, 34]. Figure 1 shows the pictorial representation of the arrangement of nucleons inside an α𝛼\alphaitalic_α–particle, and the arrangement of α𝛼\alphaitalic_α–clusters inside 12C (left) and 16O (right) nuclei. Studies involving 12C and 16O nuclei have the potential to explore the emergence of collectivity in small systems in addition to the study of α𝛼\alphaitalic_α–cluster nuclear geometry effects on the final-state observables. Additionally, studies involving forward kinematics in p–O collisions are crucial in understanding the interaction of cosmic rays (mostly protons) with the Earth’s atmosphere (mostly 14N and 16O). It could also help resolve the outstanding muon puzzle and perform precise measurements of the π0superscript𝜋0\pi^{0}italic_π start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT energy fraction (R=Eπ0/Eallhadrons𝑅subscript𝐸superscript𝜋0subscript𝐸allhadronsR=E_{\pi^{0}}/E_{\rm all~{}hadrons}italic_R = italic_E start_POSTSUBSCRIPT italic_π start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT roman_all roman_hadrons end_POSTSUBSCRIPT), a smaller value of which is potentially attributed to strangeness enhancement, and hence hints towards the formation of QGP droplets [7].

With all these tempting motivations, in recent years, several studies have been performed involving collisions of 12C and 16O nuclei [40, 41, 39, 35, 36, 37, 16, 38]. It includes the studies based on different hydrodynamic models [42, 43, 44], Glauber Monte Carlo [35, 36, 37], parton energy loss [45], and jet quenching effects [46]. Some of these studies also investigate the possible signatures of α𝛼\alphaitalic_α–clusters in 12C and 16O nuclei by studying the final-state particle production and anisotropic flow measurements [40, 41, 39, 48, 49, 50, 51, 16, 47]. In Ref. [40], an enhanced value of triangular flow in C–Au collisions is observed for an α𝛼\alphaitalic_α–clustered 12C nucleus. In Ref. [16], authors show that the presence of an α𝛼\alphaitalic_α–clustered structure in 16O nucleus leads to a high value of triangular flow in most central O–O collisions, which are well observed in the ratio v3/v2subscript𝑣3subscript𝑣2v_{3}/v_{2}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. In addition, an away-side broadening in the two-particle correlation function is also reported for O–O collisions with α𝛼\alphaitalic_α–clusters as compared to the Woods-Saxon nuclear profile [16]. Although there have been a few studies that aim to establish a clear signature of the presence of α𝛼\alphaitalic_α–clusters in 16O and 12C nucleus, the choice of p–O and p–C is novel to this study. In addition, the studies of anisotropic flow in asymmetric collision systems like p–O and p–C can accord the observations of collectivity in p–Pb collisions.

In this study, we incorporate an α𝛼\alphaitalic_α–cluster-type initial geometry for 12C and 16O nuclei, and report the first measurements of transverse momentum (pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT) and pseudorapidity (η𝜂\etaitalic_η) spectra, participant eccentricity (ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangularity (ϵ3subscriptitalic-ϵ3\epsilon_{3}italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT), and elliptic flow (v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangular flow (v3subscript𝑣3v_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) of the final-state hadrons, and their scaling in p–O and p–C collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using a multi-phase transport model (AMPT). We use the two-particle cumulant method with a relative pseudorapidity cut to minimize the effects of nonflow correlations. These results are compared with a model-independent Sum of Gaussians (SOG) type nuclear density profile for 12C and 16O nuclei.

The paper is organized as follows. We start with a brief introduction and motivation of the study in Section I followed by a description of event generation and methodology in Section II. In Section III, we present and discuss the results. Finally, in Section IV, we summarize the study with important findings and present a brief outlook for future research.

II Event generation and methodology

In this section, we begin with a brief introduction to a multi-phase transport model, which is used to simulate p–C and p–O collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV. The technical details of the implementation of different nuclear profiles for 16O and 12C nuclei are discussed next. Finally, a detailed description of the two-particle cumulant method for the anisotropic flow estimation is provided.

II.1 A multi-phase transport model

AMPT is a widely used Monte Carlo transport model constructed to describe the space-time evolution of pp, p+A, and A+A collisions across RHIC and LHC energies [53, 52]. It incorporates the initial deconfined partonic phase, the final hadronic interactions, and the transition between these two phases of matter. AMPT model has four major components as described below.

  • Initialization: The initial spatial and momentum distribution of partons, are generated by the HIJING model [54]. Multiple scatterings among incoming nucleons are treated in the framework of eikonal formalism. The hard and soft components of particle production are respectively modeled by the formation of energetic minijet partons and soft string excitations. The differential cross-section of the produced mini-jet partons and excited strings are first calculated for pp collisions and are then converted into p+A and A+A collisions by using the in-built Glauber model [53, 54].

  • Transport of partons: The produced partons are propagated to the Zhang’s Parton Cascade (ZPC) model, where the Boltzmann transport equation for partons is solved via cascade method [55]. In the string-melting version of AMPT, all excited strings are converted into hadrons and then decomposed further into their constituent quarks [56]. These partons are then combined with the mini-jet partons and evolve through the ZPC model via two-body elastic parton scatterings [53].

  • Hadronization: The transported partons are then hadronized using either the default mode or the string melting mode of AMPT. In the default mode, the transported partons are combined with their parent strings via the Lund string fragmentation model and then the strings get converted into hadrons [57]. In the string melting mode, transported partons are combined to form hadrons through a quark coalescence mechanism [58, 56]. In this model, a quark and an anti-quark sharing a close phase-space combine to form a meson while the three nearest quarks combine into baryons [53, 58].

  • Hadron transport: Evolution of the resulting hadronic matter is described via hadron cascade, which is based on a relativistic transport (ART) model [59, 60]. The produced hadrons go through a final evolution via baryon-baryon, meson-baryon, and meson-meson scatterings and decays [53].

As the quark coalescence mechanism for hadronization implemented in the string melting version of AMPT explains the particle pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra and flow at intermediate-pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT, we use the string melting mode of AMPT (version 2.26t9b) for this study [62, 63, 64, 65, 38, 67, 61, 16, 66]. For this study, in AMPT, we take the partonic scattering cross-section to be σgg=3subscript𝜎gg3\sigma_{\rm gg}=3italic_σ start_POSTSUBSCRIPT roman_gg end_POSTSUBSCRIPT = 3 mb and the value of strong coupling constant as αs=0.33subscript𝛼𝑠0.33\alpha_{s}=0.33italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0.33. The Lund symmetric splitting function parameters are set as a=0.3𝑎0.3a=0.3italic_a = 0.3 and b=0.15𝑏0.15b=0.15italic_b = 0.15 [42]. The parton screening mass in the ZPC model is fixed to μ=2.265𝜇2.265\mu=2.265italic_μ = 2.265 fm-1. With the above-mentioned settings in AMPT, we simulate p–O and p–C collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV.

Since the impact parameter (b𝑏bitalic_b), the number of participants (Npartsubscript𝑁partN_{\rm part}italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT), and the number of binary collisions (Ncollsubscript𝑁collN_{\rm coll}italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT) cannot be directly measured in experiments, we resort to Glauber model estimations for the same [68, 69]. Using the publicly available MC Glauber code (TGlauberMC3.23.2-3.2- 3.2[69], the impact parameter distribution is sliced to obtain the required centralities of collision. We have also modified the in-built HIJING model in AMPT to incorporate SOG and αlimit-from𝛼\alpha-italic_α -cluster nuclear density profiles for both O16superscriptO16{}^{16}\rm Ostart_FLOATSUPERSCRIPT 16 end_FLOATSUPERSCRIPT roman_O and C12superscriptC12{}^{12}\rm Cstart_FLOATSUPERSCRIPT 12 end_FLOATSUPERSCRIPT roman_C nuclei. Due to the unavailability of experimental data for p–O and p–C collisions at the LHC, we tune the AMPT model parameters by comparing them with that of the p–Pb system. A detailed description of the parameter tuning is provided in the Appendix.

II.2 Nuclear density profiles

This study implements two different nuclear density profiles for the 12C and 16O nuclei, which are α𝛼\alphaitalic_α–cluster type geometrical distribution and a model-independent Sum of Gaussians type nuclear density profile. Here, the technical details regarding the implementation of the two density profiles are discussed.

II.2.1 Sum Of Gaussians (SOG) density profile

To represent the nuclear charge density and extract charge density parameters in a model-independent fashion as approximated in experiments, one fits the nuclear densities with a sum of Gaussian functions (SOG) [70, 71]. SOG can accurately represent complex nuclear densities by exploiting the parameters of individual Gaussian functions, which makes SOG applicable to a broad list of nuclei. SOG can produce smooth and continuous descriptions of the nuclear density profiles, consequently simplifying many theoretical calculations due to the well-known properties of the Gaussian distribution function. In addition, SOG can produce a more realistic and precise description of the nuclear profiles as compared to the traditional uniform density or a Woods-Saxon description of the nucleus [70]. In this study, we use the sum of two Gaussian functions to simulate 12C and 16O nuclear profiles. The SOG nuclear charge density (ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r )) as a function of radial distance (r𝑟ritalic_r) is expanded as,

ρ(r)=C1ea1r2+C2ea2r2.𝜌𝑟subscript𝐶1superscript𝑒subscript𝑎1superscript𝑟2subscript𝐶2superscript𝑒subscript𝑎2superscript𝑟2\rho(r)=C_{1}e^{-a_{1}r^{2}}+C_{2}e^{-a_{2}r^{2}}.italic_ρ ( italic_r ) = italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT . (1)

Here, the coefficients C1subscript𝐶1C_{1}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and a2subscript𝑎2a_{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT for the respective nuclei are obtained from Ref. [70]. Due to the rapid decrease of the Gaussian tail, the values of the charge density at various radii (ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r )) are advantageously decoupled. Table 1 shows the parameters of SOG that have been considered in this study to simulate the 12C and 16O nuclei.

Table 1: Parameters of Sum of Gaussians nuclear density profile chosen for 12C and 16O nucleus [70].
Nucleus C1subscript𝐶1C_{1}italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT a2subscript𝑎2a_{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
Carbon (12C) -0.162 0.340 0.554 0.280
Oxygen (16O) -0.539 0.729 0.407 0.300

II.2.2 α𝛼\alphaitalic_α–clustered nuclear geometry

The 4He nucleus consisting of a pair of protons and a pair of neutrons is known as the α𝛼\alphaitalic_α–particle. As discussed earlier, nucleons of certain light nuclei can cluster into groups of α𝛼\alphaitalic_α–particles forming a stable geometrical shape. For this, the nuclei should have 4n4𝑛4n4 italic_n (n𝑛nitalic_n is a positive integer) number of nucleons. In 12C, three such α𝛼\alphaitalic_α–particles cluster into forming an equilateral triangle, whereas in 16O, four α𝛼\alphaitalic_α–particles cluster into a regular tetrahedral arrangement. This α𝛼\alphaitalic_α–cluster geometry is believed to account for the additional stability of the nucleus. In our study, for the first time, we have implemented such α𝛼\alphaitalic_α–cluster geometry for 12C and 16O nuclei to study the initial-state effects in particle production through p–C and p–O collisions at the LHC. The technical details of the implementation of α𝛼\alphaitalic_α–cluster structure for 12C and 16O are discussed below.

