2.1. The Phase Transitions in the Standard Model
In the SM of elementary particle interactions, the dynamics of fireball expansion is based on the asymptotic freedom property of underlying non-Abelian gauge theories [
31,
32]. QCD is a quantum non-Abelian field theory, an important part of the SM, that describes the fundamental interactions between colored quarks and gluons. The generalization of classical electrodynamics to non-Abelian gauge theories was first studied and exemplified in SU(2) by Yang and Mills in 1954 [
75]. The classical Lagrangian density of an SU(
N) gauge theory reads,
in terms of the field strength tensor defined in terms of YM fields
as
, where
are the structure constants of the SU(
N) group. Throughout,
denote internal indices of SU(
N) in the adjoint representation. Here, the parameter
is known as the YM coupling constant. Gauge theories based on SU(
N) are known as YM theories, and they became the target of a wider interest prompted by the discovery that massless particles may acquire a mass and a longitudinal polarization through spontaneous symmetry breaking (or Higgs) mechanism of a massless YM theory [
76,
77,
78]. The latter is a vital part of the SM framework realizing the classical Higgs mechanism of EW symmetry breaking that has found an excellent confirmation through the discovery of the Higgs boson [
79,
80].
In the framework of the quantum field theoretical approach, the YM field fluctuations are quantized around a given ground state through the first quantization procedure à la QED. However, due to self-interactions of the YM quanta, manifested via the term ∝
in the field strength tensor, the quantum YM ground state acquires, in general, a very non-trivial structure. This structure is well understood in the weakly coupled (perturbative) regime of the theory, implying
, which is the case of the EW theory or in the UV regime of QCD with the so-called asymptotic freedom of color charges. The strongly coupled (non-perturbative) regime, in which
that is realized in particular in the infrared limit of QCD, corresponds to the color confinement phenomenon and has remained the subject of active research over the last few decades. In recent years, significant progress has been made in understanding of the quantum YM ground state in SU(
N) gauge theories at finite temperatures, see e.g., Refs. [
81,
82].
As a result of a series of cosmological phase transitions that occurred in early universe during the first few microseconds after the Big Bang, new vacuum subsystems associated with breaking of the fundamental symmetries were formed. In the early universe, the SM predicts that cooling proceeds as a series of two phase transitions associated with the various spontaneous symmetry breakings of the corresponding gauge symmetries [
12,
68,
83,
84,
85,
86]. One at the temperature
GeV [
87] is responsible for spontaneous breaking of the EW symmetry providing masses to the elementary particles; see the left panel on
Figure 3. Due to the large value of its critical temperature
, it is not amenable to experimental study under the laboratory conditions. The second and the only one accessible in the laboratory, QGP-to-hadronic matter phase transition happening at
MeV [
88], is related to the spontaneous breaking of the chiral symmetry and manifesting itself in the massless quark limit of the QCD Lagrangian. Since both phase transitions are considered to be analytic crossovers, the bulk motion of the corresponding plasmas did not depart from thermal equilibrium. Therefore, such transitions, if realized in nature, are not expected to generate cosmological relics [
86,
89,
90] or to be helpful for a baryogenesis mechanism.
The QCD phase transition has occurred at characteristic temperatures of above
that correspond to a cosmological time-scale of above
and the Hubble length-scale of approximately
. The nature of the QCD phase transition is still a matter of intense debates in the literature [
74,
91,
92,
93,
94,
95,
96,
97,
98,
99,
100,
101,
102,
103], with results derived so far heavily relying either on lattice field theory methods applied to QCD [
92,
93], i.e., lattice QCD, or on holographic analyses of QCD at early cosmology [
103]. For a thorough review on various aspects of the QCD transition epoch, see e.g., Refs. [
12,
104,
105,
106,
107,
108].
It is undeniable that an abrupt QCD phase transition occurring reversibly in the early universe would lead to a promising cosmological scenario, according to which a large part of the quark excess would be condensed into invisible quark nuggets—a possible explanation for DM only relying on QCD. As Witten suggested in Ref. [
5], this would happen only if quark matter retains an energy per baryon which is less than 938 MeV: then, neutron stars might generate a quark matter component for cosmic rays, and detectable gravitational radiation could be produced during the QCD phase transition. Conversely, several recent studies drew a different conclusion, pointing toward the realization of second-order or crossover phase transition scenarios [
101,
102].
