Chinese Physics C Vol. 41, No. 5 (2017) 053107

On thermodynamic self-consistency of generic axiomatic-nonextensive statistics

Abdel Nasser Tawfik1,2;1) Hayam Yassin3 Eman R. Abo Elyazeed3 1Egyptian Center for Theoretical Physics (ECTP), Modern University for Technology and Information (MTI), 11571 Cairo, Egypt 2World Laboratory for Cosmology and Particle Physics (WLCAPP), 11571 Cairo, Egypt 3Physics Department, Faculty of Women for Arts, Science and Education, Ain Shams University, 11577 Cairo, Egypt

Abstract: Generic axiomatic-nonextensive statistics introduces two asymptotic properties, to each of which a scaling function is assigned. The first and second scaling properties are characterized by the exponents c and d, respectively. In the thermodynamic limit, a grand-canonical ensemble can be formulated. The thermodynamic properties of a relativistic ideal gas of hadron resonances are studied, analytically. It is found that this generic statistics satisfies the requirements of the equilibrium thermodynamics. Essential aspects of the thermodynamic self-consistency are clarified. Analytical expressions are proposed for the statistical fits of various transverse momentum distributions measured in most-central collisions at different collision energies and colliding systems. Estimations for the freezeout

temperature (Tch) and the baryon chemical potential (µb) and the exponents c and d are determined. The earlier are found compatible with the parameters deduced from Boltzmann-Gibbs (BG) statistics (extensive), while the latter refer to generic nonextensivities. The resulting equivalence class (c,d) is associated with stretched exponentials, where Lambert function reaches its asymptotic stability. In some measurements, the resulting nonextensive is linearly composed on extensive . Apart from power-scaling, the particle ratios and yields are excellent quantities to highlighting whether the particle production takes place (non)extensively. Various particle ratios and yields measured by the STAR experiment in central collisions at 200, 62.4 and 7.7 GeV are fitted with this novel approach. We found that both c and d < 1, i.e. referring to neither BG- nor Tsallis-type statistics, but to (c,d)- entropy, where Lambert functions exponentially rise. The freezeout temperature and baryon chemical potential are found comparable with the ones deduced from BG statistics (extensive). We conclude that the particle production at STAR energies is likely a nonextensive process but not necessarily BG or Tsallis type.

Keywords: Nonextensive thermodynamical consistency, Boltzmann and Fermi-Dirac statistics PACS: 05.70.Ln, 05.70.Fh,05.70.Ce DOI: 10.1088/1674-1137/41/5/053107

1 Introduction nonextensive statistics of Tsallis type [8–14]. These are based on the conjecture that replacing the Boltzmann factor by the q-exponential function of Tsallis statistics, In theories of extensive (such as Boltzmann-Gibbs) with q > 1, leads to a good agreement with the experi- and nonextensive (such as Tsallis) statistics, thermody- mental measurements at high energies. Recently, Taw- namic consistency gives a phenomenological description fik explained that this method seems to fail to assure a for various phenomena in high-energy experiments [1, 2]. full incorporation of nonextensivity because fluctuations, The earlier was used by Hagedorn [3] in the 1960s to correlations, interactions among the produced particles prove that fireballs or heavy resonances lead to a boot- besides the possible modification in the phase space of strap approach, i.e. further fireballs, which - in turn - such an interacting system are not properly taken into consist of smaller fireballs and so on. The implementa- account [15]. Again, the Tsallis distribution was widely tion of the nonextensive Tsallis statistics was introduced applied to describe the hadron production [8–14]. At in Refs. [5–7]. Assuming that the distribution func- high transverse momentum spectra (pT ), some authors tion can vary, due to possible symmetrical change, Taw- did not obtain a power-law, while at low pT , they ob- fik applied nonextensive concepts to high-energy parti- tained an approximate exponential distribution. cle production [4]. Recently, various papers were quite In light of such a wide discrepancy, and especially successful in explaining various aspects of high-energy to find a unified statistical description for high-energy particle production using thermodynamically consistent collisions, we introduce generic axiomatic-nonextensive

Received 21 December 2016, Revised 23 December 2016 1) E-mail: a.tawfi[email protected] ©2017 Chinese Physical Society and the Institute of High Energy Physics of the Chinese Academy of Sciences and the Institute of Modern Physics of the Chinese Academy of Sciences and IOP Publishing Ltd

053107-1 Chinese Physics C Vol. 41, No. 5 (2017) 053107 statistics, in which the phase space determines the degree and formulated for further implications in high-energy of (non)extensivity. The latter is not necessarily limited physics in Refs. [15, 27]. to extensive and/or intensive thermodynamic quantities, In thermal equilibrium, is a such as temperature and baryon chemical potential. thermodynamically self-consistent theory. It fulfils the Regarding the success of Tsallis statistics in describ- requirements of equilibrium thermodynamics, because a ing transverse momentum spectra, especially at high pT , thermodynamic potential - in its thermodynamic limit we first recall that Bialas claimed that such a good fit - can be expressed as a first-order homogeneous func- would be incomplete [16]. First, it is believed to ig- tion of extensive variables [28]. The entropy is a funda- nore the contradiction between the applicability of such mental check for thermodynamic self-consistency. For a statistical thermal approach at high energy and per- a nonextensive quantum gas, the entropy should be turbative QCD. The crucial question which should be fully constructed from boson and fermion contributions: FD BE FD BE answered is how the statistical thermal approach, which Sq = Sq +Sq , where Sq and Sq are Fermi-Dirac and describes lattice thermodynamics (non-perturbative) [1] Bose-Einstein nonextensive entropy, respectively [29]. In well, can be assumed to do such an excellent job at microcanonical [30], canonical [31], and grand canoni- ultra-relativistic energies. Second, such a statistical fit cal [32] ensembles, the thermodynamic self-consistency seems to disbelieve the role of statistical cluster decay, of Tsallis statistics has been proved. If its entropic vari- which can be scaled as power laws very similar to that able is extensive, it has been demonstrated that the ho- from Tsallis statistics. This simply means that the de- mogeneity of the thermodynamic potential leads to the cay of statistical clusters is conjectured to be capable zeroth law of thermodynamics, i.e. the additivity princi- of explaining the excellent reproduction of the measured ple and Euler theorem. Also, the first and second laws of transverse momenta rather than the Tsallis-type nonex- thermodynamics should be fulfilled. An additional ingre- tensivity [15]. dient of special importance in that particle production, Also, Bialas [16] presented within the statisti- namely the fireball self-consistency principle, should be cal cluster-decay model a numerical analysis for the guaranteed as well. hadronization processes. It was found that the result- This paper is organized as follows. Generic ing transverse-momentum distribution can be a Tsallis- axiomatic-nonextensive statistics is reviewed in Section like one. Only in a very special case, where the fluctua- 2. Thermodynamic self-consistency is discussed in Sec- tions of the Lorentz factor and the temperature are given tion 3. This is divided into nonextensive Boltzmann- by Beta and Gamma distributions, respectively, can the Gibbs statistics (Section 3.1) and generic axiomatic- well-known Tsallis distribution be obtained. The origin nonextensive quantum statistics (Section 3.2). Sections of these fluctuations was introduced in Ref. [17]. Bialas 4 and 4.2 are devoted to the fitting of the transverse mo- explained the produced hadronic cluster decays purely mentum distributions and particle ratios with different thermally, i.e. following Boltzmann-Gibbs (BG) statis- beam energies and different system sizes, respectively. tics [16]. Furthermore, it is also supposed that the pro- Section 5 gives the final conclusions. duced hadronic clusters move with a fluctuating Lorentz factor in the transverse direction, i.e. a power law. Thus, 2 Generic axiomatic-nonextensive statis- the production and the decay of such clusters would be tics be regarded as examples of superstatistics. The latter can be understood as a kind of a superposition of two Based on Hanel-Thurner entropy [25–27], which is different types of statistics corresponding to nonequilib- fully expressed in Ref. [15], rium systems [18, 19]. In other words, it was concluded [17] that even superstatistics should be based on more Ω S [p] = Γ (d+1,1 clogp ) p , (1) general distributions than the Gamma type before be- c,d A − i −B i i=1 ing applied to multiparticle productions in high energy X collisions. For the sake of completeness, we recall that where pi is the probability of the i-th state and Γ (a,b) = the various power-law distributions have been already ∞ − implemented in pp-collisions [20–24]. dtta 1 exp( t) is an incomplete gamma-function b − The crucial questions that remains unanswered are Zwith and being arbitrary parameters, the generic what the origin and degree of nonextensivity are, and axiomatic-nonextensivA B e partition function for statistical how the degree of nonextensivity can be determined in processes in high-energy physics was suggested [27]. For a strongly correlated system, such as relativistic heavy- a classical gas, ion collisions. In the present work, we analyse the ther- modynamic self-consistency of the generic axiomatic- NM|B ∞ d3 p nonextensive approach, which was introduced in [25, 26] ln Z (T,µ) = V g ε (x ), (2) cl i (2π)3 c,d,r i i 0 X Z

