Sei sulla pagina 1di 12

Int J Theor Phys

DOI 10.1007/s10773-011-1015-4
MatterAntimatter Asymmetry in Heavy-Ion Collisions
A. Tawk
Received: 23 July 2011 / Accepted: 3 November 2011
Springer Science+Business Media, LLC 2011
Abstract Implementing modied phase space and distribution function in grand-canonical
ensemble and taking into account the experimental acceptance, the ratios of antiparticle-
to-particle over the whole range of energies are well reproduced by hadron resonance gas
(HRG) model. We introduce a systematic study for antiparticle-to-particle ratios measured in
various pp and AA collisions and nd that the ratios of bosons and baryons get very close
to unity indicating that the matterantimatter asymmetry nearly vanishes at LHC energy.
We predict that the LHC heavy-ion program will produce the same particle ratios as the pp
program implying the dynamics and evolution of the system would not depend on the initial
conditions.
Keywords Collective ow Statistical physics Relativistic heavy-ion collisions Early
Universe
1 Introduction
The elimination of antimatter while leaving behind matter known as baryogenesis is one of
the most fundamental problems in the modern physics. The Big Bang and the cosmologi-
cal standard models assuming homogeneous and isotropic matter (antimatter) distribution
are conjectured to allow for an equal creation of matter and antimatter. What we likely ex-
perience in our Universe is that the matter asymmetry seems to be 100%. Furthermore,
the Universe is an obvious evidence that matter and antimatter are not fully symmetric.
There are various astrophysical and fundamental observations supporting such conclusion.
The materialistic objects under non-extreme conditions have baryons and/or leptons only.
The ratio of n
net-baryon
to n
photon
measured in cosmic microwave background (CMB) era by
WMAP are very much small 10
10
[1, 2]. In the whole solar system, there is no evidence
for the existence of antimatter. Even in the ultra highest energy cosmic rays (UHECR), the
number of antiprotons, for instance, is in a good (fully) agreement with the production as
A. Tawk ()
Egyptian Center for Theoretical Physics (ECTP), MTI University, Cairo, Egypt
e-mail: a.tawk@eng.mti.edu.eg
Int J Theor Phys
secondary particles from nuclear collisions. The observed gamma burst ux does not indi-
cate the existence of antimatter in admixture of galaxies. It is apparent that the recent results
of Large Hadron Collider (LHC) on antiparticle-to-particle ratios show that the produced
matter and antimatter are almost equal [3, 4].
Under extreme conditions, like hot and dense QCD of the nuclear collisions, the anti-
matter can be created. In present work, a phenomenological survey of the matterantimatter
asymmetry measured in nuclear interactions is introduced. The baryon and boson yield ra-
tios are studied. The collective properties of hot and dense QCD is one of main objectives
of the heavy-ion program. The possible modication of the transport coefcients, like phase
structure and effective degrees of freedom, with increasing energy and size of the interact-
ing system likely provides fruitful tools to study the collective properties. Furthermore, such
modications seem to comprehensively characterize the particle and/or antiparticle produc-
tion itself [5]. A universal description for particle and/or antiparticle production at different
center-of-mass-energies of nucleonnucleon (NN) collisions

s
NN
(hereafter simply

s)
can be achieved by studying the dynamical uctuations [5, 6] and the multiplicities of parti-
cle yield ratios [7, 8]. The particle yield ratios are not only able to determine the freeze-out
parameters, temperature T and chemical potential , but also eliminateto a very large
extendthe volume uctuations and the dependence of the freeze-out surface on the initial
conditions. On the other hand, the uctuations and multiplicities of particle ratios of parti-
cle (antiparticle) production in nuclear collisions are also very much essential regarding to
examining the statistical/thermal models [9, 10], characterizing the chemical/thermal equi-
libration [11] and nally search for unambiguous signals for the creation of the quarkgluon
plasma (QGP) [12, 13].
The freeze-out parameters (T and ) can be determined by thermal models through com-
bining various uctuations and multiplicities of integrated particle (antiparticle) ratios [7, 8].
The search for common properties of these parameters at different center-of-mass-energies
has a long tradition [14, 15]. Different models have been suggested so far [1419]. In present
work, constant entropy density s normalized to T
3
[16, 17] is used. This universal condition
obviously implies that the hadronic matter is characterized by constant degrees of freedom.
To illustrate this, the quantity s/T
3
can be scaled by
2
/4. Apparently, this quantity reects
the phase space volume. It results in the overall effective degrees of freedom of free gas [20].
It is worthwhile to mention that the phase transition (conned hadrons deconned QGP)
itself is accompanied by a drastic decrease of the degrees of freedom [10].
Using grand-canonical ensembles and modied phase space in the hadron resonance gas
(HRG) model makes it possible to fairly describe different features of the experimental data
[21, 22] including particle multiplicities, dynamical uctuations of various particle ratios
[23] and thermodynamics of the strongly interacting system below the critical tempera-
ture T
c
. With reference to scope of present work, it has been noticed that the uctuations
and multiplicities over the whole range of

s exhibit a non-monotonic behavior [21, 22].