In a hot and dense QCD matter, the
and to some extent, depending on the temperature, also the
s quarks become nearly massless, and the QCD Lagrangian acquires an approximate chiral symmetry SU
SU
, with the number of massless quark flavors
(
) or 3 (
). At low
, the QCD vacuum becomes unstable, and this symmetry is spontaneously broken by
pairing. The corresponding order parameter
MeV
, known as the chiral quark condensate, gives rise to masses of light hadrons as well as to constituent masses of
quarks and to some extent also to the
s quark; see the left panel of
Figure 3. Recent lattice calculations with
and strange quark having its physical mass reveal that the chiral phase transition occurs at
MeV [
109].
Let us note that in addition to the above scenario when the thermal history of the universe proceeds by a sequence of phase transitions from a more symmetric to a less symmetric state of matter, there is also a possibility of the reverse evolution when part of the zero-temperature unbroken gauge group of the SM or other gauge theory might have been broken in the early universe by thermal effects. As first noted by Weinberg [
113] in the context of an
gauge theory, with decreasing
T, one may encounter a transition to a state of higher symmetry
. Within the minimal extensions of the SM containing an additional color triplet scalar field, the scenario in which the early universe underwent an epoch when SU
was spontaneously broken but later restored was analyzed in Ref. [
114]. The attractiveness of such a multi-step phase transition scenario stems from the fact that it may generate the observed baryon asymmetry of the universe [
115].
To describe the evolution of energy density
and entropy density
of the early universe, it is customary to normalize both quantities to their values
and
corresponding to an ideal massless Bose gas with a single degree of freedom (DoF) [
112,
116]
and call
and
the effective numbers of DoFs in energy and entropy, respectively. For the particular case of a non-interacting gas consisting of
Dirac fermions,
massive vectors,
massless vectors and
neutral scalars, the two functions are identical and read
where the prefactors account for the DoF of each of the considered particles.
It is worth mentioning that in a generic case of interacting (non-ideal) gas
and
, they are not identical and depend on temperature. Using the relationship
between the pressure
p, the entropy density
s, the energy density
and the generalized EoS parameter
w,
and the speed of sound
can be expressed in terms of the effective DoF measures
and
,
where the prime indicates differentiation with respect to temperature
T [
116]. It is worth mentioning that the causality condition between the speed of sound and the speed of light
induces the inequality
with an upper bound saturated for
corresponding to the case of absolutely stiff fluid.
Taking into account all permissible interactions in the SM, one can calculate either directly [
117] or on a lattice the temperature dependence of
and
and extract corresponding DoFs. This is illustrated on the right panel of
Figure 3 showing the temperature dependence of
in the SM. For realistic values of both
and
for a wide temperature interval from 10 keV to 10 TeV, see Ref. [
117].
The simultaneous presence of the EW and the QCD matter in thermal equilibrium is one of the remarkable differences between the QGP produced in accelerator experiments and the primordial QGP in the early universe [
13,
118,
119]. To find out which form of matter prevails, let us use Equation (
2) and compare the number of DoFs in an ideal massless gas consisting only of EW or QCD matter. Including only the particles which at the temperature
can be considered as massless, we obtain
DoFs for the EW case. The first term in the brackets corresponds to charged leptons
, and the second term corresponds to neutrinos
. The last term corresponds to photons. For non-interacting or weakly interacting QCD matter, Equation (
4) reduces to
where the first term accounts for the two spin and
color DoFs of the gluons and the second term accounts for
colors,
flavors, two spin and two particle-antiparticle DoFs of the quarks. Including only the quarks with the mass
we obtain successively
DoFs. In thermal equilibrium at the temperature
and for
active quark flavors, the QGP contains a factor of
more energy and pressure than those for the EW matter. For temperatures
, deep inside the EW era, all six quarks
can be considered to be massless, cf. the left panel of
Figure 3, and
. At the same time, the EW matter acquires
DoFs, where
are the DoFs of massless gauge bosons,
, and the last term is due to the Higgs scalar doublet. For this case, the QGP has a factor of
larger energy density and pressure than those of the EW matter. Hence, we conclude that the QGP was the most dense form of matter filling the early universe during both the QCD and EW epochs.