053107-2 Chinese Physics C Vol. 41, No. 5 (2017) 053107

where V is the fireball volume and xi = β[µi Ei(p)] Then, we determine the thermodynamic properties of 2 2 1/2 − with Ei(p) = (p +mi ) is the dispersion relation of the that system and confirm that both laws of thermody- i-th state (particle). Straightforwardly, the quantum gas namics are verified. partition function reads The first law of thermodynamics describes the change in energy (dE) in terms of a change in volume (dV ) and N M|B ∞ d3 p entropy (dS): ln Z (T ) = V g ln[1 ε (x )], (3) FB  i (2π)3  c,d,r i i Z0 dE(V,S) = P dV +T dS, (7) X − where represent fermions (subscript F) and bosons where T and P are temperature and pressure coeffi- (subscript B), respectively. The distribution function cients, respectively. In the variation of the free energy, ε (x ) is given as [25, 26] c,d,r i F (V,T ) = E T S, the dependence on the entropy as − 1 given in Eq. (7) is to be replaced by a temperature- d x d εc,d,r(x) = exp − Wk B 1 Wk[B] , dependence 1 c − r −      −   (4) dF (V,T ) = dE T dS S dT P dV S dT. (8) where Wk is the Lambert W-function which has real so- − − ≡ − − lutions at k = 0 with d > 0 and at k = 1 with d < 0, If the system of interest contains a conserved number (N), it is necessary to introduce a chemical potential (µ). (1 c)r (1 c)r B = − exp − , (5) Equations (7) and (8) should be extended as 1 (1 c)r 1 (1 c)r − −  − −  − dE(V,S,N) = P dV +T dS +µdN, (9) with r = [1 c+cd] 1 and c,d being two constants to be − − dF (V,T,µ) = P dV S dT N dµ. (10) elaborated shortly. Equation (4) is valid for both classi- − − − cal and quantum gases. The equivalent classes (c,d) stand for two expo- The pressure is derived as nents giving estimations for two scaling functions with ∂F two asymptotic properties [25, 26]. Statistical systems p = . (11) − ∂V in their large size limit violating the fourth Shannon- T,µ

Khinchin axiom are characterized by a unique pair of From Eqs. (2), (3) and (11), thermo dynamics relations scaling exponents (c,d). Such systems have two asymp- can be deduced [33]. Their justification proves the ther- totic properties of their associated generalized entropies. modynamic self-consistency, Both properties are associated with one scaling function each. Each scaling function is characterized by one expo- ∂p ∂ ∂p ∂ n = , T = , s = , µ = . (12) nent; c for the first and d for the second property. These ∂µ ∂s ∂T ∂n T n µ s exponents define equivalence relations of entropic forms, i.e. two entropic forms are equivalent if their exponents To verify the second la w of thermo dynamics one has to are the same. prove that ∂s > 0. The various thermodynamic observables such as pres- In the section that follows, all these thermody- sure (p), number (n = N/V ), energy ( = E/V ) and en- namic quantities shall be derived from the generic tropy density (s = S/V ) [33] can be derived from Eq. (2) axiomatic-nonextensive statistics for Boltzmann-Gibbs, or Eq. (3). Bose-Einstein and Fermi-Dirac statistical ensembles [32]. ∂ lnZ ∂p 3.1 Nonextensive Boltzmann-Gibbs statistics p = T , n = , ∂V ∂µ As discussed in Ref. [34], the universality class (c,d) is conjectured not only to characterize the entropy of T 2 ∂ lnZ µT ∂ lnZ ∂p the system of interest entirely, but also to specify the  = + , s = . (6) V ∂T V ∂µ ∂T distribution functions of that system in the thermody- namic limit. Thus, it is likely able to determine the 3 Thermodynamic self-consistency (non)extensivity of the system. For instance, if (c,d) = (1,1), the system can be well described by BG statis- Here, we examine the thermodynamic self- tics, while if (c,d) = (q,0), the system has a Tsallis- consistency of generic axiomatic-nonextensive statistics type nonextensivity. Furthermore, if (c,d) = (1,d), in Boltzmann-Gibbs and quantum gases. The procedure stretched exponentials characterize that system. To our goes as follows. We start with the first and second laws knowledge, further details about the physical meaning of thermodynamics which control the system of interest. of (c,d) are being worked out by many colleagues and