In heavy-ion collisions, the enhanced antiparticles are conjectured as indicators for the
formation of deconned QGP [2426], whereas the possible annihilation might suppress
such an enhancement [27, 28]. This would explain why the values of antiparticle-to-particle
ratios in nucleonnucleon (pp) collisions are higher than in heavy-ion collisions. There-
fore, the initial conditions and formation time can be reected by the surviving antiparticle.
Decelerating or even stopping of incident particle (projectile) and its break up in inelastic
collisions have been discussed in literature [2931]. Therefore, the antiparticle-to-particle
ratios can be used to study particle (or antiparticle) transport and production and therefore
would have signicant cosmological and astrophysical consequences. By mentioning as-
trophysics, it seems in order now to recall that the n
p
/n
p
ratios among others have been
Int J Theor Phys
calculated and also observed in the cosmic rays revealing essential details on astrophysical
phenomena [3234].
In present work, different antiparticle-to-particle ratios are calculated in the HRG model
and compared with the heavy-ion collisions at

s ranging from AGS to LHC, few GeV to


couple TeV, respectively. Thus, the ratios and their description in the thermal model seem
to provide fruitful tools to study the evolution of the matterantimatter asymmetry with
changing energy. In Sect. 2, the methods are introduced. Section 3 is devoted to the results
and their discussion. Some cosmological and astrophysical consequences are discussed in
Sect. 4. The conclusions are listed in Sect. 5.
2 Methods
The antiparticle-to-particle ratios over the whole range of the center-of-mass-energy

s
are studied in framework of a thermal model. The hadron resonances treated as a free
gas [10, 3538] are conjectured to add to the thermodynamic quantities in the hadronic
phase of heavy-ion collisions. This statement is valid for free as well as strong interactions
between the hadron resonances themselves. It has been shown that the thermodynamics of
strongly interacting system can be approximated to an ideal gas composed of hadron res-
onances [3739]. Also, it has been shown that the thermodynamics of strongly interacting
system can be approximated to an ideal gas composed of hadron resonances with masses
1.8 GeV [3739]. It is obvious that the heavier constituents, the smaller thermodynamic
quantities. The main motivation of using the Hamiltonian is that it contains all relevant
degrees of freedom of conned and strongly interacting matter. It implicitly includes the in-
teractions that result in the formation of new resonances. In addition, the HRG model is used
to provide a quite satisfactory description of particle production and collective properties in
heavy-ion collisions [19, 2123, 3539].
At nite temperature T , strange
S
and isospin
I
3
and baryochemical potential
B
, the
partition function of one single particle and one single antiparticle reads [21, 22]
lnZ
gc
(T, V, ) =

g
i
2
2
V
_

0
k
2
dk
_
ln
_
1 Qexp
_

i
(k)
T
__
+ln
_
1 Qexp
_

i
(k)
T
___
, (1)
where
i
(k) =(k
2
+m
2
i
)
1/2
is the i-th particles dispersion relation (single-particle energy)
in equilibrium and stands for bosons and fermions, respectively. g is the spinisospin
degeneracy factor and
n
q

m
s
stand for the quark phase space occupancy parameters,
where n and m being the number of light and strange quarks in the hadron of interest,
respectively. is the chemical potential and =exp(c/T ) is the fugacity factor, where c is
the corresponding charge. Multiplying by c, like baryon, light and strange quark numbers,
etc., makes it possible to count for particle and antiparticle. Since no phase transition is
conjectured in the HRG model (its validity is limited below T
c
), summing over all hadron
resonances results in the nal thermodynamics in the hadronic phase, exclusively.
The generic parameter Q( x,

k) is conjectured to reect the change in the phase-space,


when the hadronic degrees of freedom are replaced by the partonic ones and vice versa.
The parameter =exp() can be interpreted as a measure for the non-extensivity, where
will be elaborated below. It gives the averaged occupancy of the phase space relative to
equilibrium limit. On the other hand, when assuming nite time evolution of the system,
Int J Theor Phys
then
i
can be seen as a ratio of the change in the particle numbers before and after the
chemical freeze-out i.e.,
i
=n
i
(t )/n
i
(). The chemical freeze-out turns to be dened as
the time scale, at which no longer particle production is possible and the collisions become
entirely elastic. In order to complete this picture, it is worthwhile to mention that in case of
phase transition,
i
is expected to be larger than unity. The reasons are the large degrees of
freedom, weak coupling and expanding phase space above the critical temperature to QGP.
The parameter is a Lagrange multiplier in the entropy maximization. Its physical mean-
ing is a controller over the number of particles in the phase space i.e., acting as chemical
potential, = lnE lnT lnN. It also combines intensive variables, T and N with an
extensive one E =