2.2. Creation of Primordial Black Holes during the Phase Transitions
According to inflationary theories, initially, very small inhomogeneities in the matter distribution were produced by the end of the exponential expansion regime. Such inhomogeneities filling the early universe are described by the metric perturbations
which can be decomposed into three irreducible pieces—scalar, vector and tensor ones, see, e.g., Refs. [
84,
120]. While the scalar part is induced by energy density fluctuations
, the vector and tensor perturbations are related to the rotational motion of the fluid and to the gravitational waves, respectively [
84]. Given the scope of this review, in the following, we focus only on one spectacular phenomenon related to metrics fluctuations in the early universe—the matter collapse into primordial black holes (PBHs).
Whereas their existence was proposed already a half-century ago first by Zeldovich and Novikov [
121] and later by Hawking [
122], it was the detection of gravitational waves from mergers of tens of solar mass
black hole binaries [
123] which has led to a surge of current interest in the PBHs as a Cold Dark Matter (CDM) candidate [
73,
124,
125,
126]. It can be shown that the creation of PBHs due to the gravitational collapse of hot and dense matter occurs for the density contrast
exceeding the critical threshold
, which generally depends on the EoS parameter
w [
73,
125]. Such large values of
can be generated, e.g., during a period of inflation in the very early universe [
126] or during an intermediate period dominated by long-lived massive particles (for recent work, see e.g., Ref. [
127] and references therein) or when the universe in the course of the phase transition passes a local minimum in the pressure-to-energy density ratio
[
125].
For the PBHs forming from Gaussian inhomogeneities with root-mean-square amplitude
, the present CDM fraction for PBHs with a mass around
M is found as [
125,
128]
where
is the fraction of horizon patches undergoing collapse to PBHs when the temperature of the universe is
T,
is the horizon mass at matter-radiation equality, and the numerical factor comes from the ratio of measured baryon
and CDM
abundances. In Equation (
10), we have explicitly taken into account dependence of the critical overdensity
on the EoS parameter
. The temperature depends on the PBH mass
M as
MeV. Note that the parameter
appearing in Equation (
10) can be always adjusted to counterbalance the theoretical uncertainties in the value of
so that the current PBH DM fraction is preserved [
128]. It is worth mentioning that in the scenarios where PBHs are formed during inflation, their abundance is larger than the Gaussian result by orders of magnitude, but also the mass function has a more pronounced tail at larger masses [
126].
In fact, there are a plethora of other mechanisms for PBHs formation (including besides the already mentioned FOPTs, bubble collisions, and the collapse of cosmic strings, necklaces, domain walls, non-standard vacua, etc., see e.g., the recent reviews [
73,
125]). In the following, in conformity with the topic of our review, we will concentrate on the softest point (SP) mechanism of creation of the PBHs discussed in Ref. [
128]. Its virtue stems from the fact that by tracing the origin of PBHs to the SM phase transitions, it is capable of explaining the provenance of part, if not all, of the CDM in the universe [
124].
The SP, corresponding to a local minimum in the pressure-to-energy density ratio
as given in Equation (
6), gives rise to elongation of the expansion time of the hot and dense matter. In HICs, the interest in locating the SP was started by the recognition that the longest-lived fireball might provide a clear signature of the QGP-to-hadron phase transition [
129]. Shortly after that, the formation of horizon-size PBHs due to a substantial reduction of pressure during adiabatic collapse in the course of the QCD transition was analyzed in the context of the early universe in Refs. [
130,
131]. Even though the previously used assumption of the first-order character of the phase transition was later on replaced by a crossover scenario, the lattice calculations have found a local minimum in
at
MeV [
132].