  • Tetrahedral geometry for 16O

    For the O16superscriptO16{}^{16}\rm Ostart_FLOATSUPERSCRIPT 16 end_FLOATSUPERSCRIPT roman_O nucleus, four α𝛼\alphaitalic_α–particles are situated at the four vertices of a regular tetrahedron with side length 3.42 fm, which makes the rms radius of O16superscriptO16{}^{16}\rm Ostart_FLOATSUPERSCRIPT 16 end_FLOATSUPERSCRIPT roman_O to be 2.699 fm [39, 16, 38]. One can visualize the nucleons inside the 16O nuclei as depicted in Fig. 1. Nucleons inside the α𝛼\alphaitalic_α–particle are sampled using the following Woods-Saxon density profile in terms of a three-parameter Fermi (3pF) distribution,

    ρ(r)=ρ0(1+w(rr0)2)1+exp(rr0a).𝜌𝑟subscript𝜌01𝑤superscript𝑟subscript𝑟021𝑟subscript𝑟0𝑎\rho(r)=\frac{\rho_{0}\Big{(}1+w\big{(}\frac{r}{r_{0}}\big{)}^{2}\Big{)}}{1+% \exp\big{(}\frac{r-r_{0}}{a}\big{)}}.italic_ρ ( italic_r ) = divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 + italic_w ( divide start_ARG italic_r end_ARG start_ARG italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG 1 + roman_exp ( divide start_ARG italic_r - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG ) end_ARG . (2)

    Here, ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r ) is the nuclear charge density at a radial distance r𝑟ritalic_r from the center of the nucleus. r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT refers to the mean radius of the nucleus, w𝑤witalic_w is the nuclear deformation parameter, and a𝑎aitalic_a is the skin depth. The values of these parameters are chosen for the 4He nucleus as r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.964 fm, w𝑤witalic_w = 0.517, and a𝑎aitalic_a = 0.322 fm, which corresponds to the rms radius of 1.676 fm for the α𝛼\alphaitalic_α–particle. To include a finite volume effect for each nucleon, the minimum separation distance between any two nucleons is set to be 0.4 fm. The spatial orientation of the tetrahedron is randomized before each collision for both the target and projectile nuclei.

  • Triangular geometry for 12C

    For the 12C nucleus, three α𝛼\alphaitalic_α–particles arrange themselves at the corners of an equilateral triangle of side length 3.10 fm [49, 72]. One can visualize the nucleons inside the 12C nuclei as depicted in Fig. 1. These parameters lead to a rms radius of 2.47 fm for the 12C nucleus. The nucleons inside the α𝛼\alphaitalic_α–particles are sampled using a similar procedure as discussed above.

Table 2: The impact parameter, average number of nucleon-nucleon binary collisions and average number of nucleon participants for different density profiles and in different centrality classes for p–O collisions at sNNsubscript𝑠NN\sqrt{s_{\rm{NN}}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV. Centrality selection is done through geometrical slicing, i.e., from the impact parameter obtained from the Glauber model.
p–O SOG α𝛼\alphaitalic_α–cluster
Centrality(%) bminsubscript𝑏minb_{\rm{min}}italic_b start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT(fm) bmaxsubscript𝑏maxb_{\rm{max}}italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT(fm) Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm{coll}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm{part}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms bminsubscript𝑏minb_{\rm{min}}italic_b start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT(fm) bmaxsubscript𝑏maxb_{\rm{max}}italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT(fm) Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm{coll}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm{part}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms
0-5 0 0.81 6.64 ±plus-or-minus\pm± 2.12 7.64 ±plus-or-minus\pm± 2.12 0 0.80 5.65 ±plus-or-minus\pm± 1.64 6.65 ±plus-or-minus\pm± 1.64
5-10 0.81 1.15 5.94 ±plus-or-minus\pm± 1.92 6.94 ±plus-or-minus\pm± 1.92 0.80 1.14 5.30 ±plus-or-minus\pm± 1.64 6.30 ±plus-or-minus\pm± 1.64
10-20 1.15 1.63 5.00 ±plus-or-minus\pm± 1.75 6.00 ±plus-or-minus\pm± 1.75 1.14 1.61 4.78 ±plus-or-minus\pm± 1.56 5.78 ±plus-or-minus\pm± 1.56
20-30 1.63 2.00 3.96 ±plus-or-minus\pm± 1.56 4.96 ±plus-or-minus\pm± 1.56 1.61 1.97 4.08 ±plus-or-minus\pm± 1.40 5.08 ±plus-or-minus\pm± 1.40
30-40 2.00 2.32 3.16 ±plus-or-minus\pm± 1.40 4.16 ±plus-or-minus\pm± 1.40 1.97 2.28 3.43 ±plus-or-minus\pm± 1.27 4.43 ±plus-or-minus\pm± 1.27
40-50 2.32 2.60 2.54 ±plus-or-minus\pm± 1.24 3.54 ±plus-or-minus\pm± 1.24 2.28 2.56 2.84 ±plus-or-minus\pm± 1.15 3.84 ±plus-or-minus\pm± 1.15
50-60 2.60 2.89 2.08 ±plus-or-minus\pm± 1.06 3.08 ±plus-or-minus\pm± 1.06 2.56 2.82 2.37 ±plus-or-minus\pm± 1.03 3.37 ±plus-or-minus\pm± 1.03
60-70 2.89 3.18 1.73 ±plus-or-minus\pm± 0.87 2.73 ±plus-or-minus\pm± 0.87 2.82 3.08 1.99 ±plus-or-minus\pm± 0.90 2.99 ±plus-or-minus\pm± 0.90
70-100 3.18 7.50 1.26 ±plus-or-minus\pm± 0.54 2.26 ±plus-or-minus\pm± 0.54 3.08 5.38 1.43 ±plus-or-minus\pm± 0.65 2.43 ±plus-or-minus\pm± 0.65
Table 3: The impact parameter, average number of nucleon-nucleon binary collisions and average number of nucleon participants for different density profiles and in different centrality classes for p–C collisions at sNNsubscript𝑠NN\sqrt{s_{\rm{NN}}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV. Centrality selection is done through geometrical slicing, i.e., from the impact parameter obtained from the Glauber model.
p–C SOG α𝛼\alphaitalic_α–cluster
Centrality(%) bminsubscript𝑏minb_{\rm{min}}italic_b start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT(fm) bmaxsubscript𝑏maxb_{\rm{max}}italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT(fm) Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm{coll}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm{part}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms bminsubscript𝑏minb_{\rm{min}}italic_b start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT(fm) bmaxsubscript𝑏maxb_{\rm{max}}italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT(fm) Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm{coll}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm{part}}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ ±plus-or-minus\pm± rms
0-5 0 0.76 5.49 ±plus-or-minus\pm± 1.84 6.49 ±plus-or-minus\pm± 1.84 0 0.71 6.05 ±plus-or-minus\pm± 1.89 7.05 ±plus-or-minus\pm± 1.89
5-10 0.76 1.08 4.91 ±plus-or-minus\pm± 1.69 5.91 ±plus-or-minus\pm± 1.69 0.71 1.00 5.46 ±plus-or-minus\pm± 1.68 6.46 ±plus-or-minus\pm± 1.68
10-20 1.08 1.53 4.14 ±plus-or-minus\pm± 1.53 5.14 ±plus-or-minus\pm± 1.53 1.00 1.42 4.70 ±plus-or-minus\pm± 1.54 5.70 ±plus-or-minus\pm± 1.54
20-30 1.53 1.90 3.25 ±plus-or-minus\pm± 1.34 4.25 ±plus-or-minus\pm± 1.34 1.42 1.74 3.85 ±plus-or-minus\pm± 1.38 4.85 ±plus-or-minus\pm± 1.38
30-40 1.90 2.18 2.61 ±plus-or-minus\pm± 1.18 3.61 ±plus-or-minus\pm± 1.18 1.74 2.02 3.14 ±plus-or-minus\pm± 1.27 4.14 ±plus-or-minus\pm± 1.27
40-50 2.18 2.46 2.15 ±plus-or-minus\pm± 1.03 3.15 ±plus-or-minus\pm± 1.03 2.02 2.27 2.60 ±plus-or-minus\pm± 1.15 3.60 ±plus-or-minus\pm± 1.15
50-60 2.46 2.74 1.80 ±plus-or-minus\pm± 0.88 2.80 ±plus-or-minus\pm± 0.88 2.27 2.51 2.18 ±plus-or-minus\pm± 1.02 3.18 ±plus-or-minus\pm± 1.02
60-70 2.74 3.03 1.54 ±plus-or-minus\pm± 0.73 2.54 ±plus-or-minus\pm± 0.73 2.51 2.76 1.85 ±plus-or-minus\pm± 0.90 2.85 ±plus-or-minus\pm± 0.90
70-100 3.03 7.44 1.19 ±plus-or-minus\pm± 0.45 2.19 ±plus-or-minus\pm± 0.45 2.76 5.86 1.36 ±plus-or-minus\pm± 0.62 2.36 ±plus-or-minus\pm± 0.62
Table 4: Centrality dependence of average charged-particle multiplicity density for various nuclear profiles in p–O and p–C collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV in the range |ηlab|<subscript𝜂lababsent|\eta_{\rm lab}|<| italic_η start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT | < 0.5
Centrality(%) p–O p–C
dNch/dηSOGsubscriptdelimited-⟨⟩𝑑subscript𝑁𝑐𝑑𝜂SOG\langle dN_{ch}/d\eta\rangle_{\rm SOG}⟨ italic_d italic_N start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT / italic_d italic_η ⟩ start_POSTSUBSCRIPT roman_SOG end_POSTSUBSCRIPT dNch/dηαclustersubscriptdelimited-⟨⟩𝑑subscript𝑁𝑐𝑑𝜂𝛼cluster\langle dN_{ch}/d\eta\rangle_{\alpha{\rm-cluster}}⟨ italic_d italic_N start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT / italic_d italic_η ⟩ start_POSTSUBSCRIPT italic_α - roman_cluster end_POSTSUBSCRIPT dNch/dηSOGsubscriptdelimited-⟨⟩𝑑subscript𝑁𝑐𝑑𝜂SOG\langle dN_{ch}/d\eta\rangle_{\rm SOG}⟨ italic_d italic_N start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT / italic_d italic_η ⟩ start_POSTSUBSCRIPT roman_SOG end_POSTSUBSCRIPT dNch/dηαclustersubscriptdelimited-⟨⟩𝑑subscript𝑁𝑐𝑑𝜂𝛼cluster\langle dN_{ch}/d\eta\rangle_{\alpha{\rm-cluster}}⟨ italic_d italic_N start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT / italic_d italic_η ⟩ start_POSTSUBSCRIPT italic_α - roman_cluster end_POSTSUBSCRIPT
0-5 35.56 ±plus-or-minus\pm± 0.02 26.02 ±plus-or-minus\pm± 0.02 30.43 ±plus-or-minus\pm± 0.02 25.26 ±plus-or-minus\pm± 0.04
5-10 30.97 ±plus-or-minus\pm± 0.02 25.03 ±plus-or-minus\pm± 0.02 26.33 ±plus-or-minus\pm± 0.02 24.14 ±plus-or-minus\pm± 0.03
10-20 25.48 ±plus-or-minus\pm± 0.02 23.44 ±plus-or-minus\pm± 0.02 21.92 ±plus-or-minus\pm± 0.02 22.29 ±plus-or-minus\pm± 0.04
20-30 19.91 ±plus-or-minus\pm± 0.02 20.92 ±plus-or-minus\pm± 0.02 17.28 ±plus-or-minus\pm± 0.02 19.71 ±plus-or-minus\pm± 0.02
30-40 15.88 ±plus-or-minus\pm± 0.01 18.00 ±plus-or-minus\pm± 0.02 13.97 ±plus-or-minus\pm± 0.02 17.12 ±plus-or-minus\pm± 0.04
40-50 12.94 ±plus-or-minus\pm± 0.01 15.45 ±plus-or-minus\pm± 0.02 11.55 ±plus-or-minus\pm± 0.02 14.71 ±plus-or-minus\pm± 0.02
50-60 10.65 ±plus-or-minus\pm± 0.01 13.02 ±plus-or-minus\pm± 0.03 9.66 ±plus-or-minus\pm± 0.01 12.48 ±plus-or-minus\pm± 0.03
60-70 8.80 ±plus-or-minus\pm± 0.01 10.67 ±plus-or-minus\pm± 0.03 8.05 ±plus-or-minus\pm± 0.01 10.47 ±plus-or-minus\pm± 0.03
70-100 5.61 ±plus-or-minus\pm± 0.01 5.89 ±plus-or-minus\pm± 0.02 5.28 ±plus-or-minus\pm± 0.01 5.03 ±plus-or-minus\pm± 0.46