n
i
g
i

i
, where
i
is energy density of i-th cell in the phase space of in-
terest. The most probable state density can be found by the Lagrange multipliers, where one
of them, , has been expressed in term of the second one, 1/T , and the occupation num-
bers of the system. Apparently, gives how the energy E is distributed in the microstates
of the equilibrium system and therefore, can be understood as another factor controlling the
number of occupied states at the micro-canonical level.
The quark chemistry is given by relating the hadronic chemical potentials and to the
quark constituents.
n
q

m
s
with n and m being the number of light and strange quarks,
respectively.
B
= 3
q
and
S
=
q

s
, with q and s are the light and strange quark
quantum number, respectively. The baryochemical potential for the light quarks is
q
=
(
u
+
d
)/2. The iso-spin chemical potential
I
3
=(
u

d
)/2. Therefore, vanishing
I
3
obviously means that light quarks are conjectured to be degenerate and
q
=
u
=
d
.
The strangeness chemical potential,
S
, is calculated in HRG model as a function of T
and
B
. At nite T and
B
,
S
is calculated under the condition of the overall strangeness
conservation. The procedure is given in Ref. [37].
Absolving chemical freeze-out means that the hadrons nally decay to stable hadrons or
resonances
n
nal
i
=n
direct
i
+

j=i
b
ji
n
j
, (2)
where b
ji
being decay branching ratio of j-th hadron resonance into i-th stable particle
or antiparticle. The channel in which the particle and its antiparticle simultaneously appear
is not taken into account. Such a channel of mixture decay seems to play an essential role in
describing dynamical uctuations of particle (antiparticle) yield ratios [16, 17, 19, 2123].
In grand-canonical ensemble, the particle number density is no longer constant. In an en-
semble of N
r
hadron resonances,
n =
N
r

i
g
i
2
2
_

0
k
2
dk
_

i
e

i
(k)/T
(Q )
1

i
e

i
(k)/T
+

1
i
e

i
(k)/T
(Q )
1

1
i
e

i
(k)/T
_
. (3)
In carrying out the calculations given in present work, full grand-canonical statistical
sets of the thermodynamic quantities have been used. Corrections due to van der Waals
repulsive interactions have not been taken into account [37]. Seeking for simplicity, one
canfor a momentassume the Boltzmann approximations, especially at ultra-relativistic
energies, in order nd out the parameters, on which n/n ratios depend. Then, at nite isospin
fugacity
I
3
, the ratios of antiproton-to-proton and antikaon-to-kaon, respectively, can be
approximated as
n
p
n
p

p
=
_

2
u

d
_
2
, (4)
n
K

n
K
+

K
+
=(
u

S
)
2
. (5)
Int J Theor Phys
These two expressions can be used to calculate the light and strange quarks chemical po-
tentials at various energies using the experimentally measured ratios. As discussed above,
both baron and strange potentials are correlated to assure an overall strangeness conserva-
tion [37].
3 Results and Discussion
As mentioned above, the formation of deconned QGP strongly depends on the initial con-
ditions in the early stages in the nuclear interactions, where the particle (antiparticle) number
transport (deceleration of the incoming projectile) or even its entire stopping is assumed to
be achieved [4042]. Therefore, it is widely conjectured that such deconned matter would
not be produced, especially in pp collisions at low energies as will be given shortly. On
the other hand, the degrees of transport likely affect the overall dynamical evolution of
the system. The initial parton equilibrium [43, 44], thermal/chemical equilibrium [45], time
evolution [46, 47] and the nal particle production [48] are affected as well.
In Fig. 1, the baryonic n
p
/n
p
ratios are given in dependence on center-of-mass-energy

s ranging from AGS to LHC, few GeV to couple TeV, respectively. The pp collisions
(empty symbols) are compared with the heavy-ion collisions (solid symbols). In pp colli-
sions, the spatial and time evolution of the system is too short to assure initial conditions
required to drive the parton matter into deconned QGP. We notice that n
p
/n
p
ratios up
to the RHIC energy are smaller than 100% indicating an overall excess of proton n
p
over
antiproton n
p
. Increasing

s makes the accessibility of p and p nearly equal.