053107-3 Chinese Physics C Vol. 41, No. 5 (2017) 053107

1 d should be published in the near future. The last case, 2 x ∞ p (x T ) W B 1 for instance, requires that d > 0 and c 1 so that g 0 − r   p 1/d → 2   1 d . limc→1 εc,d,r(x) = exp( dr[1 x/r] 1). In light of −2π T d − − − Z0 (1 c)[r x] 1+W B 1 x this, for a single particle, the Boltzmann distribution in − − 0 − r      a nonextensive system, Eq. (2), can be expressed as (17) 1 f(x) = , (13) In an ideal gas approach, such as the hadron resonance εc,d,r(x) gas (HRG) model, the thermodynamic quantities get contributions from each hadron resonance. Thus, a sum- from which various thermodynamic quantities can de- mation over the different hadron resonances of which the duced statistical ensemble consists should be added in front of the right-hand side. gT ∞ p = p2 ln[ε (x)]dp, (14) To check and satisfy the thermodynamic self- π2 c,d,r 2 0 consistency, let us rewrite Eq. (17) as Z 1 d 2 x ∞ p W B 1 p  µ n g 0 − r s = + , (18) T T T n =     1 dp, − 2π2 x d Z0 (1 c)[r x] 1+W B 1 − − 0 − r and take its derivative with respect to . We then get      (15) ∂s 1 1 = . (19) x d 2 ∂ n T ∞ p (x T ) W0 B 1 g − r  = −     1 dp 2π2 x d Furthermore, we derive from Eq. (2) [33], Z0 (1 c)[r x] 1+W B 1 − − 0 − r    p = +T s+µn. (20)  1  d − 2 x ∞ p W B 1 g µ 0 − r At constant T , the derivative of pressure with respect to +   1 dp, π2   d µ reads 2 0 x Z (1 c)[r x] 1+W0 B 1 ∂p ∂ ∂s ∂n − − − r = +T +n+µ . (21)      ∂µ −∂µ ∂µ ∂µ (16) T

g ∞ The differentiations of Eqs. (15), (16), and (17) with s = p2 ln[ε (x)]dp 2π2 c,d,r respect to µ lead to Z0

1 2 d 2 x ∞ p W B 1 ∂n g 0 − r =   dp π2   1 3 ∂µ 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p W B 1 g 0 − r   dp π2   1 2 −2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p W B 1 g 0 − r +     1 dp, (22) 2π2T x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 2 d 2 x ∞ p W B 1 ∂ g µ 0 − r =    dp π2   1 3 ∂µ 2 T x d Z0 (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r     

053107-4 Chinese Physics C Vol. 41, No. 5 (2017) 053107

1 d 2 x ∞ p W B 1 g µ 0 − r   dp π2   1 2 −2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p W B 1 g µ 0 − r +     1 dp 2π2T x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 2 d 2 x ∞ p (x T ) W B 1 g 0 − r   dp π2   1 3 −2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p (x T ) W B 1 g 0 − r +   dp π2   1 2 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p (x T ) W B 1 g 0 − r     1 dp, (23) −2π2T x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 2 d 2 x ∞ p (x T ) W B 1 ∂s g 0 − r = −   dp π2 2   1 3 ∂µ 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p (x T ) W B 1 g 0 − r +   dp π2 2   1 2 2 T x d Z0 (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r      1 d 2 x ∞ p (x T ) W B 1 g 0 − r     1 dp. (24) −2π2T 2 x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r      Then, the substitution from Eqs. (15), (22), (23) and By substituting from Eq. (25) into this previous equa- (24) into Eq. (21) gives tion, we get ∂p ∂ = n. (25) = µ, (28) ∂µ ∂n s T

Also, from Eq. (18) we can express the energy density which is a proof of thermodynamic self-consistency. as To prove the fourth equation in Eq. (12), let us differ-  = T s p+µ n, (26) − entiate Eq. (20) with respect to temperature at constant Its differentiation with respect to n is µ ∂ ∂p ∂µ ∂p ∂µ ∂µ ∂p ∂s ∂ ∂n = +µ+n = +µ+n . (27) = s+T +µ , (29) n n n n ∂T ∂T − ∂T ∂T ∂ s −∂ ∂ −∂µ ∂ ∂n µ

053107-5 Chinese Physics C Vol. 41, No. 5 (2017) 053107

The differentiations of Eqs. (15), (16) and (17) with respect to T read

1 2 d 2 x ∞ p (x T ) W B 1 ∂n g 0 − r = −   dp π2 2   1 3 ∂T 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r     1  d 2 x ∞ p (x T ) W B 1 g 0 − r +   dp π2 2   1 2 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r    1  d 2 x ∞ p (x T ) W B 1 g 0 − r     1 dp, (30) −2π2T 2 x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r    1 2  d 2 2 x ∞ p (x T ) W B 1 ∂ g 0 − r =   dp π2 2   1 3 ∂T 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r     1  d 2 2 x ∞ p (x T ) W B 1 g 0 − r   dp π2 2   1 2 −2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r    1  d 2 2 x ∞ p (x T ) W B 1 g 0 − r +     1 dp 2π2T 2 x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r     1 2 d 2 x ∞ p (x T ) W B 1 g µ 0 − r    dp π2 2   1 3 −2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r    1  d 2 x ∞ p (x T ) W B 1 g µ 0 − r +   dp π2 2   1 2 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r    1  d 2 x ∞ p (x T ) W B 1 g µ 0 − r     1 dp, (31) −2π2T 2 x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r    1 2  d 2 2 x ∞ p (x T ) W B 1 ∂s g 0 − r =   dp π2 3   1 3 ∂T 2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r     1  d 2 2 x ∞ p (x T ) W B 1 g 0 − r   dp π2 3   1 2 −2 T 0 x d Z (d d c)r2 (1 x/r)2 1+W B 1 − − 0 − r     

053107-6 Chinese Physics C Vol. 41, No. 5 (2017) 053107

1 d 2 2 x ∞ p (x T ) W B 1 g 0 − r +     1 dp. (32) 2π2T 3 x d Z0 (1 c)r2 (1 x/r)2 1+W B 1 − − 0 − r      Then, the substitution from Eqs. (17), (30), (31) and ∂ > µ∂n. (35) (32) into Eq. (29) leads to This inequality is fulfilled, when comparing Eq. (30) ∂p = s. (33) with Eq. (31). Two cases can be classified. Firstly, ∂T µ ∂ > µ∂n is obvious, at arbitrary µ. Secondly, ∂ = µ∂n

is also obtained, at c = 1 or µ =  or µ = +T r. It is apparent that the given definitions of temperature, number density, chemical potential and entropy density 3.2 Generic axiomatic-nonextensive quantum are thermodynamically consistent. statistics The first law of thermodynamics describes the conse- quences of heat transfer, Eq. (18), while the second law For quantum statistics, the distribution function sets constraints on it, i.e. δs > 0, reads 1 1 f(x) = , (36) ∂s [∂ µ∂n], (34) 1 ε (x) ' T −  c,d,r where ∂µ/∂T is conjectured to vanish. Therefore, the Accordingly, the various thermodynamic quantities can second law of thermodynamics is fulfilled, if be deduced

gT ∞ p = p2 ln[1 ε (x)]dp, (37) 2π2  c,d,r Z0

1 d 2 x ∞ p ε (x) W B 1 g c,d,r 0 − r n =      1 dp, (38) 2π2 x d Z0 (1 c)[1 ε (x)] (r x) 1+W B 1 −  c,d,r − 0 − r     