II.3 Two-particle cumulants

Azimuthal anisotropy is one of the key observables that can characterize the medium formed in relativistic heavy-ion collisions. The asymmetric pressure gradient created in the medium due to the initial spatial anisotropy of the nuclear overlap region can transform into the final state momentum space azimuthal anisotropy. This is also known as the anisotropic flow. The anisotropic flow can be quantified as the coefficients of Fourier expansion of the azimuthal momentum distribution of the final-state particles as follows [73],

dNdϕ=12π(1+n=12vncos[n(ϕψn)]).𝑑𝑁𝑑italic-ϕ12𝜋1superscriptsubscript𝑛12subscript𝑣𝑛𝑛italic-ϕsubscript𝜓𝑛\frac{dN}{d\phi}=\frac{1}{2\pi}\Big{(}1+\sum_{n=1}^{\infty}2v_{n}\cos[n(\phi-% \psi_{n})]\Big{)}.divide start_ARG italic_d italic_N end_ARG start_ARG italic_d italic_ϕ end_ARG = divide start_ARG 1 end_ARG start_ARG 2 italic_π end_ARG ( 1 + ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT 2 italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT roman_cos [ italic_n ( italic_ϕ - italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ] ) . (3)

Here, ϕitalic-ϕ\phiitalic_ϕ is the azimuthal angle, ψnsubscript𝜓𝑛\psi_{n}italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the n𝑛nitalic_nth harmonic event-plane angle, and vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT characterize the n𝑛nitalic_nth-order anisotropic flow coefficient. v1subscript𝑣1v_{1}italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT stands for the directed flow, v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT for elliptic flow and v3subscript𝑣3v_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT characterises the triangular flow, and so on. One can estimate the anisotropic flow of different orders using the following expression [73],

vn=cos[n(ϕψn)].subscript𝑣𝑛delimited-⟨⟩𝑛italic-ϕsubscript𝜓𝑛v_{n}=\langle\cos[n(\phi-\psi_{n})]\rangle.italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ⟨ roman_cos [ italic_n ( italic_ϕ - italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ] ⟩ . (4)

Here, delimited-⟨⟩\langle\dots\rangle⟨ … ⟩ represents the average over all particles in an event. The estimation of anisotropic flow coefficients from Eq. (4) is not slick since it requires ψnsubscript𝜓𝑛\psi_{n}italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT whose measurement is not trivial in experiments. In addition, the flow coefficients estimated using Eq. (4) are prone to non-flow effects, such as contributions from jets, and short-range resonance decays. These effects are more pronounced in small collision systems. Thus, to avoid these issues, multi-particle cumulant method is prescribed to estimate the anisotropic flow coefficients. In this method, one does not require the information of ψnsubscript𝜓𝑛\psi_{n}italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and the non-flow effects can also be significantly reduced by implementing a proper relative pseudorapidity (ΔηΔ𝜂\Delta\etaroman_Δ italic_η) cut between the particles.

In this study, the anisotropic flow coefficients are estimated using a two-particle Q𝑄Qitalic_Q-cumulant method. In this method, the two-particle azimuthal correlations are expressed in terms of a Q𝑄Qitalic_Q-vector as follows [74, 76, 75],

Qn=j=1Meinϕj.subscript𝑄𝑛superscriptsubscript𝑗1𝑀superscript𝑒𝑖𝑛subscriptitalic-ϕ𝑗Q_{n}=\sum_{j=1}^{M}e^{in\phi_{j}}.italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . (5)

Here, M𝑀Mitalic_M is the multiplicity of the event, and ϕjsubscriptitalic-ϕ𝑗\phi_{j}italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the azimuthal angle of j𝑗jitalic_jth hadron. The index ‘j𝑗jitalic_j’ runs over all the charged hadrons. The single-event average two-particle azimuthal correlations are estimated using the following expression,

2delimited-⟨⟩2\displaystyle\langle 2\rangle⟨ 2 ⟩ =|Qn|2MM(M1).absentsuperscriptsubscript𝑄𝑛2𝑀𝑀𝑀1\displaystyle=\frac{|Q_{n}|^{2}-M}{M(M-1)}.= divide start_ARG | italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_M end_ARG start_ARG italic_M ( italic_M - 1 ) end_ARG . (6)

Using Eq. (6), one can estimate the two-particle cumulants as follows,

cn{2}subscript𝑐𝑛2\displaystyle c_{n}\{2\}italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } =2=i=1Nev(W2)i2ii=1Nev(W2)i.absentdelimited-⟨⟩delimited-⟨⟩2superscriptsubscript𝑖1subscript𝑁evsubscriptsubscript𝑊delimited-⟨⟩2𝑖subscriptdelimited-⟨⟩2𝑖superscriptsubscript𝑖1subscript𝑁evsubscriptsubscript𝑊delimited-⟨⟩2𝑖\displaystyle=\langle\langle 2\rangle\rangle=\frac{\sum_{i=1}^{N_{\rm ev}}(W_{% \langle 2\rangle})_{i}\langle 2\rangle_{i}}{\sum_{i=1}^{N_{\rm ev}}(W_{\langle 2% \rangle})_{i}}.= ⟨ ⟨ 2 ⟩ ⟩ = divide start_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ev end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_W start_POSTSUBSCRIPT ⟨ 2 ⟩ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟨ 2 ⟩ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ev end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_W start_POSTSUBSCRIPT ⟨ 2 ⟩ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG . (7)

Here, delimited-⟨⟩\langle\langle~{}\rangle\rangle⟨ ⟨ ⟩ ⟩ denotes the average over all particles over all the events. Nevsubscript𝑁evN_{\rm ev}italic_N start_POSTSUBSCRIPT roman_ev end_POSTSUBSCRIPT is the total number of events used for the calculations, and (W2)isubscriptsubscript𝑊delimited-⟨⟩2𝑖(W_{\langle 2\rangle})_{i}( italic_W start_POSTSUBSCRIPT ⟨ 2 ⟩ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the weight factor for the i𝑖iitalic_ith event which takes into account the number of different two-particle combinations in an event with multiplicity M𝑀Mitalic_M. W2subscript𝑊delimited-⟨⟩2W_{\langle 2\rangle}italic_W start_POSTSUBSCRIPT ⟨ 2 ⟩ end_POSTSUBSCRIPT can be estimated as follows,

W2=M(M1).subscript𝑊delimited-⟨⟩2𝑀𝑀1W_{\langle 2\rangle}=M(M-1).italic_W start_POSTSUBSCRIPT ⟨ 2 ⟩ end_POSTSUBSCRIPT = italic_M ( italic_M - 1 ) . (8)

One can obtain the value of event-averaged reference flow with the two-particle cumulants using the following expression,

vn{2}=cn{2}.subscript𝑣𝑛2subscript𝑐𝑛2v_{n}\{2\}=\sqrt{c_{n}\{2\}}.italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } = square-root start_ARG italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } end_ARG . (9)

However, to estimate the differential flow of the Particles Of Interest (POIs), one can define pnsubscript𝑝𝑛p_{n}italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and qnsubscript𝑞𝑛q_{n}italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT vectors with specific kinematic cuts as follows,

pnsubscript𝑝𝑛\displaystyle p_{n}italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =j=1mpeinϕj,absentsuperscriptsubscript𝑗1subscript𝑚𝑝superscript𝑒𝑖𝑛subscriptitalic-ϕ𝑗\displaystyle=\sum_{j=1}^{m_{p}}e^{in\phi_{j}},= ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (10)
qnsubscript𝑞𝑛\displaystyle q_{n}italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =j=1mqeinϕj,absentsuperscriptsubscript𝑗1subscript𝑚𝑞superscript𝑒𝑖𝑛subscriptitalic-ϕ𝑗\displaystyle=\sum_{j=1}^{m_{q}}e^{in\phi_{j}},= ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ,

where, mpsubscript𝑚𝑝m_{p}italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the total number of particles labeled as POIs, and mqsubscript𝑚𝑞m_{q}italic_m start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT is the total number of particles tagged both as reference flow particles (RFP) and POI. One can estimate the single-event averaged differential two-particle azimuthal correlation using the following expression,