From the historical point-of-view, the formation of QGP in laboratory is governed by the
Bjorken ow [49], which determines the critical energy density as

m
T

R
2

0
dN(
0
)
dy
, (6)
where y is the rapidity. R is the spacial extension of the interacting system, while
0
is the
formation time. At low incident energies,
pp
is much higher than
AA
making it much easier
for AA collisions to form QGP than for pp collisions. Currently, we assume that the QGP
formation strongly depends on the initial conditions [50] and the time evolution of the inter-
acting system. On the other hand, an unambiguous signature for the QGP formation is the jet
quenching in the partonic dense medium [51]. At RHIC and LHC energies [5254], the pp
collisions show that there is always an away side jet for each jet. This is not the case in the
AA collisions, where R
AA
N
AA
/R
2
N
pp
is suppressed with increasing transverse momen-
tum p
t
. It would indicate that QGP in AA collisions is much accessible by HRG than pp
collisions. The HRG model basically takes into account the thermodynamic contributions
by heavy resonances (3). It is therefore suitable to compare it with the heavy-ion collisions.
It was not possible to compare pp collisions with HRG consisting of one component, like
pion of proton, [37]. Indeed, we have shown that the HRG model can be successfully used to
describe the heavy-ion collisions and the thermodynamics results by the lattice QCD simu-
lations as well [10, 16, 17, 19, 3537]. The collective ow of the strongly interacting matter
in heavy-ion collisions makes it very much obvious that the HRG model underestimates
the n
p
/n
p
ratios measured in pp collisions, especially at low energies. On the other hand,
the ratios from the heavy-ion collisions [55] are very well reproduced by the HRG model.
In both types of collisions, there is a dramatic increase in the n
p
/n
p
ratio [3, 56, 57]. At
LHC energy [3], while the differences between n
p
/n
p
ratios of pp and heavy-ion collisions
apparently almost disappear, their values approach 100% i.e., nearly 0% matterantimatter
asymmetry.
Int J Theor Phys
Fig. 1 n
p
/n
p
ratios depicted in
the whole available range of

s.
Open symbols stand for the
results from various pp
experiments (labeled). The
heavy-ion results from AGS, SPS
and RHIC, respectively, are
drawn in solid symbols. The solid
curve represents results from
thermal HRG model at Q=0.5
and =1.0, which apparently
perfectly describes all heavy-ion
collisions, besides the ALICE pp
results. At LHC energy, the ratios
are very much close to unity
indicating almost 100%
baryonantibaryon symmetry
Fig. 2 n
K
/n
K
+ ratios given
in the whole available range of

s in heavy-ion collisions of
AGS, SPS and RHIC,
respectively. The solid curve
represents HRG results. At RHIC
energy, the ratios reach 93%,
whereas at LHC energy very
much close to unity indicating an
almost 100% antibosonboson
symmetry. The values predicted
at ALICE energies are marked by
upwards arrows
In Fig. 1, the solid line gives the results from the HRG model, Sect. 2. No tting has
been utilized. Over the whole range of

s, the two parameters and Q remain constant,
1.0 and 0.5, respectively. HRG obviously describes very well the heavy-ion results [10, 16,
17, 19, 3537]. It is interesting to notice that the ALICE pp results are also reproduced by
means of the HRG model. Therefore, the solid line seems to be a universal curve describ-
ing the behavior of antiproton-to-proton ratios, while

s raises from few GeV to couple
TeV. Furthermore, it is clear that the LHC results indicate an almost vanishing net-baryonic
matter.
In Fig. 2, the ratios n
K
/n
K
+ are given in dependence on

s ranging from AGS to LHC.


Once again in this bosonic ratio, there is a drastic raise enhancing the accessibility of nega-
tive against positive kaons with increasing

s [5865]. At RHIC energies, the symmetry of


these bosons approaches 93%, which is obviously higher than the baryonic ratio in Fig. 1
(73%). Same behavior is also observed at lower energies. At the same energy, the values
of n
K
/n
K
+ are higher than the values of n
p
/n
p
implying that the accessibility of bosons is
relatively much higher than that of baryons. Comparing the two gures illustrates these dif-
ferences. In pp collisions, there are few measurements at low energy [66]. Therefore, their
comparison with the heavy-ion collisions would be incomplete. As shown in the previous
gure, the HRG model is not suitable to be compared with pp results. At LHC energy, the
Int J Theor Phys
kaon ratio approaches very much close to 100% indicating a nearly fully antibosonboson
symmetry.
4 Cosmological and Astrophysical Consequences
It is essential to mention potential similarities between heavy-ion collisions and Big Bang
cosmology [67]. There is a generic similarity between early stages of the early Universe
and those of the heavy-ion collisions [68], which apparently reects the explosive start
and successive phases, as well [69]. The formation of QGP in relativistic heavy-ion col-
lisions would constrain even the inhomogeneous cosmologies [70]. It is conjectured that the
rst-order phase transition might take place in heavy-ion collisions and/or in lattice QCD
simulations. Such a prompt transition seems to have essential astrophysical consequences.
Despite of the order of phase transitions, the QGP erain early Universeseems not to be
followed by an extreme expansion (ination). This is apparently the case in heavy-ion col-
lisions, because of the baryon number conservation and the limitation of baryon-to-photon
ratio (n
b
n

b
)/n

10
11
[1]. Therefore, n
p
n
p
recently measured by ALICE experi-
ment at 7 GeV can be used to estimate photon number density, n