1 d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r  = ∓     1 dp 2π2 x d Z0 (1 c)[1 ε (x)](r x) 1+W B 1 −  c,d,r − 0 − r      1 d 2 x ∞ p ε (x) W B 1 gµ c,d,r 0 − r     1 dp, (39) 2π2 x d Z0 (1 c)[1 ε (x)] (r x) 1+W B 1 −  c,d,r − 0 − r     

1 d 2 x ∞ ∞ p εc,d,r(x) (x T ) W0 B 1 g 2 g − r s = p ln[1 εc,d,r(x)] dp     1 dp. (40) 2π2   2π2T x d Z0 Z0 (1 c)[1 ε (x)](r x) 1+W B 1 −  c,d,r − 0 − r      The differentiations of Eqs. (38), (39) and (40) with respect to µ give

1 2 d 2 x ∞ p ε (x) W B 1 ∂n g c,d,r 0 − r =    dp π2   1 3 ∂µ 2 T 0 x d Z (d d c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r     

053107-7 Chinese Physics C Vol. 41, No. 5 (2017) 053107

1 d 2 x ∞ p ε (x) W B 1 g c,d,r 0 − r   dp π2   1 2 ∓2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2   d 2 x ∞ p ζ (x) W B 1 g c,d,r 0 − r   dp π2   1 2 −2 T x d Z0 (1 c)2 [1 ε (x)]2 r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ε (x) W B 1 g c,d,r 0 − r   dp π2   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 d 2 x ∞ p ε (x) W B 1 g c,d,r 0 − r     1 dp, (41) 2π2T x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      where 1 2d x d ζ (x) = exp − W B 1 W [B] . (42) c,d,r 1 c k − r − k  −      

1 2 d 2 x ∞ p ε (x) W B 1 ∂ g µ c,d,r 0 − r =     dp π2   1 3 ∂µ 2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 d 2 x ∞ p ε (x) W B 1 g µ c,d,r 0 − r   dp π2   1 2 ∓2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ζ (x) W B 1 g µ c,d,r 0 − r   dp π2   1 2 −2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ε (x) W B 1 g µ c,d,r 0 − r   dp π2   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 d 2 x ∞ p ε (x) W B 1 g µ c,d,r 0 − r     1 dp 2π2T x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2   1 3 ∓2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r     

053107-8 Chinese Physics C Vol. 41, No. 5 (2017) 053107

1 d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2   1 2 2 T 0 x d Z (d d c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 x ∞ p ζ (x) (x T ) W B 1 g c,d,r 0 − r +   dp π2   1 2 2 T x d Z0 (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2   1 2 ∓2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1   d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r     1 dp, (43) ∓2π2T x d Z0 (1 c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 x ∞ p ε (x) (x T ) W B 1 ∂s g c,d,r 0 − r = ∓   dp π2 2   1 3 ∂µ 2 T 0 x d Z (d d c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1  d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 2   1 2 2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 x ∞ p ζ (x) (x T ) W B 1 g c,d,r 0 − r +   dp π2 2   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]2 r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 2   1 2 ∓2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1   d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r     1 dp, (44) ∓2π2T 2 x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      ∂s 1 From the substitution from Eqs. (41), (43), (44) and = , (47) (55) into Eq. (21), we get ∂ T ∂p which proves the thermodynamic consistency, Eq. (12). = n. (45) ∂µ Also, from Eq. (46) we can write another form of T energy density Also, from Eqs. (39) and (40), we get a direct relation  = T s p+µ n. (48) between entropy and energy density, − p  µ n By differentiating the energy density with respect to n, s = + . (46) we get T T − T Thus, at constant n and p the differentiation of entropy ∂ ∂p ∂µ ∂p ∂µ ∂µ = +µ+n = +µ+n . (49) density with respect to energy density leads to ∂n −∂n ∂n −∂µ ∂n ∂n s

053107-9 Chinese Physics C Vol. 41, No. 5 (2017) 053107

Then, by substituting from Eq. (45) into the previous which is a proof of the thermodynamic self-consistency. equation, To prove the fourth equation in Eq. (12), the differ- ∂ entiations of Eqs. (38), (39) and (40) with respect to T = µ, (50) ∂n read s

1 2 d 2 x ∞ p ε (x) (x T ) W B 1 ∂n g c,d,r 0 − r = −   dp π2 2   1 3 ∂T 2 T 0 x d Z (d d c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 2   1 2 2 T x d Z0 (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ζ (x) (x T ) W B 1 g c,d,r 0 − r +   dp π2 2   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]2 r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 2   1 2 ∓2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 d 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r     1 dp, (51) ∓2π2T 2 x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 2 x ∞ p ε (x) (x T ) W B 1 ∂ g c,d,r 0 − r =    dp π2 2   1 3 ∂T 2 T 0 x d Z (d d c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1  d 2 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 2   1 2 ∓2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 2 x ∞ p ζ (x) (x T ) W B 1 g c,d,r 0 − r   dp π2   1 2 −2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1 2  d 2 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 2   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1   d 2 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r     1 dp 2π2T 2 x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r     

053107-10 Chinese Physics C Vol. 41, No. 5 (2017) 053107

1 2 d 2 x ∞ p ε (x) (x T ) W B 1 g µ c,d,r 0 − r    dp π2 2   1 3 ∓2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1   d 2 x ∞ p ε (x) (x T ) W B 1 g µ c,d,r 0 − r   dp π2 2   1 2 2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ζ (x) (x T ) W B 1 g µ c,d,r 0 − r +   dp π2 2   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 x ∞ p ε (x) (x T ) W B 1 g µ c,d,r 0 − r   dp π2 2   1 2 ∓2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1   d 2 x ∞ p ε (x) (x T ) W B 1 g µ c,d,r 0 − r     1 dp, (52) ∓2π2T 2 x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 2 x ∞ p ε (x) (x T ) W B 1 ∂s g c,d,r 0 − r =    dp π2 3   1 3 ∂T 2 T 0 x d Z (d d c)[1 ε (x)] r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1  d 2 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 3   1 2 ∓2 T 0 x d Z (d d c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 2 x ∞ p ζ (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 3   1 2 −2 T 0 x d Z (1 c)2 [1 ε (x)]2 r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      1 2 d 2 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r   dp π2 3   1 2 2 T 0 x d Z (1 c)2 [1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r    1   d 2 2 x ∞ p ε (x) (x T ) W B 1 g c,d,r 0 − r     1 dp. (53) 2π2T 3 x d Z0 (1 c)[1 ε (x)]r2 (1 x/r)2 1+W B 1 −  c,d,r − 0 − r      Then, the substitution from Eqs. (40), (51), (52) and So far, we have proved that the definitions of temper- (53) into Eq. (29) gives ature, number density, chemical potential and entropy density within our formalism for nonextensive quantum ∂p = s. (54) statistics lead to expressions which satisfy the first law ∂T µ