2delimited-⟨⟩superscript2\displaystyle\langle 2^{{}^{\prime}}\rangle⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ =pnQnmqmpMmq.absentsubscript𝑝𝑛superscriptsubscript𝑄𝑛subscript𝑚𝑞subscript𝑚𝑝𝑀subscript𝑚𝑞\displaystyle=\frac{p_{n}Q_{n}^{*}-m_{q}}{m_{p}M-m_{q}}.= divide start_ARG italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - italic_m start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_M - italic_m start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG . (11)

Finally, the differential flow, v2(pT)subscript𝑣2subscript𝑝Tv_{2}(p_{\rm T})italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ), can be estimated using the following equation,

vn{2}(pT)subscript𝑣𝑛2subscript𝑝T\displaystyle v_{n}\{2\}(p_{\rm T})italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) =dn{2}cn{2}.absentsubscript𝑑𝑛2subscript𝑐𝑛2\displaystyle=\frac{d_{n}\{2\}}{\sqrt{c_{n}\{2\}}}.= divide start_ARG italic_d start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } end_ARG start_ARG square-root start_ARG italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } end_ARG end_ARG . (12)

Here, dn{2}subscript𝑑𝑛2d_{n}\{2\}italic_d start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } is the differential nthsuperscript𝑛thn^{\rm th}italic_n start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT-order cumulant given as,

dn{2}=2=i=1Nev(w2)i2ii=1Nev(w2)i.subscript𝑑𝑛2delimited-⟨⟩delimited-⟨⟩superscript2superscriptsubscript𝑖1subscript𝑁evsubscriptsubscript𝑤delimited-⟨⟩superscript2𝑖subscriptdelimited-⟨⟩superscript2𝑖superscriptsubscript𝑖1subscript𝑁evsubscriptsubscript𝑤delimited-⟨⟩superscript2𝑖d_{n}\{2\}=\langle\langle 2^{{}^{\prime}}\rangle\rangle=\frac{\sum_{i=1}^{N_{% \rm ev}}(w_{\langle 2^{{}^{\prime}}\rangle})_{i}\langle 2^{{}^{\prime}}\rangle% _{i}}{\sum_{i=1}^{N_{\rm ev}}(w_{\langle 2^{{}^{\prime}}\rangle})_{i}}.italic_d start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 } = ⟨ ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ ⟩ = divide start_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ev end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_w start_POSTSUBSCRIPT ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ev end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_w start_POSTSUBSCRIPT ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG . (13)

The weight factor, w2subscript𝑤delimited-⟨⟩superscript2w_{\langle 2^{{}^{\prime}}\rangle}italic_w start_POSTSUBSCRIPT ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ end_POSTSUBSCRIPT, is given by,

w2=mpMmq.subscript𝑤delimited-⟨⟩superscript2subscript𝑚𝑝𝑀subscript𝑚𝑞w_{\langle 2^{{}^{\prime}}\rangle}=m_{p}M-m_{q}.italic_w start_POSTSUBSCRIPT ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_M - italic_m start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT . (14)

Unfortunately, the vn(pT)subscript𝑣𝑛subscript𝑝Tv_{n}(p_{\rm T})italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) obtained from Eq. (12) possesses contributions from non-flow effects which can be suppressed by appropriate kinematic cuts. For example, one can introduce a pseudorapidity gap between the particles in the two-particle Q𝑄Qitalic_Q-cumulant method [76]. Consequently, the whole event is divided into two sub-events, A𝐴Aitalic_A and B𝐵Bitalic_B, which are separated by a |Δη|Δ𝜂|\Delta\eta|| roman_Δ italic_η | gap. This modifies Eq. (6) as,

2Δη=QnAQnBMAMB.subscriptdelimited-⟨⟩2Δ𝜂superscriptsubscript𝑄𝑛𝐴superscriptsubscript𝑄𝑛𝐵subscript𝑀𝐴subscript𝑀𝐵\langle 2\rangle_{\Delta\eta}=\frac{Q_{n}^{A}\cdot Q_{n}^{B*}}{M_{A}\cdot M_{B% }}.⟨ 2 ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT = divide start_ARG italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ⋅ italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ⋅ italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG . (15)

Here, QnAsuperscriptsubscript𝑄𝑛𝐴Q_{n}^{A}italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT and QnBsuperscriptsubscript𝑄𝑛𝐵Q_{n}^{B}italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT are the flow vectors from sub-events A𝐴Aitalic_A and B𝐵Bitalic_B, respectively. MAsubscript𝑀𝐴M_{A}italic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT and MBsubscript𝑀𝐵M_{B}italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT are the multiplicities corresponding to the sub-events A𝐴Aitalic_A and B𝐵Bitalic_B, respectively.

The two-particle Q𝑄Qitalic_Q-cumulant with a |Δη|Δ𝜂|\Delta\eta|| roman_Δ italic_η | gap is given by,

cn{2,|Δη|}=2Δηsubscript𝑐𝑛2Δ𝜂subscriptdelimited-⟨⟩delimited-⟨⟩2Δ𝜂c_{n}\{2,|\Delta\eta|\}=\langle\langle 2\rangle\rangle_{\Delta\eta}italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 , | roman_Δ italic_η | } = ⟨ ⟨ 2 ⟩ ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT (16)

where, one can estimate 2Δηsubscriptdelimited-⟨⟩delimited-⟨⟩2Δ𝜂\langle\langle 2\rangle\rangle_{\Delta\eta}⟨ ⟨ 2 ⟩ ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT by using 2Δηsubscriptdelimited-⟨⟩2Δ𝜂\langle 2\rangle_{\Delta\eta}⟨ 2 ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT from Eq. (15) with MAMBsubscript𝑀𝐴subscript𝑀𝐵M_{A}\cdot M_{B}italic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ⋅ italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT as the event weights. If we select RFP from one sub-event and POIs from another to estimate the differential flow with a pseudorapidity gap, there is no overlap between POIs and RFP. This modifies Eq. (11) as,

2Δη=pn,AQn,Bmp,AMB.subscriptdelimited-⟨⟩superscript2Δ𝜂subscript𝑝𝑛𝐴superscriptsubscript𝑄𝑛𝐵subscript𝑚𝑝𝐴subscript𝑀𝐵\langle 2^{{}^{\prime}}\rangle_{\Delta\eta}=\frac{p_{n,A}Q_{n,B}^{*}}{m_{p,A}M% _{B}}.⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT = divide start_ARG italic_p start_POSTSUBSCRIPT italic_n , italic_A end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_n , italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT italic_p , italic_A end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG . (17)

Now, one can obtain the two-particle differential cumulant by taking an event average over 2Δηsubscriptdelimited-⟨⟩superscript2Δ𝜂\langle 2^{{}^{\prime}}\rangle_{\Delta\eta}⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT from Eq. (17) with an event weight factor mp,AMBsubscript𝑚𝑝𝐴subscript𝑀𝐵m_{p,A}M_{B}italic_m start_POSTSUBSCRIPT italic_p , italic_A end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. Thus, the two-particle differential cumulant is given by,

dn{2,|Δη|}=2Δη.subscript𝑑𝑛2Δ𝜂subscriptdelimited-⟨⟩delimited-⟨⟩superscript2Δ𝜂d_{n}\{2,|\Delta\eta|\}=\langle\langle 2^{{}^{\prime}}\rangle\rangle_{\Delta% \eta}.italic_d start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 , | roman_Δ italic_η | } = ⟨ ⟨ 2 start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ ⟩ start_POSTSUBSCRIPT roman_Δ italic_η end_POSTSUBSCRIPT . (18)

Finally, the two-particle differential flow coefficient can be estimated using the following equation,

vn{2,|Δη|}(pT)=dn{2,|Δη|}cn{2,|Δη|}.subscript𝑣𝑛2Δ𝜂subscript𝑝Tsubscript𝑑𝑛2Δ𝜂subscript𝑐𝑛2Δ𝜂v_{n}\{2,|\Delta\eta|\}(p_{\rm T})=\frac{d_{n}\{2,|\Delta\eta|\}}{\sqrt{c_{n}% \{2,|\Delta\eta|\}}}.italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 , | roman_Δ italic_η | } ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) = divide start_ARG italic_d start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 , | roman_Δ italic_η | } end_ARG start_ARG square-root start_ARG italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { 2 , | roman_Δ italic_η | } end_ARG end_ARG . (19)

To estimate the anisotropic flow coefficients, the multi-particle Q𝑄Qitalic_Q-cumulants method is adopted in many experiments [76, 78, 77]. In this study, pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT differential second- and third-order flow coefficients, v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and v3subscript𝑣3v_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, are estimated using the above equations by setting the n=𝑛absentn=italic_n = 2 and 3, respectively. To estimate the anisotropic flow coefficients, we use all the charged hadrons within the pseudorapidity region, |ηlab|<2.5subscript𝜂lab2.5|\eta_{\rm lab}|<2.5| italic_η start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT | < 2.5. The RFP are charged hadrons selected within |ηlab|<2.5subscript𝜂lab2.5|\eta_{\rm lab}|<2.5| italic_η start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT | < 2.5 and 0.2<pT<5.00.2subscript𝑝T5.00.2<p_{\rm T}<5.00.2 < italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT < 5.0 GeV/c. In addition, to reduce the non-flow effects from the two-particle Q𝑄Qitalic_Q-cumulant method, we use a pseudorapidity gap, |Δη|>1.0Δ𝜂1.0|\Delta\eta|>1.0| roman_Δ italic_η | > 1.0, in the two-subevent method.

III Results and discussions

Refer to caption
Refer to caption
Figure 2: (Color online) Upper panel shows pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT-spectra of all charged particles for different nuclear density profiles in p–O (left) and p–C (right) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV for (0–5)% and (30–40)% centrality classes. The ratios of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra of SOG to α𝛼\alphaitalic_α–cluster density profile for respective centrality classes are shown in the lower panel.

In this section, we begin with a comparison of transverse momentum (pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT) and pseudorapidity (η𝜂\etaitalic_η) spectra for the final-state charged particles followed by a discussion on the participant eccentricity (ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangularity (ϵ3subscriptitalic-ϵ3\epsilon_{3}italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) in p–C and p–O collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT. Then, we discuss the centrality and pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT dependence of elliptic flow (v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangular flow (v3subscript𝑣3v_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT), and v3/v2subscript𝑣3subscript𝑣2v_{3}/v_{2}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as a function of collision centrality. Finally, the results for v2/ϵ2subscript𝑣2subscriptitalic-ϵ2v_{2}/\epsilon_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and v3/ϵ3subscript𝑣3subscriptitalic-ϵ3v_{3}/\epsilon_{3}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT as a function of collision centrality have been reported. All these results include a comparison between the SOG and α𝛼\alphaitalic_α–cluster type density profiles for the 12C and 16O nucleus.