5.5 10
4
, while in
CMB era, n

411.4(T/2.73 K) cm
3
. The partonic is followed by hadronization through
thermal and chemical equilibrium. Therefore, the QGP era seems to be the last symmetry
breaking of strongly interacting matter. With this statement, we mean deconnement and
chiral symmetry breaking and/or restoring. Going back to the analogy with CMB, it is not
restricted to the baryon-photon ratio. The quantum uctuations from the early Universe are
conjectured to be explained by WMAP [1]. The temperature-correlations in the heavy-ion
collisions have been simulated in hydrodynamic models for Au nuclei (12 fm across) [71].
A power-spectrum from the heavy-ion data has been extracted and compared with CMB.
In an isotropic and homogeneous background, the volume of the Universe seems to be
related directly to the scale parameter a(t ), where t is the co-moving time. Implement-
ing barotropic equation of state for the background matter makes it possible to calculate
among othersthe Hubble parameter H(t ) = a(t )/a(t ). One dot refers to the rst derivative
with respect to time t . Focusing the discussion about the QCD era of early Universe, which
likely turns to be fairly accessible in the high-energy experiments, the equation of state is
very well dened, experimentally. In Ref. [72, 73], a viscous equation of state for QGP
matter has been introduced. Different solutions for the evolution equation of H have been
worked out. For simplicity, we can assume that the deviation of H in viscous background
matter from the one characterizing the ideal matter can perturbatively be given as [74]
H(t ) =H
0
(t ) +f (t ), (7)
where H
0
(t ) is the Hubble parameter corresponding to a vanishing bulk viscosity. The pa-
rameter will be given in (10). In this treatment, is taken a perturbative variable. The
barotropic equation of state of viscous QCD matter (QGP matter) are
P =, (8)
T =
r
, (9)
=, (10)
besides an expression for the relaxation time, . The quantities P, , and T are pressure,
energy density, bulk viscosity and temperature, respectively. The parameters , , and r
are to be determined from the high-energy experiments and/or the lattice QCD simulations
Int J Theor Phys
Fig. 3 The time evolution of the
Hubble parameter H(t ) at
different values of the
perturbative parameter for bulk
viscosity, . The vertical lines
determine the co-moving time t
and temperature T boundaries of
the QCD era in early Universe,
which can be assigned to the
LHC energy
[10, 3538]. The function f (t ) has been evaluated [74]. The resulting H(t ) is depicted in
Fig. 3. The vertical lines dene the temperatures available to the QGP matter according to
heavy-ion experiments and lattice QCD simulations and simultaneously the time (in 1/GeV
units) according to the Big Bang cosmology. By doing this, the volume V and scale factor
a in the early Universe can be roughly estimated.
The ratio of baryon density asymmetry to the photon density, , has been measured in
WMAP data [1]. Then n
p
n
p
from ALICE experiment at 7 GeV can be used to give an
estimation of the photon number density.
4.4 10
4
<n

< 7.7 10
4
. (11)
When introducing a model for the volume of the early Universe as given in Fig. 3 at ALICE
energies 7 TeV, the photon number, N

=n

V , can be calculated. It is obvious that the re-


sulting N

will be extraordinarily large indicating dominant radiation against all other equa-
tions of state in this era, where matter and antimatter are supposed to be almost symmetric. In
the CMB era, where T 2.73 K, the photon number density n

411.4(T/2.73 K) cm
3
.
In a co-moving volume V a
3
(t ), the number density of non-interacting photons likely
remains constant. Therefore
n

1/a
3
(t ). (12)
Nevertheless, the decrease of temperature, when the Universe expanded, would affect the
quantity a
3
(t )n

. As a consequence, a
3
(t )n

seems to change in the expanding Universe.


The values in CMB and in (11), where the latter should be related to an era characterized
by ALICE temperature or energy, can be taken as an indicator for varying gamma number
with cooling down the early Universe. There is a conserved quantity, namely the entropy
density s. In a perfectly closed system like the Universe, s likely remain unchanged. It can
be deduced from the rst-law of thermodynamics and (1) [72, 73]. At vanishing chemical
potential, (8) and (9), which are valid for the QGP matter, lead to
s(T ) =
P(T ) +(T )
T
=
+1