To become acquainted with the influence of the SPs on the cooling of the universe during its radiation-dominated era, let us follow Ref. [
128] and inspect the behavior of the function
shown on the right panel of
Figure 3. Let us focus on the temperatures when a part of the radiation matter transforms into a non-relativistic matter. Starting from
GeV downwards, this happens first to the top quark at
GeV, which is followed by the Higgs boson at 125 GeV and the
Z and
W bosons at 92 and 81 GeV, respectively. The fact that these particles become non-relativistic at nearly the same time of universe expansion induces a significant drop in the number of relativistic DoFs, from
down to
. Further changes at the
b,
c-quark and
-lepton thresholds are too small to be noticeable. Hence, further on,
remains approximately constant until the QCD transition at around 160 MeV. The number of relativistic DoFs then falls abruptly to
. A little later, pions become non-relativistic, and then muons, yielding
. Thereafter,
remains constant until
annihilation and neutrino decoupling at around 1 MeV, when it drops down to
[
128].
Provided that total entropy is conserved, an abrupt reduction of
leads to a sudden drop in the speed of sound
, cf. Equation (
7), and hence to a drop of pressure,
, cf. Equation (
6). The effect is clearly visible on the left panel of
Figure 4 showing the four periods in thermal history of the universe when
reaches its local minimum. After each period,
w returns back to its relativistic value of 1/3, but each sudden drop modifies the probability of gravitational collapse of any large curvature fluctuations present at that time [
128].
Consider one cooling period
with
centered around the local minimum
and define the quantity,
measuring departure from the
case; see Equation (
6). At the endpoints
but for
, it is always positive
, cf. Equation (
7). Hence, the initial drop in the entropy DoFs,
, always precedes the jump in the energy density DoFs,
. This leads to the following “coarse-grained" scenario for the PBH formation: the reduction in
occurring for
is followed by a fall in
for
. An excess in entropy
lost by the radiation during its cooling is dumped into the collapsing matter, emerging eventually in the form of PBHs—the matter with the largest entropy density in the universe [
133]. This may explain why even at the present stage of the universe evolution, there is by a huge factor far more entropy in supermassive black holes (BHs) in galactic centers than in all other sources of entropy put together [
134].
Assuming that the amplitude of the primordial curvature fluctuations is approximately scale-invariant [
128], one obtains from Equation (
10) the mass spectrum of PBHs
shown on the right panel of
Figure 4. The peaks at
and
correspond to the EW and QCD phase transitions, to pions and muons becoming non-relativistic and to
annihilation, respectively. The latter may also provide seeds for the supermassive BHs’ formation in galactic nuclei. The largest contribution to
comes from the PBHs formed at the QCD transition epoch and that would naturally have the Chandrasekhar mass (
) [
128]. Moreover, the peak in the range 1–
could explain the LIGO/Virgo observations [
123]. The latter favor mergers with low effective spins as expected for PBHs, but it is hard to explain BHs of stellar origin [
135].
The simple analytical models that describe the dynamical process of gravitational collapse which may be relevant for PBH formation were studied in Ref. [
136]. It is also worth noting that the gravitational collapse of large inhomogeneities during the quark-hadron transition epoch may also explain the baryon asymmetry of the universe [
137]. The asymmetry can be generated in local hot spots through the violent process of PBH formation at the QCD transition triggered by a sudden drop in the radiation pressure and the presence of large amplitude curvature fluctuations caused by the axion field—the subject to be discussed in
Section 2.6.
2.3. Perturbative and Strongly Coupled Regimes of QCD
An important contribution to the effective number of relativistic DoFs,
, comes from the hadron-to-QGP phase transition—see a big jump in the interval
MeV in
Figure 3 (right panel). At higher temperatures, in the QGP region, the strength of the interactions between the quarks and gluons is set by the QCD coupling
which at the one-loop order of perturbation theory takes the form,
where
Q is the momentum transferred during the interaction,
MeV is the characteristic QCD scale, and
is the number of active quark flavors. The logarithmic decrease of
with increasing
Q, i.e., with decreasing distance among the quarks and gluons, is due to the fact that, in contrast to the photon in QED, the force carriers in QCD, the gluons, have color charge. Their exchanges in higher-order processes involving both the quarks and the gluons occur more frequently with increasing
Q and lead to a color charge spread (or anti-screening). Indeed, the gluon multiplicity increases at low momentum fractions corresponding to the limit of large energies. Dilution of the initial color charge is responsible for the weakening of
at small distances
, i.e., when the quark experiences a large momentum transfer
Q, see Equation (
12). This effect known as asymptotic freedom [
31,
32,
138] is illustrated on the left panel of
Figure 5 where the values of the
extracted from proton-(anti)proton and lepton-proton collisions are shown [
139]. In agreement with Equation (
12), a slow logarithmic decrease from
to
is observed.