As discussed earlier, we define the collision centrality by the geometrical slicing of the impact parameter (b𝑏bitalic_b) distributions for p–C and p–O collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV. Table 2 and  3 shows the impact parameter values against each centrality class, along with the mean number of binary collisions (Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm coll}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩) and average number of participants (Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm part}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩) for both SOG and α𝛼\alphaitalic_α–cluster type nuclear density profiles in p–O and p–C collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV, respectively. These values are obtained from the Glauber model simulation. For a given centrality class, the upper-cut on the impact parameter (bmaxsubscript𝑏maxb_{\rm max}italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT) is always smaller for the case of α𝛼\alphaitalic_α–cluster nuclear density profile compared to the SOG for both p–C and p–O collision systems. For the central p–O collisions (up to (10-20)%), Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm part}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ and Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm coll}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ are higher for the SOG density profile than the α𝛼\alphaitalic_α–cluster nuclear distribution. However, beyond (20-30)% centrality class, in mid-central and peripheral collisions, Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm part}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ and Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm coll}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ values are higher for the case of α𝛼\alphaitalic_α–cluster as compared to the SOG. In contrast, for the p–C collisions, Npartdelimited-⟨⟩subscript𝑁part\langle N_{\rm part}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ and Ncolldelimited-⟨⟩subscript𝑁coll\langle N_{\rm coll}\rangle⟨ italic_N start_POSTSUBSCRIPT roman_coll end_POSTSUBSCRIPT ⟩ values for α𝛼\alphaitalic_α–cluster is always higher than the SOG type density profile irrespective of the centrality of the collision.

Since the collision systems are asymmetric, the center-of-mass frame of p–O or p–C collisions do not coincide with the laboratory frame. Thus, the shift in rapidity of the center-of-mass frame (ycmssubscript𝑦cmsy_{\rm cms}italic_y start_POSTSUBSCRIPT roman_cms end_POSTSUBSCRIPT) from the lab frame (ylabsubscript𝑦laby_{\rm lab}italic_y start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT) needs to be included. For asymmetric collisions, the shift in rapidity (ΔyΔ𝑦\Delta yroman_Δ italic_y) can be expressed in terms of the atomic number (Z𝑍Zitalic_Z) and mass number (A𝐴Aitalic_A) of the colliding nuclei as,

Δy=|ylabycms|=12ln[Z1A2/Z2A1].Δ𝑦subscript𝑦labsubscript𝑦cms12subscript𝑍1subscript𝐴2subscript𝑍2subscript𝐴1\Delta y=|y_{\rm lab}-y_{\rm cms}|=\frac{1}{2}\ln{[Z_{1}A_{2}/Z_{2}A_{1}]}.roman_Δ italic_y = | italic_y start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT - italic_y start_POSTSUBSCRIPT roman_cms end_POSTSUBSCRIPT | = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ln [ italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] . (20)

The subscripts 1111 and 2222 denote the projectile and target nuclei, respectively. For both p–O and p–C collisions, the rapidity shift is estimated to be ΔyΔy\Delta\rm yroman_Δ roman_y = 0.346 units in the direction of the proton beam. As a result, a detector coverage of |ηlab|<0.8subscript𝜂lab0.8|\eta_{\rm lab}|<0.8| italic_η start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT | < 0.8 would imply the nucleon-nucleon c.m.s to be roughly 1.15<ηcms<0.451.15subscript𝜂cms0.45-1.15<\eta_{\rm cms}<0.45- 1.15 < italic_η start_POSTSUBSCRIPT roman_cms end_POSTSUBSCRIPT < 0.45. Table 4 shows the mean charged particle multiplicity density (dNch/dηdelimited-⟨⟩𝑑subscript𝑁ch𝑑𝜂\langle dN_{\rm ch}/d\eta\rangle⟨ italic_d italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT / italic_d italic_η ⟩) in |ηlab|<0.5subscript𝜂lab0.5|\eta_{\rm lab}|<0.5| italic_η start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT | < 0.5 as a function of collision centrality in p–O and p–C collisions for α𝛼\alphaitalic_α–clustered and SOG nuclear density profiles. One finds that dNch/dηdelimited-⟨⟩𝑑subscript𝑁ch𝑑𝜂\langle dN_{\rm ch}/d\eta\rangle⟨ italic_d italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT / italic_d italic_η ⟩ is higher for the SOG case for both p–O and p–C collisions for the central collisions, while towards mid-central and peripheral collisions, the multiplicity density is higher for the α𝛼\alphaitalic_α–cluster case.

III.1 Transverse momentum (pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT) and pseudorapidity (η𝜂\etaitalic_η) distributions

Refer to caption
Refer to caption
Figure 3: (Color online) η𝜂\etaitalic_η-spectra of charged particles (pT>0.15subscript𝑝T0.15p_{\rm T}>0.15italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT > 0.15 GeV/c𝑐citalic_c) for different nuclear density profiles in p–O (upper) and p–C (lower) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV for (0–5)% and (30–40)% centrality classes.

Figure 2 shows the event normalized pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT-spectra for unidentified charged hadrons in |ηlab|<0.8subscript𝜂lab0.8|\eta_{\rm lab}|<0.8| italic_η start_POSTSUBSCRIPT roman_lab end_POSTSUBSCRIPT | < 0.8 for central (0–5)% and mid-central (30–40)% p–O (left), p–C (right) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT. The pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra include the cases for the two nuclear density profiles considered in this study, i.e., the SOG and the α𝛼\alphaitalic_α–cluster. For the most central p–O and p–C collisions, the charged particle yield is higher for the case of SOG as compared to the α𝛼\alphaitalic_α–cluster density profile. This enhancement of yield for the SOG density profile for the most central collisions is larger in p–O collisions compared to the same centrality in p–C collisions. In contrast, an opposite trend is observed for the mid-central collisions, where the charged particle yield is enhanced for α𝛼\alphaitalic_α–cluster density profile compared to SOG. This is also consistent with the results of dNch/dηdelimited-⟨⟩𝑑subscript𝑁ch𝑑𝜂\langle dN_{\rm ch}/d\eta\rangle⟨ italic_d italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT / italic_d italic_η ⟩ shown in Table 4. On the bottom panels of Fig. 2, the ratio of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra of SOG to α𝛼\alphaitalic_α–cluster type density profile has been shown for both collision systems. The ratios show a hardening of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra in the most central collisions for the SOG density profile compared to the α𝛼\alphaitalic_α–cluster case. On the other hand, for the mid-central collisions, the SOG type density profile produces a slightly softer pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT-spectra than α𝛼\alphaitalic_α–cluster density profile.

These effects of hardening in the central and softening in the peripheral collisions for the SOG density profile than the α𝛼\alphaitalic_α–cluster density profile are significantly prominent in p–O than in p–C collisions [79, 80]. In hydrodynamics, the hardening of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra is attributed to the collective transverse flow generated due to hydrodynamic expansion [81]. As transport models like AMPT incorporate many of the experimentally measured collective features [62, 63, 64, 65, 38, 67, 61, 16, 66], a hardening of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT spectra may hint to the presence of a hydrodynamically expanding system.

Figure 3 represents the η𝜂\etaitalic_η-spectra of all charged hadrons with pT>0.15subscript𝑝T0.15p_{\rm T}>0.15italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT > 0.15 GeV/c𝑐citalic_c for (0–5)% and (30–40)% centrality classes in p–O (upper) and p–C (lower) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV for SOG and α𝛼\alphaitalic_α–clustered density profiles. One can observe a similarity in terms of collision centrality and nuclear density profile dependence of charged particle yield between Fig. 2 and 3. The yield in the most central p–O and p–C collisions is higher for the SOG nuclear density profile as compared to the clustered geometry of the nucleons. However, in the mid-central collisions, the yield is higher for α𝛼\alphaitalic_α–clustered nuclear density profile as compared to the SOG case.

The double-peaks appearing in the pseudorapidity distribution seem to be symmetric around η=0𝜂0\eta=0italic_η = 0 towards peripheral p–O and p–C collisions, which is similar to pp collisions, possibly due to their similarity in geometry [82]. In other words, due to the smaller number of participants in the mid-central and peripheral p–O and p–C collisions, the scenario becomes equivalent to pp collisions with Npart24similar-to-or-equalsdelimited-⟨⟩subscript𝑁part24\langle N_{\rm part}\rangle\simeq 2-4⟨ italic_N start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ⟩ ≃ 2 - 4. However, as we move towards the central collisions in asymmetric collision systems, the pseudorapidity distribution becomes progressively more asymmetric around η=0𝜂0\eta=0italic_η = 0, such that fewer particles are produced in the direction of the proton beam while there is a predominant emission from the participant nucleons of the heavier nucleus. Unlike A–A nuclear collisions, the scenario in p–A collisions is that of a single nucleon probing the nucleons of a target nucleus in a narrow cylinder [83]. In a central p–A collision, the projectile proton interacts with a denser volume of the target nucleus than in the case of a peripheral p–A collision. Consequently, the yield in the nucleus-beam-going direction will increase from peripheral to central collisions. As one can see in Fig. 3, the multitude of particles produced is higher in the oxygen- and carbon-going directions than in the proton-going direction for the most central p–O and p–C collisions, respectively. In addition, one observes a slightly higher asymmetry in the double-peak structure of pseudorapidity distribution in p–O collisions as compared to p–C collisions, presumably due to a denser 16O nuclei. The effects are further enhanced for an SOG system as compared to α𝛼\alphaitalic_α–cluster nuclear density profile, indicating that the core part of an SOG is slightly denser as compared to an α𝛼\alphaitalic_α–cluster nuclear geometry. This observation is consistent with our observation in Table 4.

III.2 Eccentricity and triangularity

Refer to caption
Refer to caption
Figure 4: (Color online) Centrality dependence of average eccentricity (ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩) and triangularity (ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩) for p–O (left) and p–C (right) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT for SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles are shown in the upper panels. Lower panels shows the ratio ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩/ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ as a function of centrality.