(T )
1r
, (13)
implying that s(T ) is related to
1r
(T ), where = 0.319, = 0.718 0.054 and r =
0.230.196. At low energy, where the dominant degrees of freedom are given by all hadron
Int J Theor Phys
Fig. 4 The relative abundant of
kaon and proton yields as
functions of

s in a loglog
chart. The net boson and proton
relative to their summation
decrease almost linearly with
increasing

s. Apparently, kaon
relative abundant is about one
order of magnitude faster than
that of proton
resonances, (1). The entropy density can be derived from (1). Doing this, we can study
baryon and boson relative abundant (n(T ) n(T ))/(n(T ) + n(T )). In Fig. 4, HRG results
for the relative abundant as a function of

s are given for both spices. We notice that the
kaon relative abundant is by about one order of magnitude aster than the proton. It is obvious
that the abundances of light elements (
7
Li,
4
He,
3
He and
2
H) produced in the early Universe
are sensitive indicators of the baryonic number density [75].
5 Conclusions
So far, it can be concluded that the future LHC heavy-ion program likely will produce
antiparticle-to-particle ratios similar, when not entirely identical, to the ones measured in
current pp program. Secondly, the ratios obviously run very close to unity implying al-
most vanishing matterantimatter asymmetry. Some astrophysical consequences of these
results are given in Sect. 4. On the other hand, it can also be concluded that the thermal
models including the HRG model seem to be able to perfectly describe the hadronization
at very large energies and the condition deriving the chemical freeze-out at the nal state
of the hadronization, the constant degrees of freedom or S(

s, T ) = 7(4/
2
)VT
3
, where
V (or V as in (1)) is the volume, seems to be valid at all center-of-mass-energies raging
from AGS to LHC. Furthermore, at very high energy, the initial conditions likely do not
affect the particle (antiparticle) production so that the ratios in pp equal the ratios in AA
collisions. The rate and amount of annihilation of antiparticle obviously do not differ in
both types of collisions. At the freeze-out boundaries, the particle (antiparticle) production
seems to be well described by the degrees of freedom. At LHC energy, either antiboson-
to-boson or antibaryon-to-baryon ratios are equal implying nearly vanishing net-bosons and
net-baryons. The vanishing net-bosons are much faster than the vanishing net-baryons.
The expression (4) and (5) can be used to calculate the antiquark and strange chemi-
cal potentials, respectively. Seeking for simplicity, we assume vanishing isospin chemical
potential
I
3
i.e., degenerate light quarks
n
p
n
p
exp(6
q
/T
ch
), (14)
n
K

n
K
+
exp(2
u
/T
ch
) exp(2
s
/T
ch
), (15)
Int J Theor Phys
Fig. 5 The baryochemical
potential
B
is depicted versus
the center-of-mass-energy

s in
a loglog chart. With increasing

s,
B
decreases almost linearly
indicating that the higher
collision energies, the smaller
energy required to insert or
remove (create or annihilate) new
particles
where T
ch
is the freeze-out temperature. Substituting the value of n
p
/n
p
measured by AL-
ICE experiment at 7 GeV and T
ch
=0.175 0.02 GeV [3] in (14) results is

q
0.179 MeV. (16)
Then, extracting a value for n
K
/n
K
+ at the same energy (7 TeV, for instance) from Fig. 2
and substituting it in (15) leads to

s
0.0038 MeV. (17)
The evolution of baryochemical potential
B
for the hadronic baryons with increasing

s
is illustrated in Fig. 5. The larger

s, the smaller
B
. The relations between
B
and the
quark chemistry is given in Sect. 2. At LHC energy,
B
is very small indicating a very small
difference between particle and antiparticle. This situation likely fairly simulates the early
Universe during the QCD era [7274, 7679].
The matterantimatter asymmetry is one of the greatest mysteries in modern physics.
According to the well-known Sakharovs conditions [80]; the baryon number violation and
C and CP violation besides the out-of-equilibrium processes are conjectured to give a solid
framework to explain why should matter become dominant against antimatter with the Uni-
verse cooling-down through expansion. Recently, a fourth condition has been suggested;
b l violation, where b and l being baryon and lepton quantum number, respectively. The
rst condition means different interactions of particles and antiparticles. The second condi-
tion means non-conservation of the baryonic charge. These two conditions are not given in
the standard model, where both b and l are conserved, explicitly. Non-perturbatively, b +l
can be violated through sphalerons, for instance, which drive the two-direction conversion
b l with a certain number of selection rules [81, 82]. That the heavy-ion program seems
to give systematic tools to survey the evolution of matterantimatter asymmetry from 0%
to 100% in the nuclear interactions would make it possible to devote such experimental
facilities to study the evolution of Sakharovs conditions with the center-of-mass-energy