Before proceeding further, let us recall that quantum field theory (QFT) at finite temperature
T is often considered to be equivalent to Euclidean QFT in a space which is periodic, with period
along the “imaginary time" axis (for a recent review of this subject, see e.g., Refs. [
26,
140] and references therein). Thus, in order to formulate the theory at
using its variant at
, one should replace zero components of all 4-momenta
in the Euclidean integrals by the discrete Matsubara frequencies—
for bosons and
for fermions, and sum over
instead of integrating over
. Consequently, the average momentum transferred during the interactions in the hot medium
Q can be related to the temperature as
. In particular, the maximum value of the momentum transferred
, which has been so far measured in
collisions at the LHC, roughly corresponds to the “temperature”
.
The asymptotic freedom formula given by Equation (
12) is based on the applicability of QCD perturbation theory to the processes at high momentum transfers, as it is a well-known fact that the perturbative expansion is an example of asymptotic series. It looses its validity with decreasing
Q when the perturbative approximation breaks down. Interestingly, by performing the matching of the fundamental theory onto the effective chiral Lagrangian at the infrared scale
GeV, at which the ranges of validity of perturbative QCD and chiral perturbation theory (describing interactions among low-momentum hadrons) descriptions meet, one can infer the information about the behavior of
at large distances. Such a matching implies that the QCD coupling in the infrared region is “frozen" at
[
143], incidentally at twice the upper scale value of
shown on the left panel of
Figure 5.
Let us note that evolution of the strong coupling parameter
described at the leading order by Equation (
12) is valid only for DoFs that dominate the thermodynamical evolution, i.e., for the partons (quarks and gluons) with momenta of order
, and it does not apply to the long wavelength non-perturbative modes residing at the length scales of
. Those modes are occupied by a liquid in which neighboring “unit cells” are tightly coupled to each other [
144]. This strongly coupled regime [
10] makes the QGP behave as the ideal fluid [
145,
146]. The fluidity of the QGP was first established in the collisions of ultra-relativistic nuclei at RHIC [
53,
54,
55] and later confirmed at higher energies of the LHC [
147,
148,
149]. The most prominent signals of the strong interaction in the deconfined bulk manifest in a collective flow of matter [
8,
146] and in a spectacular phenomenon of suppression of very energetic partons passing through the QGP medium [
8,
150,
151]. Direct evidence for the non-perturbative character of deconfined matter comes from the low-momentum spectra of direct photons measured in Au+Au collisions at RHIC. The temperatures obtained from the spectra
MeV [
152] point to the initial temperatures
300–600 MeV at early times of
0.6–0.15 fm/
c, which are way below the perturbative regime of QCD.
By definition, plasma is a state of matter in which charged particles interact via long-range (massless) gauge fields [
153]. This distinguishes it from neutral gases, liquids, or solids in which the inter-particle interaction is of short range. So, plasmas themselves can be gases, liquids, or solids, depending on the value of the plasma parameter
, which is the ratio of interaction energy to kinetic energy of the particles forming the plasma [
39]. For a classical plasma of
N particles with charge
occupying a volume
V,
where
is the temperature-dependent particle number density. While most plasmas are ideal with
, a strongly interacting plasma has
. For plasma with
at the temperature
K
eV, the number density
n must be as high as
to make
[
39]. Ion plasma in a white dwarf has
= 10–200, in transient plasmas produced in explosive shock tubes, the values of
= 1–5 are found [
39]. A more down-to-earth example is table salt (NaCl), which can be considered as a crystalline plasma made of permanently charged ions
and
[
153]. At temperatures of
K, still too small to ionize non-valence electrons, the table salt transforms into a molten salt, which is a liquid plasma with
.