In non-central nuclear collisions, the geometry of the transverse nuclear overlap region is non-spherical (more like almond-shaped) and it has finite spatial anisotropy, which induces asymmetric pressure gradients in the medium. If the pressure gradient is strong, the spatial anisotropy in the initial state is transformed into azimuthal anisotropy in the momentum space of the final state particles through the collective expansion of the medium. The initial spatial anisotropy can be quantified in terms of eccentricity (ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and triangularity (ϵ3subscriptitalic-ϵ3\epsilon_{3}italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) of the participant nucleons. Eccentricity quantifies the extent to which the transverse overlap region is elliptical, and triangularity tells how triangular the transverse participant plane is. While eccentricity has a major contribution from the asymmetry in the collision geometry, triangularity can develop due to the event-by-event density fluctuations in the collision overlap region. In experiments, the measurement of ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and ϵ3subscriptitalic-ϵ3\epsilon_{3}italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is nontrivial. However, Monte Carlo models like AMPT can provide the coordinates of the participant nucleons. Thus, one can quantify the spatial anisotropy, ϵnsubscriptitalic-ϵn\epsilon_{\rm n}italic_ϵ start_POSTSUBSCRIPT roman_n end_POSTSUBSCRIPT, using the following expression [84],

ϵn=rncos(nϕpart)2+rnsin(nϕpart)2rn.subscriptitalic-ϵ𝑛superscriptdelimited-⟨⟩superscript𝑟𝑛𝑛subscriptitalic-ϕpart2superscriptdelimited-⟨⟩superscript𝑟𝑛𝑛subscriptitalic-ϕpart2delimited-⟨⟩superscript𝑟𝑛\epsilon_{n}=\frac{\sqrt{\langle r^{n}\cos({n\phi_{\rm part}})\rangle^{2}+% \langle r^{n}\sin({n\phi_{\rm part}})\rangle^{2}}}{\langle r^{n}\rangle}.italic_ϵ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG square-root start_ARG ⟨ italic_r start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_cos ( italic_n italic_ϕ start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ) ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⟨ italic_r start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_sin ( italic_n italic_ϕ start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT ) ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG ⟨ italic_r start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ⟩ end_ARG . (21)

Here, r𝑟ritalic_r and ϕpartsubscriptitalic-ϕpart\phi_{\rm part}italic_ϕ start_POSTSUBSCRIPT roman_part end_POSTSUBSCRIPT are the polar coordinates of the participant nucleons in the transverse plane, and n𝑛nitalic_n denotes the order of spatial anisotropy, i.e., ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and ϵ3subscriptitalic-ϵ3\epsilon_{3}italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT correspond to the second and third order spatial anisotropy, respectively. In Eq. (21), delimited-⟨⟩\langle\dots\rangle⟨ … ⟩ denotes the average over all the participating nucleons in an event.

Although the initial state in A–A collisions can be characterized by the overall geometrical shape of the interaction region, the scenario becomes different as we come to p–A collision systems. In the latter, the number of participants is comparatively less than nucleus-nucleus collisions, and the system’s geometry grows sensitive to the proton’s size. In short, the eccentricity and triangularity of the source in proton-induced interactions are fluctuation-dominated, and the size of the system is dominated by the incoming proton [85].

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: (Color online) Elliptic flow (upper) and triangular flow (lower) of charged particles as a function of transverse momentum in p–O (left) and p–C (right) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV for SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles in (0-5)% and (30-40)% centrality classes.

The upper panels in Fig. 4 show the event-averaged eccentricity (ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩) and triangularity (ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩) as a function of centrality for p–O (left) and p–C (right) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV, and the ratio ϵ3/ϵ2delimited-⟨⟩subscriptitalic-ϵ3delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{3}\rangle/\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ is plotted in the lower panels. It is observed that for the SOG density profile, ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ increases linearly with an increase in centrality for both p–O and p–C collision systems. This trend is also quite similar to that of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for the Woods-Saxon density profile in O–O collisions obtained using AMPT as can be in Ref. [16]. In A–A collisions, the reason for the increasing trend of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ with centrality is a growing elliptic geometry of the collision overlap region. However, this is not applicable for p–A collisions. In p–A collisions, the contributions from increasing fluctuations or decreasing matter density with an increase in impact parameter towards the peripheral collisions can lead to a significant increase in eccentricity. On the other hand, the behavior of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ with an increase in collision centrality becomes much more interesting for the α𝛼\alphaitalic_α–cluster case. Here, ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ increases non-linearly till (20–30)% centrality and remains almost unaffected with the increase in centrality till (60–70)% centrality, and finally increases sharply for the (70–100)% centrality bin. Noticeably, this behavior of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ with an increase in collision centrality is similar for both p–O and p–C collisions for the α𝛼\alphaitalic_α–cluster case. In addition, a similar observation for ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ is made for O–O collisions with an α𝛼\alphaitalic_α–cluster nuclear density profile using AMPT [16]. This observation indicates that even if the colliding nuclear species are the same, the distribution of nucleons within the colliding nuclei does play a major role in determining the initial spatial anisotropy of the collision overlap region. Further, a clustered nuclear structure shows a similar footprint for ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ with varying centrality throughout p–O, p–C, and O–O collisions. Moreover, if we draw a comparison between the two collision systems, for a given collision centrality, the values of eccentricity are higher for p–C collisions than p–O collisions for both the density profiles. This is presumably because of a smaller system size in p–C collisions than in p–O collisions, which leads to a larger fluctuation in the participant nucleon distribution and a larger value of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for p–C collisions compared to p–O collisions.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: (Color online) Centrality dependence of elliptic flow (v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 }) (upper) and triangular flow (v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 }) (lower) in p–O (left) and p–C (right) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV for SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles using AMPT model.

The variation of ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ as a function of collision centrality appears slightly different for p–O and p–C collisions in Fig. 4. ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ increases with centrality, almost in a similar fashion for both the density profiles in p–C collisions. In contrast, in p–O collisions, ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ for α𝛼\alphaitalic_α–cluster density profile dominates over that of SOG density profile in the most central case, drops in the mid-central collisions followed by a rise in the peripheral cases. As this pattern of ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ for α𝛼\alphaitalic_α–cluster density profile in p–O collisions is similar to that of ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ in Ref. [16] for O–O collisions, this trend could be attributed specifically to the presence of an α𝛼\alphaitalic_α–clustered O16superscriptO16{}^{16}\rm Ostart_FLOATSUPERSCRIPT 16 end_FLOATSUPERSCRIPT roman_O nucleus. Additionally, the complimentary trend between ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ and ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ in the case of α𝛼\alphaitalic_α–cluster density profile for p–O system is worth noting. The ratio ϵ3/ϵ2delimited-⟨⟩subscriptitalic-ϵ3delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{3}\rangle/\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for the SOG density profile in both p–O and p–C collision systems show a very similar trend, which gradually decreases with increasing centrality and approaches unity (ϵ3ϵ2similar-to-or-equalsdelimited-⟨⟩subscriptitalic-ϵ3delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{3}\rangle\simeq\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ ≃ ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩) for mid-central to peripheral collisions. This trend indicates that the spatial anisotropy in central p–A collisions have a larger contribution from event-by-event fluctuations as compared to the shape of the nuclear overlap region since ϵ3/ϵ2>1.0delimited-⟨⟩subscriptitalic-ϵ3delimited-⟨⟩subscriptitalic-ϵ21.0\langle\epsilon_{3}\rangle/\langle\epsilon_{2}\rangle>1.0⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ > 1.0. However, in the case of an α𝛼\alphaitalic_α–cluster density profile, ϵ3/ϵ2delimited-⟨⟩subscriptitalic-ϵ3delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{3}\rangle/\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ shows a sharp increase for the most central p–O collisions, which then decreases to unity towards mid-central collisions to rise again in peripheral collisions. This is once again consistent with the results for O–O collisions for α𝛼\alphaitalic_α–cluster density profile using AMPT, as shown in Ref. [16]. This dominance of ϵ3/ϵ2delimited-⟨⟩subscriptitalic-ϵ3delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{3}\rangle/\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for the most central p–O collisions with the α𝛼\alphaitalic_α–clustered profile is absent in p–C collisions. This behavior may be attributed to the presence of an extra α𝛼\alphaitalic_α–cluster in the oxygen nucleus than in the carbon, which leads to an additional density fluctuation in the most central case for the p–O collisions resulting in a higher value of ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩.

III.3 Elliptic flow and triangular flow

Refer to caption
Refer to caption
Figure 7: (Color online) v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 }/v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } as a function of centrality in p–O (left) and p–C (right) collisions at sNNsubscript𝑠NN\sqrt{s_{\rm NN}}square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV for SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles from AMPT.

Figure 5 represents pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT dependence of elliptic flow (upper) and triangular flow (lower) for all charged particles in |η|<2.5𝜂2.5|\eta|<2.5| italic_η | < 2.5 measured using two-particle Q𝑄Qitalic_Q-cumulant method for SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles in p–O (left) and p–C (right) collisions for (0–5)% and (30–40)% centrality bins, using AMPT model. v2{2}(pT)subscript𝑣22subscript𝑝Tv_{2}\{2\}(p_{\rm T})italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) increases at low-pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT, and becomes maximum towards the intermediate-pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT in both p–O and p–C collisions for both the nuclear density profiles. For (0–5)% and (30-40)% centrality classes shown here, v2{2}(pT)subscript𝑣22subscript𝑝Tv_{2}\{2\}(p_{\rm T})italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) in both p–O and p–C collisions are estimated to have almost similar values despite having a sizeable difference in the initial eccentricity for these two centrality bins. Again, for these centrality bins, the effect of the density profiles on the estimated v2{2}(pT)subscript𝑣22subscript𝑝Tv_{2}\{2\}(p_{\rm T})italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) is also very minimal. However, a clear distinction for the case of v3{2}(pT)subscript𝑣32subscript𝑝Tv_{3}\{2\}(p_{\rm T})italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } ( italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT ) values are seen as a function of centrality and the choice of nuclear density profile.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: (Color online) Ratio v2{2}/ϵ2subscript𝑣22delimited-⟨⟩subscriptitalic-ϵ2v_{2}\{2\}/\langle\epsilon_{2}\rangleitalic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ (upper) and v3{2}/ϵ3subscript𝑣32delimited-⟨⟩subscriptitalic-ϵ3v_{3}\{2\}/\langle\epsilon_{3}\rangleitalic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ (lower) as a function of centrality in p–O (left) and p–C (right) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT for both SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles.

Figure 6 shows the centrality dependence of elliptic flow (upper) and triangular flow (lower) using the two-particle Q𝑄Qitalic_Q-cumulant method for SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles in p–O (left) and p–C (right) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT. Interestingly, the SOG nuclear density profile shows a similar centrality dependence of v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } for both p–O and p–C collisions. Here, v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } rises from (0–5)% to (5–10)% centrality class followed by a smooth drop towards mid-central and peripheral collisions. However, one finds that v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } is slightly higher for p–C collisions as compared to p–O collisions which can be attributed to the higher value of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for the p–C collisions as seen in Fig. 4.

Further, one observes a similar centrality dependence of v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } between p–O and p–C collisions with the α𝛼\alphaitalic_α–cluster nuclear density profile. However, in p–O collisions, v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } for the α𝛼\alphaitalic_α–cluster nuclear density profile is smaller in magnitude compared to the same in the SOG type nuclear density, excluding (05-20)% centrality class. Again, in p–O collisions, the crossing of v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } between α𝛼\alphaitalic_α–cluster and SOG density profiles within (0–5)% and (5–10)% centrality class is consistent with the crossing of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for both nuclear density profiles seen in Fig. 4. However, towards the mid-central and peripheral collisions, due to a smaller number of participants, the centrality dependence of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ might not be well reflected in the corresponding values of v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 }. In p–C collisions, v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } for the α𝛼\alphaitalic_α–cluster case is always higher than the corresponding values of v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } for the SOG case. This is consistent with corresponding ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ as shown in Fig. 4 till (20-30)% central collisions, but, for mid-central to peripheral collisions, the centrality dependence of ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ is not seen to be well reflected in the corresponding values of v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 }.