053107-11 Chinese Physics C Vol. 41, No. 5 (2017) 053107 of thermodynamics. 4.1 Transverse momentum distributions

4 Confronting our calculations with var- The particle distributions at large transverse mo- ious experimental results menta are indeed very interesting phenomena in high- energy particle production, but as discussed by Bialas The statistical-thermal models have been successful [16], the applicability of the statistical-thermal models in reproducing particle ratios and yields at different ener- in this regime of the transverse momenta is debatable. gies [1, 8–10, 35–37]. In these models, the hadronic phase These models, either extensive or nonextensive, are ex- can be modelled at chemical and thermal equilibrium. cellent approaches at low pT . The scope of this paper is Accordingly, fitting parameters can then be identified. the examination of the thermodynamic self-consistency They construct a set of various thermal parameters. The and then implementing the proposed approach in char- most significant ones are the chemical freeze-out temper- acterizing both transverse momentum distributions and ature and baryon chemical potential [1]. particle ratios and yields. The deconfined phase is dominated by quarks and As discussed in the introduction, the hadronic clus- gluons degrees-of-freedom. Most of their information ters are assumed to undergo thermal decays, but simulta- cannot be recognized due to the nature of such partonic neously move in the transverse direction with a fluctua- QCD matter. While the integrated particle-yields are tion Lotentz factor [16]. Recently, this statistical cluster- successfully constructed in the final state, the transverse decay model was implemented in high-energy physics [17] momentum distribution should be described by a com- and it was concluded that the well-known Tsallis distri- bination of transverse flow and statistical distributions bution can be obtained in a very special case, namely, (particle ratios and yields). In other words, the latter the fluctuations of the Lorentz factor and the relativis- contains contributions from earlier stages of the colli- tic temperature are given by Beta and Gamma distribu- sion, while the particle ratios and yields are conjectured tions, respectively. The role of the statistical cluster- to be fixed during the chemical and thermal equilibrium decay and its possible connections with the approach stages. Thus, we plan to confront this new generic ap- proposed in the present work shall be subjects of future proach with both types of experimental results, namely, works. transverse momentum spectra, and particle ratios and We first implement the generic nonextensive statis- yields. This shall be introduced in the sections that fol- tical approach [27] in order to reproduce various trans- low. We intend to prove whether the proposed approach verse momentum distributions measured in different ex- is indeed able to reflect the statistical nature of the sys- periments [23, 38–42]. For quantum statistics, the total tem of interest. number of particles can be determined from Eq. (55)

1 d 2 xi ∞ p ε (x ) W B 1 V c,d,r i 0 − r   3p N = 3 gi   1 d , (55) 8π xi d i Z0 (1 c)[1 ε (x )](r x ) 1+W B 1 X −  c,d,r i − i 0 − r      + + −1 where xi = (µ Ei)/T , i [π ,K ,p], and r = [1 c + cd] . Their degeneracy factors read gπ+ = gK+ = gp = 1. For antiparticles, µ−is replaced∈ by µ. − The corresponding momentum− distribution for particles (or antiparticles) is given as

1 x d ε (x ) W B 1 i 1 Ed3N V c,d,r i 0 − r   3 = 3 giEiT   1 , (56) 2π d p 8π xi d i (1 c)[1 ε (x )](rT µ+E ) 1+W B 1 X −  c,d,r i − i 0 − r      2 2 where εc,d,r (xi) is defined in Eq. (4). In terms of rapidity (y) and transverse mass (mT i = pT +mi ), the transverse momentum distribution can be given as p 1 d2N V = 3 giT mT i coshy (57) 2π dymT dmT 8π i X 1 µ m coshy µ m coshy d ε − T i W B 1 − T i c,d,r T 0 − rT   "   # 1 . × µ m coshy µ m coshy d (1 c) 1 ε − T i (rT +m coshy µ) 1+W B 1 − T i −  c,d,r T T i − 0 − rT    "   #!

053107-12 Chinese Physics C Vol. 41, No. 5 (2017) 053107

At mid-rapidity, i.e. y = 0, and µ = 0, 1 d2N V = 3 giT mT i 2π dymT dmT 8π i X 1 m ( m ) d ε − T i W B 1 − T i c,d,r T 0 − rT   "   # 1 . (58) × m ( m ) d (1 c) 1 ε − T i (rT +m ) 1+W B 1 − T i −  c,d,r T T i 0 − rT    "   #! The transverse momentum distribution reads 1 d2N V = 3 giT mT i 2π pTdydpT 8π y=0 i X 1 m ( m ) d ε − T i W B 1 − T i c,d,r T 0 − rT   "   # 1 . (59) × m ( m ) d (1 c) 1 ε − T i (rT +m ) 1+W B 1 − T i −  c,d,r T T i 0 − rT    "   #! At mid-rapidity, i.e. y = 0, but µ = 0 6 1 d2N V = 3 giT mT i 2π pTdydpT 8π y=0 i X 1 µ m (µ m ) d ε − T i W B 1 − T i c,d,r T 0 − rT   "   # 1 . (60) × µ m (µ m ) d (1 c) 1 ε − T i (rT µ+m ) 1+W B 1 − T i −  c,d,r T − T i 0 − rT    "   #!

We can now fit various transverse momentum distri- a smaller pT-window. In the panel (a), we can compare butions measured in different experiments, i.e. different between the pp transverse momentum distributions at types of collisions and different collision energies, rapid- different collision energies. There are obvious trends with 2 ity, centrality, etc. by using Eq. (59) and Eq. (60) at increasing collision energy d N/dydpT. Furthermore, at 2 mid-rapidity. Expression (60) takes into consideration a given collision energy, d N/dydpT exponentially de- vanishing and finite baryon chemical potentials [1, 43]. creases rapidly with increasing pT. Figure 1 depicts the transverse momentum distribu- All measurements are performed at mid-rapidity and tions for the charged particles K+,π+ and p and their we assume a vanishing chemical potential. The qual- antiparticles measured in different types of collisions at ity of the statistical fit looks excellent, see Table 1. In various collision energies; p+p collisions at 0.54 [38], 0.9, p+p collisions at 0.546 TeV, we find that c 1, while 2.36 and 7 TeV [23], p+Pb collisions at 5.02 TeV [41], d 1.35. Similar values are also obtained in'p+A and Pb+Pb collisions at 2.76 TeV [42], d+Au collisions and A+A' collisions, panels (b) and (c). These two values of Au+Au collisions at 0.2 TeV [39, 40]. This set of mea- the equivalence class (c,d) = (1,d) are associated with sured particles combines the momentum spectra of six asymptotically stable systems. At the resulting d, which charged particles. Here, we zoom out the smallest pT re- is positive, asymptotically stable systems - in turn - are gion. The entire pT spectra are illustrated in Fig. 2. All associated with stretched exponential distributions [45], these distributions are fitted to the generic nonextensive 1/d statistical approach, Eq. (59). The resulting fit param- lim εd,r(x) = exp dr (1 x/r) 1 . (61) eters are given in Table 1. For a better comparison, the c→1 − − − n h io p+p, p+A and A+A results are separately depicted in The entropy becomes a stretched exponential [44] leads Fig. 1(a), (b), and (c), respectively. to the special case The entire pT-range is presented in Fig. 2. It is worth highlighting that the statistical fit was performed over 1−d −d εd,r(x) = r d exp(dr)Γ (1+d,dr ln x) r x. (62) the complete range of pT. Figure 1 is there to zoom out − −