s.
Phenomenologically, we need to nd an explanation for the different relative abundances
of bosons and baryons and for the very small charge asymmetry during early Universe. This
would be the rst step in phenomenological studies for Sakharovs conditions. Although all
effective models including thermal ones and the lattice QCD thermodynamics apparently as-
sume an equilibrium system, thermal out-of-equilibrium processes have to be implemented.
Int J Theor Phys
Acknowledgements This work is partly supported by the Science and Technology Development Fund
(STDF) under the GermanEgyptian Scientic Projects (GESP) grant ID-1378. It is a pleasure to thank Prof.
Hagen Kleinert for the stimulating discussion and the friendly hospitality in the Research Center for Einstein
Physics, while part of this work was carried out. The author likes to thank ALICE collaboration for providing
the latest measurement of antiproton-to-proton ratios at 0.7 and 9 TeV.
References
1. Bennett, C., et al. (WAMP Collaboration): Appl. J. Suppl. 148, 1 (2003)
2. Tawk, A.: AIP Conf. Proc. 1115, 239247 (2009)
3. Aamodt, K., et al. (Alice Collaboration): Phys. Rev. Lett. 105, 072002 (2010)
4. Tawk, A.: arXiv:1011.5612 [hep-ph]
5. Jeon, S., Koch, V.: Phys. Rev. Lett. 83, 5435 (1999)
6. Stephanov, M., Rajagopal, K., Shuryak, E.: Phys. Rev. D 60, 114028 (1999)
7. Sollfrank, J., Heinz, U.W., Sorge, H., Xu, N.: J. Phys. G 25, 363 (1999)
8. Cleymans, H., Oeschler, J., Redlich, K.: J. Phys. G 25, 281285 (1999)
9. Torrieri, G., Jeon, S., Rafelski, J.: Phys. Rev. C 74, 024901 (2006)
10. Karsch, F., Redlich, K., Tawk, A.: Eur. Phys. J. C 29, 549556 (2003)
11. Zhang, Q.H., Topor Pop, V., Jeon, S., Gale, C.: Phys. Rev. C 66, 014909 (2002)
12. Bialas, A., Hwa, R.C.: Phys. Lett. B 253, 436 (1991)
13. Hegyi, S., Csorgo, T.: Phys. Lett. B 296, 256 (1992)
14. Cleymans, J., Redlich, K.: Phys. Rev. C 60, 054908 (1999)
15. Braun-Munzinger, P., Stachel, J.: Nucl. Phys. A 606, 320328 (1996)
16. Tawk, A.: Europhys. Lett. 75, 420 (2006). hep-ph/0602094
17. Tawk, A., Toublan, D.: Phys. Lett. B 623, 4854 (2005)
18. Magas, V., Satz, H.: Eur. Phys. J. C 32, 115119 (2003)
19. Tawk, A.: Nucl. Phys. A 764, 387392 (2006)
20. Cleymans, J., Stankiewicz, M., Steinberg, P., Wheaton, S.: nucl-th/0506027
21. Tawk, A.: Prog. Theor. Phys. 126, 279292 (2011)
22. Tawk, A.: arXiv:1007.4585 [hep-ph]
23. Tawk, A.: Fizika B 18, 141150 (2009). 0805.3612 [hep-ph]; hep-ph/0602094
24. Heinz, U., Subramanian, P.R., Greiner, W.: Z. Phys. A 318, 247 (1984)
25. Koch, P., Muller, B., Stocker, H., Greiner, W.: Mod. Phys. Lett. A 3, 737 (1988)
26. Ellis, J., Heinz, U., Kowalski, H.: Phys. Lett. B 233, 223 (1989)
27. Gavin, S., Gyulassy, M., Plumer, M., Venugopalan, R.: Phys. Lett. B 234, 175 (1990)
28. Sorge, H., Keitz, A.V., Mattiello, R., Stocker, H., Greiner, W.: Z. Phys. C 47, 629 (1990)
29. Rossi, G.C., Veneziano, G.: Nucl. Phys. B 123, 507 (1977)
30. Capella, A., et al.: Phys. Rep. 236, 225 (1994)
31. Kaidalov, A.B., Ter-Martirosyan, K.A.: Sov. J. Nucl. Phys. 39, 1545 (1984)
32. Labrador, A.W., Mewaldt, R.A.: Astrophys. J. 480, 480 (1997)
33. Ahlen, S.P., et al.: Phys. Rev. Lett. 61, 145148 (1988)
34. Gaisser, T.K., Mauger, B.G.: Astrophys. J. 252, L57L59 (1982)
35. Karsch, F., Redlich, K., Tawk, A.: Phys. Lett. B 571, 6774 (2003)
36. Redlich, K., Karsch, F., Tawk, A.: J. Phys. G 30, S1271S1274 (2004)