Moving to the centrality dependence of triangular flow (v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 }), as shown in the lower plots of Fig. 6, for both α𝛼\alphaitalic_α–cluster and SOG nuclear density profiles, v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } is found to decrease with an increase in centrality for both p–O and p–C collisions; however, the fall is found to be less steep for collisions involving α𝛼\alphaitalic_α–clusters. Interestingly, one observes that while ϵ3delimited-⟨⟩subscriptitalic-ϵ3\langle\epsilon_{3}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ shows an increasing trend towards the peripheral collisions, v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } shows a decreasing trend, which is expected in small collision systems like p–O and p–C collisions. Further, one may notice that v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } varies with a factor of maximum 8% between its lowest and highest values in both the nuclear profiles and in both p–O and p–C collision systems. This weak dependence of elliptic flow coefficients is consistent with the experimental observation of elliptic flow in p–Pb [26], and p–Au [87, 86, 88, 89] collisions. In contrast, for v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 }, one finds that the variation is more than 50% for SOG nuclear density profiles, while an α𝛼\alphaitalic_α–cluster density profile shows a small centrality dependence in both p–O and p–C collisions.

Figure 7 shows the centrality dependence of ratio v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 }/v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } for SOG and α𝛼\alphaitalic_α–cluster density profiles for p–O and p–C collisions. As one can see, the ratio is strictly less than one, which is consistent with the observations made in Pb–Pb and p-Pb collisions in experiments [26]. We observe that the trend for v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 }/v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } is very much identical to the trend for v3delimited-⟨⟩subscript𝑣3\langle v_{3}\rangle⟨ italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩, where a gradually decreasing ratio of SOG density profile is cut across in the mid-centralities by an almost flat curve of α𝛼\alphaitalic_α–cluster density profile, for both the collision systems. This is because v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } has a smaller centrality dependence, whereas v3{2}subscript𝑣32v_{3}\{2\}italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } is affected significantly with a change in collision centrality.

The ratio of vn/ϵnsubscript𝑣𝑛subscriptitalic-ϵ𝑛v_{n}/\epsilon_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT / italic_ϵ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT can be considered as the hydrodynamic medium response to the initial spatial anisotropy. Figure 8 shows the centrality dependence of ratio v2{2}/ϵ2subscript𝑣22delimited-⟨⟩subscriptitalic-ϵ2v_{2}\{2\}/\langle\epsilon_{2}\rangleitalic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ (upper) and v3{2}/ϵ3subscript𝑣32delimited-⟨⟩subscriptitalic-ϵ3v_{3}\{2\}/\langle\epsilon_{3}\rangleitalic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ (lower) in p–O (left) and p–C (right) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT for both SOG and α𝛼\alphaitalic_α–cluster nuclear density profiles. Naively, one finds that the centrality dependence of v2{2}/ϵ2subscript𝑣22delimited-⟨⟩subscriptitalic-ϵ2v_{2}\{2\}/\langle\epsilon_{2}\rangleitalic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ is nearly proportional to 1/ϵ21delimited-⟨⟩subscriptitalic-ϵ21/\langle\epsilon_{2}\rangle1 / ⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ for both the nuclear density profiles in both p–O and p–C collisions. This is because v2{2}subscript𝑣22v_{2}\{2\}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 2 } has a weak dependence on collision centrality (Fig. 6), whereas ϵ2delimited-⟨⟩subscriptitalic-ϵ2\langle\epsilon_{2}\rangle⟨ italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ shows a relatively stronger centrality dependence, shown in Fig. 4. In contrast, we find an interesting observation of the centrality dependence of v3{2}/ϵ3subscript𝑣32delimited-⟨⟩subscriptitalic-ϵ3v_{3}\{2\}/\langle\epsilon_{3}\rangleitalic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩. The SOG nuclear density profiles show a decreasing behavior from central to mid-central collisions, indicating a significant impact of the system formed in the evolution of the anisotropic flow coefficient. Interestingly, for the α𝛼\alphaitalic_α–cluster case, the ratio v3{2}/ϵ3subscript𝑣32delimited-⟨⟩subscriptitalic-ϵ3v_{3}\{2\}/\langle\epsilon_{3}\rangleitalic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ is less sensitive to collision centrality in both p–O and p–C collisions. The sensitivity is even less for p–O collisions than for p–C collisions. This marks one of the significant observations in p–O and p–C collisions, which relates to the presence of a clustered geometry in 16O and 12C nuclei.

IV Summary

In summary, for the first time, we report a systematic study of the effect of the nuclear density profile through p–O and p–C collisions at the LHC. We study the transverse momentum (pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT) and pseudorapidity (η𝜂\etaitalic_η) spectra, eccentricity and triangularity, elliptic and triangular flow in p–O and p–C collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV using AMPT with α𝛼\alphaitalic_α–cluster type geometry and a model-independent Sum of Gaussians (SOG) type nuclear density profiles.

With the study of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT and η𝜂\etaitalic_η spectra, the study establishes the differences in the yield for both the nuclear density profiles for different centrality classes. In addition, one finds a significant ϵnsubscriptitalic-ϵ𝑛\epsilon_{n}italic_ϵ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT dependence on the collision centrality, where an α𝛼\alphaitalic_α–cluster nuclear density profile maintains a similar qualitative behavior with an increase in collision centrality in p–O, p–C and O–O collisions. This is one of the important findings of this paper; although it can not be confronted in experiments, it retains significant importance in understanding the collision overlap region for an α𝛼\alphaitalic_α–clustered geometry. Further, p–Pb collisions and the α𝛼\alphaitalic_α–cluster nuclear density profile of p–O and p–C collisions show a weak dependence of triangular flow on collision centrality. In contrast, the triangular flow for the SOG density profile shows a strong centrality dependence for both p–O and p–C collision systems. Finally, for an α𝛼\alphaitalic_α–cluster nuclear density profile, v3{2}/ϵ3subscript𝑣32delimited-⟨⟩subscriptitalic-ϵ3v_{3}\{2\}/\langle\epsilon_{3}\rangleitalic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT { 2 } / ⟨ italic_ϵ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ has a small dependence on collision centrality where SOG nuclear profile shows a linear drop with an increase in collision centrality. This is one of the important findings of this paper and can be confronted in experiments in the upcoming p–O collisions.

Refer to caption
Refer to caption
Figure 9: (Color online) Upper panel shows comparison of transverse momentum spectra measured at ALICE [90] for minimum bias p–Pb collisions at sNN=5.02subscript𝑠NN5.02\sqrt{s_{\rm NN}}=5.02square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 5.02 TeV with the predictions from AMPT with σggsubscript𝜎gg\sigma_{\rm gg}italic_σ start_POSTSUBSCRIPT roman_gg end_POSTSUBSCRIPT = 3 mb for p–Pb collisions at sNN=5.02subscript𝑠NN5.02\sqrt{s_{\rm NN}}=5.02square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 5.02 TeV as well as p–O (left) and p–C (right) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV. The lower panel shows the ratio of pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT-spectra of p–Pb, p–O and p–C collisions in AMPT to ALICE measurements for p–Pb collisions.

Acknowledgement

A.M.K.R. acknowledges the doctoral fellowships from the DST INSPIRE program of the Government of India. S.P. acknowledges the University Grants Commission (UGC), Government of India. The authors gratefully acknowledge the DAE-DST, Government of India funding under the mega-science project “Indian participation in the ALICE experiment at CERN” bearing Project No. SR/MF/PS-02/2021-IITI(E-37123).

Appendix

Figure 9 shows the comparison of transverse momentum spectra in minimum bias p–Pb collisions from AMPT with the ALICE measurements [90] at sNN=5.02subscript𝑠NN5.02\sqrt{s_{\rm NN}}=5.02square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 5.02 TeV. In addition, we compare the transverse momentum spectra of minimum bias p–O (left) and p–C (right) collisions at sNN=9.9subscript𝑠NN9.9\sqrt{s_{\rm NN}}=9.9square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 9.9 TeV. Due to a lack of experimental results for p–O and p–C collisions, we have first tuned AMPT for p–Pb collisions to the best possible match with minimum bias pTsubscript𝑝Tp_{\rm T}italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT-spectra from ALICE measurements for p–Pb collisions at sNN=5.02subscript𝑠NN5.02\sqrt{s_{\rm NN}}=5.02square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 5.02 TeV. Further, we keep similar tuning of AMPT, as discussed in Section II, to simulate p–O and p–C collisions. One finds a fair agreement between AMPT and ALICE results for p–Pb collisions at sNN=5.02subscript𝑠NN5.02\sqrt{s_{\rm NN}}=5.02square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 5.02 TeV, at pT>3subscript𝑝T3p_{\rm T}>3italic_p start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT > 3 GeV/c. Similarly, the qualitative trends for p–O and p–C collisions from AMPT for both the nuclear density profiles match quite well with AMPT results for p–Pb collisions at sNN=5.02subscript𝑠NN5.02\sqrt{s_{\rm NN}}=5.02square-root start_ARG italic_s start_POSTSUBSCRIPT roman_NN end_POSTSUBSCRIPT end_ARG = 5.02 TeV.

Refer to caption
Figure 10: (Color Online) Probability distribution of the radial position of the nucleons inside 12C and 16O nucleus considering SOG and α𝛼\alphaitalic_α-cluster nuclear density profiles.

Figure 10 shows the probability distribution of the radial position of the nucleons inside 12C and 16O nucleus. The nucleon probability distribution considering an α𝛼\alphaitalic_α-cluster nuclear distribution is compared with the SOG nuclear distribution for both 12C and 16O nucleus where the α𝛼\alphaitalic_α-cluster nuclear distribution retains a more compact structure as compared to the SOG nuclear density profile. In addition, one finds that the peak positions of the probability distribution for both the nuclear density profiles are slightly lowered for the 12C case when compared with the 16O nucleus.