053107-13 Chinese Physics C Vol. 41, No. 5 (2017) 053107

102

101

100

10−1

10−2

10−3

10−4

Fig. 1. (color online) Transverse momentum distributions for the charged particles K+,π+ and p and their antipar- ticles measured in the experiments UA1 [38], PHENIX [39], STAR [40], ATLAS [23] and ALICE [41, 42] (symbols) are compared with calculations from the generic nonextensive statistical approach (curves). The resulting fit pa- rameters are listed in Table 1. The pp, pA and AA results are separately depicted in panels (a), (b) and (c), respectively.

102 alence class defines a nonextensive entropy which is lin- UA1 0.54 TeV ATLAS 0.9 TeV early dependent on or - in other words - composed of 100 ATLAS 2.36 TeV ATLAS 7 TeV extensive entropies, such as Renyi [46]. Its optimized T p

d − y 10 2 entropy is given by probability distribution (pi), which d /

N includes the Lambert function, 2 −4 )d i 10 p

T β 1 p Sβ = p ln , (63) −6 i 10 pi (1/2 i X   −8 10 where β c and pβ is the probability that the entire ≡ i 10−10 phase-space is occupied. Although this type of entropy 0 10 20 30 40 50 looks analogous to the Tsallis one, (1 pq)/(1 q), it − i − i − pT is apparently different, at least, as the latter reaches sin- Fig. 2. (color online) The same as in Fig. 1 includ- gularity, at q = 1, the so-called ShannonP limit. Further ing the largest pT-region. differences have been elaborated in Ref. [46]. It is worth highlighting that the resulting freezeout It has been proved that the numerical analysis becomes temperatures (Tch) rise with increasing collision energy impossible without the general case of the Gamma- (√sNN). Possible explanations of this observation are ∞ postponed for a future work. We focus on the discussion function, Γ (a,b) = dtta−1 exp( t). b − of their phenomenologies. The Tsallis nonextensive ap- Furthermore, wZe observe that the ATLAS measure- proach was also used in conducting the same study [47]. ments are associated with positive c (close to unity), Accordingly, there is a remarkable difference between Tch while d can be approximated as d 1. Such an equiv- obtained from our approach and the one deduced from '

053107-14 Chinese Physics C Vol. 41, No. 5 (2017) 053107 the Tsallis-type approach [47]. The latter are relatively tential drops from 29 MeV at 0.2 TeV to 0 MeV at smaller than the earlier results (compare Table 1 of the 5.02 TeV. present work with Table 1 of Ref. [47]). Figure 1(c) presents a comparison between A+A col- Figure 1(b) shows the transverse momentum distri- lisions at different collision energies; Au+Au collisions butions of the three charged particles K+,π+ and p at 0.2 TeV and Pb+Pb collisions at 2.76 TeV for the and their antiparticles measured in d+Au collisions at momentum spectra of the three charged particles K+,π+ 0.2 TeV and in p+Pb collisions at 5.02 TeV. The equiv- and p and their antiparticles. Despite the system size alence class can be constricted from c 1, and d. difference, we also find that c 1, while d slightly in- The latter is found to be almost energy-indep' endent; creases from 1.221 0.055 at 0.2'TeV to 1.357 0.023 at 1.291 0.006 at 0.2 TeV to 1.331 0.004 at 5.02 TeV. 2.76 TeV. The freezeout temperature shows an energy- The freezeout parameters are very interesting. The independent behavior; 155 12.450 MeV at 0.2 TeV to resulting freezeout temperature slightly increases from 165 12.845 MeV at 2.76 TeV. The baryon chemical po- 170 13.038 MeV at 0.2 TeV to 175 13.228 MeV at tential behaves similarly to in p+A collisions, shown in 5.02TeV, while the corresponding baryon chemical po- Fig. 1(b).

Table 1. Various parameters deduced for the statistical fit of the transverse momentum calculations based on the generic nonextensive statistical approach (present work) to various measurements, shown in Fig. 1.

2 experiment √sNN/TeV centrality Tch/MeV µb/MeV d c χ UA1 (p+p) 0.546 0% 110 10.488 0.0 1.35 0.135 0.999 0.100 0.384    ATLAS (p+p) 0.9 0% 140 11.832 0.0 1.03 0.103 0.945 0.095 0.766    ATLAS (p+p) 2.36 0% 160 12.649 0.0 1.06 0.106 0.945 0.095 1.651    ATLAS (p+p) 7.0 0% 150 12.247 0.0 1.05 0.105 0.930 0.093 1.115    PHENIX/STAR (d+Au) 0.2 0% 20% 170 13.038 29 1.291 0.006 0.999 0.001 1.884 −    ALICE (p+Pb) 5.02 0% 5% 175 13.228 0.0 1.333 0.004 0.999 0.001 0.902 −    PHENIX/STAR (Au+Au) 0.2 0% 10% 155 12.450 25 1.221 0.055 0.999 0.001 1.584 −    ALICE (Pb+Pb) 2.76 0% 5% 165 12.845 0.0 1.357 0.023 0.999 0.001 0.164 −   