37. Tawk, A.: Phys. Rev. D 71, 054502 (2005)
38. Tawk, A.: J. Phys. G 31, S1105 (2005)
39. Venugopalan, R., Prakash, M.: Nucl. Phys. A 546, 718 (1992)
40. Busza, W., Goldhaber, A.S.: Phys. Lett. B 139, 235 (1984)
41. Vance, S.E., Gyulassy, M., Wang, X.N.: Phys. Lett. B 443, 45 (1998)
42. Vance, S.E.: Nucl. Phys. A 661, 230c (1999)
43. Gyulassy, M., Wang, X.N.: Nucl. Phys. B 420, 583 (1994)
44. Baier, R., Dokshitzer, Yu.L., Peigne, S., Schiff, D.: Phys. Lett. B 345, 277 (1995)
45. Braun-Munzinger, P., Heppe, I., Stachel, J.: Phys. Lett. B 465, 15 (1999)
46. Bearden, I.G., et al. (NA44 Collaboration): Phys. Rev. Lett. 78, 2080 (1997)
47. Ackermann, K.H., et al. (STAR Collaboration): Phys. Rev. Lett. 86, 402 (2001)
48. Ackermann, K.H., et al. (STAR Collaboration): Nucl. Phys. A 661, 681685 (1999)
49. Bjorken, J.D.: Phys. Rev. D 27, 140 (1983)
50. Levin, E.: J. Phys. Conf. Ser. 5, 127147 (2005)
51. Gyulassy, M., Pluemer, M.: Phys. Lett. B 243, 432 (1990)
Int J Theor Phys
52. dInterria, D., Betz, B.: Lect. Notes Phys. 785, 285339 (2010)
53. Chatrchyan, S., et al. (CMS Collaboration): Phys. Rev. C 84, 024906 (2011)
54. Adare, A., et al. (PHENIX Collaboration): Phys. Rev. C 77, 011901 (2008)
55. Adler, C., et al.: Phys. Rev. Lett. B 86, 4778 (2001)
56. Sickler, F., et al. (NA49 Collaboration): Nucl. Phys. A 661, 45c (1999)
57. Ahle, L., et al. (E802 Collaboration): Phys. Rev. Lett. 81, 2650 (1998)
58. Bearden, I.G., et al. (BRAHMS Collaboration): Phys. Rev. Lett. 94, 162301 (2005)
59. Bearden, I.G., et al. (BRAHMS Collaboration): Phys. Rev. Lett. 93, 102301 (2004)
60. Staszel, P., et al. (PHOBOS Collaboration): Nucl. Phys. A 774, 7792 (2006)
61. Alve, B., et al. (PHOBOS Collaboration): J. Phys. G 34, S11031108 (2007)
62. Adler, S.S., et al. (PHENIX Collaboration): Phys. Rev. C 69, 034909 (2004)
63. Laue, F., et al. (KaoS Collaboration): Phys. Rev. Lett. 82, 16401643 (1999)
64. Forster, A., et al. (KaoS Collaboration): J. Phys. G 28, 18951902 (2002)
65. Cassing, W., Bratkovskaya, E.L., Mosel, U., Teis, S., Sibirtsev, A.: Nucl. Phys. A 614, 415432 (1997)
66. Mitrovski, M., et al. (NA49 Collaboration): J. Phys. G 37, 094003 (2010)
67. Schramm, D., Olive, K.: Nucl. Phys. A 418, 289c (1984)
68. McLerran, L.: AIP Conf. Proc. 917, 219230 (2007)
69. Gelis, F.: Acta Phys. Pol. Suppl. 1, 395402 (2008)
70. Kajino, T.: Phys. Rev. Lett. 66, 125128 (1991)
71. Mocsya, A., Sorensen, P.: Nucl. Phys. A 855, 241244 (2011)
72. Tawk, A.: Can. J. Phys. 88, 822831 (2010)
73. Tawk, A., Magdy, H.: arXiv:1109.6469 [gr-qc]
74. Tawk, A., Wahba, M., Mansour, H., Harko, T.: Ann. Phys. (Berlin) 522, 912923 (2010)
75. Olive, K.A., Thomas, D.: Astropart. Phys. 7, 2734 (1997)
76. Tawk, A., Wahba, M.: Ann. Phys. (Berlin) 522, 928956 (2010)
77. Tawk, A., Wahba, M., Mansour, H., Harko, T.: Ann. Phys. 523, 194207 (2010)
78. Tawk, A., Wahba, M., Mansour, H.: Talk given at 12th Marcel Grossmann Meeting on General Rela-
tivity (MG 12), Paris, France, 1218 Jul 2009. arXiv:0912.0115 [gr-qc]
79. Tawk, A., Harko, T., Wahba, M., Mansour, H.: Invited talk at 7th International Conference on Modern
Problems of Nuclear Physics, Tashkent, Uzbekistan, 2225 Sep 2009. arXiv:0911.4105 [gr-qc]
80. Sakharov, A.D.: JETP Lett. 5, 24 (1967)
81. Kuzmin, V.A., Rubakov, V.A., Shaposhnikov, M.E.: Phys. Lett. 155, 36 (1985)
82. Riotto, A., Trodden, M.: Annu. Rev. Nucl. Part. Sci. 49, 35 (1999)

Potrebbero piacerti anche