References

  • [1] J. Adam et al. [ALICE Collaboration], Nature Phys. 13, 535 (2017).
  • [2] B. B. Abelev et al. [ALICE Collaboration], Phys. Lett. B 726, 164 (2013).
  • [3] V. Khachatryan et al. [CMS Collaboration], Phys. Rev. Lett. 116, 172302 (2016).
  • [4] B. B. Abelev et al. [ALICE Collaboration], Phys. Lett. B 728, 25 (2014).
  • [5] J. Adam et al. [ALICE Collaboration], Phys. Lett. B 760, 720 (2016).
  • [6] V. Khachatryan et al. [CMS Collaboration], Phys. Lett. B 765, 193 (2017).
  • [7] J. Brewer, A. Mazeliauskas and W. van der Schee, [arXiv:2103.01939 [hep-ph]].
  • [8] R. Katz, C. A. G. Prado, J. Noronha-Hostler and A. A. P. Suaide, Phys. Rev. C 102, 041901 (2020).
  • [9] U. Heinz and R. Snellings, Ann. Rev. Nucl. Part. Sci. 63, 123 (2013).
  • [10] J. Y. Ollitrault, Phys. Rev. D 46, 229 (1992).
  • [11] S. A. Voloshin, A. M. Poskanzer and R. Snellings, Landolt-Bornstein 23, 293 (2010).
  • [12] S. Voloshin and Y. Zhang, Z. Phys. C 70, 665 (1996).
  • [13] J. Adams et al. [STAR Collaboration], Nucl. Phys. A 757, 102 (2005).
  • [14] G. Giacalone, J. Jia and C. Zhang, Phys. Rev. Lett. 127, 242301 (2021).
  • [15] M. R. Haque, M. Nasim and B. Mohanty, J. Phys. G 46, 085104 (2019).
  • [16] D. Behera, S. Prasad, N. Mallick and R. Sahoo, Phys. Rev. D 108, 054022 (2023).
  • [17] S. Acharya et al. [ALICE Collaboration], JHEP 10, 152 (2021).
  • [18] S. Acharya et al. [ALICE Collaboration], Phys. Lett. B 784, 82 (2018).
  • [19] A. M. Sirunyan et al. [CMS Collaboration], Phys. Rev. C 100, 044902 (2019).
  • [20] G. Aad et al. [ATLAS Collaboration], Phys. Rev. C 101, 024906 (2020).
  • [21] C. Aidala et al. [PHENIX Collaboration], Nature Phys. 15, 214 (2019).
  • [22] S. Chatrchyan et al. [CMS Collaboration], Phys. Lett. B 718, 795 (2013).
  • [23] B. Abelev et al. [ALICE Collaboration], Phys. Lett. B 719, 29 (2013).
  • [24] G. Aad et al. [ATLAS Collaboration], Phys. Rev. Lett. 110, 182302 (2013).
  • [25] S. Chatrchyan et al. [CMS Collaboration], Phys. Lett. B 724, 213 (2013).
  • [26] S. Acharya et al. [ALICE Collaboration], Phys. Rev. Lett. 123, 142301 (2019).
  • [27] B. B. Abelev et al. [ALICE Collaboration], Phys. Rev. C 90, 054901 (2014).
  • [28] S. Y. Tang, L. Zheng, X. M. Zhang and R. Z. Wan, Nucl. Sci. Tech. 35, 32 (2024).
  • [29] G. Gamow, Constitution of Atomic Nuclei and Radioactivity, International series of monographs on physics PCMI collection, Clarendon Press (1931).
  • [30] J. A. Wheeler, Phys. Rev. 52, 1083 (1937).
  • [31] R. Bijker and F. Iachello, Phys. Rev. Lett. 112, 152501 (2014).
  • [32] X. B. Wang, G. X. Dong, Z. C. Gao, Y. S. Chen and C. W. Shen, Phys. Lett. B 790, 498 (2019).
  • [33] W. B. He, Y. G. Ma, X. G. Cao, X. Z. Cai and G. Q. Zhang, Phys. Rev. Lett. 113, 032506 (2014).
  • [34] J. He, W. B. He, Y. G. Ma and S. Zhang, Phys. Rev. C 104, 044902 (2021).
  • [35] M. Rybczyński and W. Broniowski, Phys. Rev. C 100, 064912 (2019).
  • [36] M. D. Sievert and J. Noronha-Hostler, Phys. Rev. C 100, 024904 (2019).
  • [37] S. Huang, Z. Chen, J. Jia and W. Li, Phys. Rev. C 101, 021901 (2020).
  • [38] D. Behera, N. Mallick, S. Tripathy, S. Prasad, A. N. Mishra and R. Sahoo, Eur. Phys. J. A 58, 175 (2022).
  • [39] Y. A. Li, S. Zhang and Y. G. Ma, Phys. Rev. C 102, 054907 (2020).
  • [40] P. Bozek, W. Broniowski, E. Ruiz Arriola and M. Rybczynski, Phys. Rev. C 90, 064902 (2014).
  • [41] W. Broniowski and E. Ruiz Arriola, Phys. Rev. Lett. 112, 112501 (2014).
  • [42] S. H. Lim, J. Carlson, C. Loizides, D. Lonardoni, J. E. Lynn, J. L. Nagle, J. D. Orjuela Koop and J. Ouellette, Phys. Rev. C 99, 044904 (2019).
  • [43] N. Summerfield, B. N. Lu, C. Plumberg, D. Lee, J. Noronha-Hostler and A. Timmins, Phys. Rev. C 104, L041901 (2021).
  • [44] B. Schenke, C. Shen and P. Tribedy, Phys. Rev. C 102, 044905 (2020).
  • [45] A. Huss, A. Kurkela, A. Mazeliauskas, R. Paatelainen, W. van der Schee and U. A. Wiedemann, Phys. Rev. C 103, 054903 (2021).
  • [46] B. G. Zakharov, JHEP 09, 087 (2021).
  • [47] D. Behera, S. Deb, C. R. Singh and R. Sahoo, Phys. Rev. C 109, 014902 (2024).
  • [48] C. Ding, L. G. Pang, S. Zhang and Y. G. Ma, Chin. Phys. C 47, 024105 (2023).
  • [49] Y. Z. Wang, S. Zhang and Y. G. Ma, Phys. Lett. B 831, 137198 (2022).
  • [50] M. Rybczyński, M. Piotrowska and W. Broniowski, Phys. Rev. C 97, 034912 (2018).
  • [51] A. Svetlichnyi, S. Savenkov, R. Nepeivoda and I. Pshenichnov, MDPI Physics 5, 381 (2023).
  • [52] B. Zhang, C. M. Ko, B. A. Li and Z. w. Lin, Phys. Rev. C 61, 067901 (2000).
  • [53] Z. W. Lin, C. M. Ko, B. A. Li, B. Zhang and S. Pal, Phys. Rev. C 72, 064901 (2005).
  • [54] X. N. Wang and M. Gyulassy, Phys. Rev. D 44, 3501 (1991).
  • [55] B. Zhang, Comput. Phys. Commun. 109, 193 (1998).
  • [56] Y. He and Z. W. Lin, Phys. Rev. C 96, 014910 (2017).
  • [57] B. Andersson, G. Gustafson, G. Ingelman and T. Sjostrand, Phys. Rept. 97, 31 (1983).
  • [58] Z. w. Lin and C. M. Ko, Phys. Rev. C 65, 034904 (2002).
  • [59] B. Li, A. T. Sustich, B. Zhang and C. M. Ko, Int. J. Mod. Phys. E 10, 267 (2001).
  • [60] B. A. Li and C. M. Ko, Phys. Rev. C 52, 2037 (1995).
  • [61] S. Prasad, N. Mallick, S. Tripathy and R. Sahoo, Phys. Rev. D 107, 074011 (2023).
  • [62] S. K. Das, et al. Int. J. Mod. Phys. E 31, 12 (2022).
  • [63] J. Altmann, et al. Eur. Phys. J. C 84, 421 (2024).
  • [64] N. Mallick, R. Sahoo, S. Tripathy and A. Ortiz, J. Phys. G 48, 045104 (2021).
  • [65] N. Mallick, S. Tripathy and R. Sahoo, Eur. Phys. J. C 82, 524 (2022).
  • [66] N. Mallick, S. Prasad, A. N. Mishra, R. Sahoo and G. G. Barnaföldi, Phys. Rev. D 107, 094001 (2023).
  • [67] N. Mallick, S. Prasad, A. N. Mishra, R. Sahoo and G. G. Barnaföldi, Phys. Rev. D 105, 114022 (2022).
  • [68] C. Loizides, J. Kamin and D. d’Enterria, Phys. Rev. C 97, 054910 (2018). [erratum: Phys. Rev. C 99, 019901 (2019)]
  • [69] C. Loizides, et. al., https://tglaubermc.hepforge.org/.
  • [70] A. Shukla, S. Åberg and S. K. Patra, J. Phys. G: Nucl. Part. Phys. 38, 095103 (2011).
  • [71] H. de Vries, C. W. de Jager and C. de Vries, At. Data Nucl. Data Tables 36, 495 (1987).
  • [72] I. Angeli and K. P. Marinova, Atom. Data Nucl. Data Tabl. 99, 69 (2013).
  • [73] B. B. Abelev et al. [ALICE Collaboration], JHEP 06, 190 (2015).
  • [74] A. Bilandzic, R. Snellings and S. Voloshin, Phys. Rev. C 83, 044913 (2011).
  • [75] Y. Zhou, X. Zhu, P. Li and H. Song, Phys. Rev. C 91, 064908 (2015).
  • [76] Y. Zhou [ALICE Collaboration], Nucl. Phys. A 931, 949 (2014).
  • [77] M. Aaboud et al. [ATLAS Collaboration], Eur. Phys. J. C 77, 428 (2017).
  • [78] W. Li [CMS Collaboration], Nucl. Phys. A 932, 373 (2014).
  • [79] B. B. Abelev et al. [ALICE Collaboration], Eur. Phys. J. C 74, 3054 (2014).
  • [80] B. Abelev et al. [ALICE Collaboration], Phys. Rev. Lett. 110, 082302 (2013).
  • [81] P. Bozek, Phys. Rev. C 85, 014911 (2012)
  • [82] J. Q. Tao, H. B. He, H. Zheng, W. C. Zhang, X. Q. Liu, L. L. Zhu and A. Bonasera, Nucl. Sci. Tech. 34, 172 (2023).
  • [83] S. Acharya et al. [ALICE Collaboration], Phys. Lett. B 845, 137730 (2023).
  • [84] H. Petersen, G. Y. Qin, S. A. Bass and B. Muller, Phys. Rev. C 82, 041901 (2010).
  • [85] A. Bzdak, B. Schenke, P. Tribedy and R. Venugopalan, Phys. Rev. C 87, 064906 (2013).
  • [86] J. D. Orjuela Koop, A. Adare, D. McGlinchey and J. L. Nagle, Phys. Rev. C 92, 054903 (2015).
  • [87] C. Aidala et al. [PHENIX Collaboration], Phys. Rev. C 95, 034910 (2017).
  • [88] N. J. Abdulameer et al. [PHENIX Collaboration], Phys. Rev. C 107, 024907 (2023).
  • [89] U. A. Acharya et al. [PHENIX Collaboration], Phys. Rev. C 105, 024901 (2022).
  • [90] S. Acharya et al. [ALICE Collaboration], JHEP 11, 013 (2018).