From the fit parameters listed in Table 1, we can process, and if the latter, whether this follows Tsallis summarize that except for the ATLAS measurements, approach. c 1, while d > 1. These values are obtained for 4.2 Particle ratios and yields differen' t system sizes and at various collision energies. They refer to nonextensivity but not of Tsallis type. We The particle ratios π−/π+, K−/K+, p¯/p, Λ¯/Λ, Ω¯/Ω, note again that the BG extensivity is guaranteed only Ξ¯/Ξ, K−/π−, K+/π+, p¯/π−, p/π+, Λ/π−, Ω/π−, and at c = d = 1. It has been pointed out [16] that the Ξ¯/π+ measured in Au+Au collisions in the STAR ex- Tsallis algebra, which is mainly implemented through periment at energies 200 GeV, 62.4 GeV, and 7.7 GeV replacing the exponential and logarithmic functions by are statistically fitted by means of the HRG model, in their counterpart expressions in the Tsallis nonexten- which the generic nonextensive statistical approach is im- sive approach, can be scaled as power laws. Such a plemented. The number density can be derived from the scaling can also be obtained in the so-called statisti- partition function and accordingly the particle ratios can cal cluster-decay. Thus, the good fitting of the trans- be determined. verse momentum distributions by the Tsallis nonexten- We take into account all possible decays into the par- sivity would be misleading [16], as the role of the sta- ticle of interest and their branching ratios. µ, T , c, and tistical cluster-decay is apparently ignored. In a forth- d are taken as free parameters. The strangeness chem- coming work, we shall estimate the contribution of the ical potential is calculated at each value assigned to µ, statistical cluster-decay to the possible power-laws in the T , c, and d so that an overall strangeness conservation nonextensive fit, with a special emphasis on Tsallis-type is guaranteed. The results depicted in Fig. 3 represent nonextensivity. the best agreement between measurements and calcu- In the section that follows, we confront our approach lations, i.e. at the smallest χ2 value. The results at with different particle ratios and yields, which likely 200 GeV [panel (a)], 62.4 GeV [panel (b)] and 7.7 GeV cannot be scaled as power laws, as such. The result- [panel (c)] are depicted in Fig. 3. We observe that the ing (non)extensivity parameters are conjectured to shed quality of the fits weakens with the collision energies; light on the statistical nature of the particle produc- χ2/dof = 1.105, χ2/dof = 2.771 and χ2/dof = 8.146 at tion and whether it is an extensive or a nonextensive 200 GeV, 62.4 GeV and 7.7 GeV, respectively.

053107-15 Chinese Physics C Vol. 41, No. 5 (2017) 053107

Fig. 3. Panel (a): different particle ratios deduced from the generic axiomatic-nonextensive statistical approach (dashed lines) are compared with the experimental results at 200 GeV (symbols). Panels (b) and (c) show the same but at 62.4 and 7.7 GeV, respectively. The exponents c, d and χ2 are given in the top right corner.

As mentioned, our fits for the transverse momentum unity, referring neither to BG extensivity nor to Tsallis distributions and the debatable interpretation of whether nonextensivity. This conclusion still needs further anal- the resulting parameters are due to power laws stem- ysis at other collision energies. We shall devote a future ming from Tsallis-algebra should be a subject of a fu- work to examine such a relation. ture work. Here, we have analysed another thermody- In Fig. 4, we present fits for various particle yields, namic quantity, the particle ratios, which are conjectured π−, π+, K−, K+, p,¯ p, Λ,¯ Λ, Ω,¯ Ω, Ξ,¯ and Ξ, measured in to highlight the (non)extensivity, as they are not to be Au+Au collisions in the STAR experiment (symbols) at scaled as power laws. In light of this assumption, we 200 GeV, 62.4 GeV, and 7.7 GeV, respectively, fitted to can discuss the resulting parameters and their physical calculations from the HRG model with generic nonexten- meanings. sive statistical approach (dashed lines). The equivalence

The resulting freezeout temperature (Tch) and baryon class is given in the top-right corners of each graph. To- chemical potential (µb) can be summarized as follows. gether with the resulting freezeout parameters, they are at 200 GeV, Tch = 148.05 12.168 MeV and listed in Table 2. µ =•23.94 4.89 MeV,  In summary, we have found that b  at 62.4 GeV, Tch = 179.13 13.384 MeV and for pT spectra except ATLAS, c 1 and d > 1. The µ =•57.33 7.57 MeV, and  nonextensiv• e entropy is associated with' stretched expo- b  at 7.7 GeV, Tch = 145.32 12.055 MeV and nentials, where the Lambert function reaches its asymp- µ =•384.3 19.6 MeV,  totic stability. b  which are obviously very compatible with those de- for ATLAS pT spectra, c < 1, while d 1. The duced from BG statistics [43]. The equivalence class nonextensiv• e entropy is linearly composed of'extensive reads entropies. The reasons why the ATLAS measurements At 200 GeV, c = 0.971 0.097 and d = 0.949 0.09, look different from the others shall be discussed in a fu- • at 62.4 GeV, c = 0.9750.098 and d = 0.923 0.091 ture work. and•   for particle ratios and yields, c < 1 and d < 1. This at 7.7 GeV, c = 0.995 0.1 and d = 0.972 0.097. is kno• wn as (c,d)-entropy, where d > 0 and Lambert •   These can be approximated as both c and d lie below functions W0 rise exponentially.

053107-16 Chinese Physics C Vol. 41, No. 5 (2017) 053107

Fig. 4. (color online) The same as in Fig. 3 but for the particle yields.

Table 2. The freezeout parameters and equivalence class deduced for the statistical fit of the generic nonextensive statistical approach (present work) compared with the experimental results on particle yields, Fig. 4.

2 √sNN/GeV centrality Tch/MeV µb/MeV d c R/fm χ 200 0% 160.54 12.67 25.20 5.02 0.965 0.097 0.975 0.098 2.044 1.126 5.395      62.4 0% 170.00 13.04 57.83 7.60 0.962 0.096 0.965 0.097 1.796 1.056 2.816      7.7 0% 147.31 12.14 388.08 19.7 0.972 0.097 0.995 0.101 1.848 1.071 2.701     

5 Conclusions and outlook pressions which satisfy the laws of thermodynamics. The second part of this paper introduced implemen- We have presented a systematic study of the ther- tations of the generic axiomatic-nonextensive statistical modynamic self-consistency of the generic axiomatic- approach to high-energy physics. We started with the nonextensive statistical approach, which is characterized transverse momentum distributions measured in central by two asymptotic properties. To each of these, a scaling collisions, in different system sizes and at various collision function was assigned. These scaling functions are esti- energies. We found that c 1 and d > 1, which describe ' mated by exponents c and d for the first and second prop- nonextensive entropy associated with stretched exponen- erty, respectively, in the thermodynamic limit of grand- tials, in which the Lambert function behaves asymptot- canonical ensembles of classical (Boltzmann) and quan- ically. For the ATLAS pT spectra, c < 1, while d 1. ' tum gas (Fermi-Dirac and Bose-Einstein) statistics of a This is nonextensive entropy, which linearly involves ex- gas composed of various hadron resonances. We started tensive entropies, such as Renyi. All these values differ with the first and second laws of thermodynamics, which from the equivalence class characterizing BG and Tsallis characterize the statistical system of interest. The ther- statistics. The role of statistical cluster decays was also modynamic properties of that system were determined highlighted. Their contributions to the possible power- and it was confirmed that both laws of thermodynamic laws should be first evaluated, so that those stemming are fully verified. We have proved that the definitions of from Tsallis-type nonextensivity can be determined. temperature, number density, chemical potential and en- We also calculated different particle ratios and yields, tropy density within the generic axiomatic-nonextensive which likely are not to be scaled as power laws. From classical and quantum statistical approach lead to ex- the resulting exponents c and d, we can judge whether

053107-17 Chinese Physics C Vol. 41, No. 5 (2017) 053107

the particle production is a (non)extensive process. We Finally, we conclude that the proposed generic nonex- found that c < 1 and d < 1 referring to (c,d)-entropy, tensive statistical approach is thermodynamically self- where the Lambert functions rise exponentially. This consistent and able to reveal the statistical nature of finding points to neither BG extensivity nor Tsallis various processes taking place in high-energy collisions, nonextensivity. such as transverse momentum spectra and particle ratios From the statistical fit of both sets of experimental and yields.

results [transverse momentum distribution (pT) and par- Many authors endorse the assumption that Tsallis- ticle ratios and yields], the resulting freezeout parame- type nonextensivity originates in fluctuations, correla-

ters, Tch and µb, are fairly compatible with those deduced tions and inter-particle interactions. We briefly discussed from BG statistics. We believe that the statistical prop- all these in the introduction. Wilk et al. [48] proposed erties, whether extensive or nonextensive, should not be that the nonextensive parameter (q) is related to fluc- related to intensive or extensive thermodynamic quanti- tuations in the inverse temperature. It is believed that ties, such as temperature and baryon chemical potential. the temperature reflects the impact of the nonextensiv- These are strongly associated with the proposed equiva- ity. Since this interesting topic lies out of the scope of lence class (c,d). The latter characterizes BG, or Tsallis this paper, we plan in a future work to examine the pos- or even generic statistics, where (1,1), or (c,0) or (c,d), sibilities that the equivalence class (c,d), instead of the respectively. The thermodynamic quantities are conjec- temperature, might affect some or all of the physical pro- tured to partly manifest the physical properties of the cesses, fluctuations, correlations and inter-particle inter- system of interest, while the nature of (non)extensivity actions. characterizes the statistical properties.

References 27 A. N. Tawfik, Lattice QCD thermodynamics and RHIC-BES particle production within generic nonextensive statistics, sub- 1 A. N. Tawfik, Int. J. Mod. Phys. A, 29: 1430021 (2014) mitted to EPJA 2 A. Tawfik, J. Phys. G, 40: 055109 (2013) 28 A. S. Parvan, Eur. Phys. J. A, 51: 108 (2015) 3 R. Hagedorn, Nuovo Cimento Suppl., 3: 147 (1965) 29 A. Deppman, Physica A, 391: 6380 (2012) 4 A. Tawfik, Prog. Theor. Phys., 126: 279 (2011) 30 A. S. Parvan, Phys. Lett. A, 350: 331 (2006) 5 C. Beck, Physica, A, 286: 164 (2000) 31 A. S. Parvan, Phys. Lett. A, 360: 26 (2006) 6 I. Bediaga, E. M. F. Curado, and J. M. de Miranda, Physica 32 J. Cleymans and D. Worku, J. Phys. G, 39: 025006 (2012) A, 286: 156 (2000) 33 J. Letessier and J. Rafelski, Hadrons and Quark-Gluon Plasma, 7 W. M. Alberico, A. Lavagno, and P. Quarati, Eur. Phys. J. C, (Cambridge University Press, UK, 2004), p. 210 12: 499–506 (2000) 34 S. Thurner and R. Hanel, Int. J. Mod. Phys.: Conf. series, 16: 8 S. Tripathy et al., Eur. Phys. J. A, 52: 289 (2016) 105 (2012) 9 A. Khuntia et al, Eur. Phys. J. A, 52: 292 (2016) 35 J. Cleymans, H. Oeschler, K. Redlich, and S. Wheaton, Phys. 10 T. Bhattacharyya et al, Eur. Phys. J. A, 52: 30 (2016) Rev. C, 73: 034905 (2006) 11 A. Deppman, J. Phys. G, 41: 055108 (2014) 36 A. Andronic, P. Braun-Munzinger, and J. Stachel, Nucl. Phys. 12 W. M. Alberico et al, Physica A, 387: 467 (2008) A, 772: 167 (2006) 13 T. Bhattacharyya, R. Sahoo, and P. Samantray, Eur. Phys. J. 37 F. Becattini, J. Manninen, and M. Gazdzicki, Phys. Rev. C, A, 52: 283 (2016) 73: 044905 (2006) 14 Q. A. Wong, Entropy, 5: 220 (2003). 38 C. Albajar et al (UA1 Collaboration), Nucl. Phys. B, 335, 261 15 A. N. Tawfik, Eur. Phys. J. A, 52: 253 (2016) (1990) 16 A. Bialas, Phys. Lett. B, 747: 190 (2015) 39 A. Adare et al (PHENIX Collaboration), Phys. Rev. C, 88: 17 G. Wilk and Z. Wlodarcyk, Phys. Lett. A, 379: 2941 (2015) 024906 (2013) 18 C. Beck and E. G. D. Cohen, Physica A, 322: 267 (2003) 40 B. I. Abelev et al (STAR Collaboration), Phys. Rev. C, 79: 19 F. Sattin, Eur. Phys. J. B, 49: 219 (2006) 034909 (2009) 20 B. I. Abelev et al (STAR Collaboration), Phys. Rev. C, 75: 41 B. Abelev et al (ALICE Collaboration), Phys. Lett. B, 728: 25 064901 (2007) (2014) 21 A. Adare et al (PHENIX Collaboration), Phys. Rev. C, 83: 42 B. Abelev et al (ALICE Collaboration), Phys. Rev. Lett. 109: 052004 (2010) 252301 (2012) 22 K. Aamodt et al (ALICE Collaboration), Eur. Phys. J. C, 71: 43 A. N. Tawfik and E. Abbas, Phys. Part. Nucl. Lett., 12: 521 1655 (2011). (2015) 23 G. Aad et al (ATLAS Collaboration), New J. Phys., 13: 053033 44 C. Anteneodo and A. R. Plastino, J. Phys A: Math. Gen., 32: (2011) 1089 (1999) 24 V. Khachatryan et al (CMS Collaboration), JHEP, 05: 064 45 R. Hanel and S. Thurner, arXir: 1005.0138 [physics.class-ph] (2011) 46 Fariel Shafee, IMA J. Appl. Math., 72: 785 (2007) 25 R. Hanel and S. Thurner, Europhys. Lett., 93: 20006 (2011) 47 J. Cleymans et al,Phys. Lett. B, 723: 351 (2013) 26 R. Hanel and S. Thurner, Europhys. Lett., 96: 50003 (2011) 48 G. Wilk and Z. Wlodarczyk, Phys. Rev. Lett., 84: 2770 (2000)

053107-18