Generalization of Equation (
13) to the QGP case was suggested in Ref. [
154]
where the strong coupling
is a slowly varying function of temperature,
and
are the Casimir invariants of fundamental and adjoint irreducible representations of the color SU
group corresponding to quarks and gluons, respectively, and
is the average inter-parton distance at a given temperature
T as follows from Equation (
13). The factor 2 in Equation (
14) takes into account the equal importance of chromoelectric (CE) and chromomagnetic (CM) interactions in ultra-relativistic systems. For ideal massless QCD gas with
active quarks and
degrees of freedom, see Equation (
9), the particle number density reads
From Equations (
13) and (
15), it follows that
, and so, the term
, appearing in the denominator of Equation (
14), depends on
T through the temperature-dependent number of active quark flavours
only. Consequently, for an ideal massless QCD gas, the temperature-dependence of the plasma coupling parameter
is driven by
. For the QGP created in HICs at RHIC
MeV and
0.3–0.5 with
, Equations (
15) and (
14) yield
and
2–8 well inside the strongly coupled regime. At much higher temperatures, say, at
, with
and
, we obtain
and
0.5–1.5—the value located in the vicinity of the strongly coupled regime. At even higher temperatures, the number of active quark DoFs saturates at
and the evolution of
becomes solely driven by the (logarithmically decreasing) QCD coupling
, cf. (
12). Let us note that the ideal gas approximation serves only as a lower estimate of
because it ignores the interactions in the partonic liquid. The latter will slow down the temperature dependence of the average inter-parton distance
a, thus weakening the strong coupling parameter dependence on
T.
A more in-depth approach to strongly coupled non-Abelian plasmas [
155] was expected to come from the gauge/string duality [
156]—a correspondence between
d-dimensional conformal QFT and
-dimensional string or gravity theory. In these theories, the graviton needs not live in the same spacetime as the QFT, but due to the holographic principle, the description of gravity within a volume of spacetime can be thought of as encoded on a lower-dimensional boundary to the region in the formalism of conformal field theory [
157,
158]. However, the inherently conformal character of the gauge QFT used in the duality with anti-de Sitter gravity (AdS) is at variance with QCD where the scale invariance is broken by the confinement scale (for a recent review of confinement dynamics, see Ref. [
22]), causing the running of the coupling. This limits the applicability of gauge/string duality [
156] to temperatures
and hence to weak couplings.
2.4. QCD at High Parton Densities and Saturation
To proceed further, a different type of analysis of the QCD dynamics of partonic matter is needed. There are two independent paths along which the density of partons can evolve, and these are illustrated in the right panel of
Figure 5. The two together form the basis of our current understanding of high-energy scattering in QCD.
The first path follows the development of partonic cascade in variable
Q. For partons that occupy a transverse area
, the increase of
Q and hence of the temperature
leads to dilution of their density. The process is controlled by the Dokshitzer–Gribov–Lipatov–Altarelli–Parisi (DGLAP) [
159,
160,
161] equations describing the evolution of partonic density as a function of evolution variable
[
17,
162].
The second path follows the development of a parton shower in variable
, which is a fraction of the light cone momentum
1 of the parent parton, which has radiated a parton emerging with the light-cone momentum
. In the
plane transverse to the direction of the fast-moving primary parton, the partonic cascade initiated by the primary parton can be visualized as a Brownian motion-like trajectory developing from
toward
. The corresponding Gribov diffusion process is controlled by the so-called evolution parameter
, leading to a difference in rapidity between the primary and radiated partons, with a coefficient being the diffusion constant proportional to
. Its evolution in the
Y variable is described by the Balitsky-Fadin-Kuraev-Lipatov (BFKL) equation that is complementary to the DGLAP evolution realized in
(see standard textbooks, e.g., Refs. [
17,
162] and references therein).
At fixed
Q, the radiated partons (mostly, soft gluons with
) are typically of the same size. When a parton-parton interaction cross-section ∼
multiplied by
—the probability to find at fixed
Q a parton carrying a fraction
x of the parent parton momentum—becomes comparable to the geometrical cross section
of the object
A occupied by the gluons, the partons “overlap”. The repulsive interactions among the gluons ensure that their occupation number
(the number of gluons with a given
x multiplied by the area each gluon fills up divided by the transverse size of the object) saturates at
. Note that this is a very generic behavior—the same density scaling as the inverse interaction strength
is characteristic of a number of condensation phenomena such as the Higgs condensate, see, e.g., Ref. [
163], or superconductivity [
164].
The phenomenon of saturation [
165] is thus important for gluons with transverse momenta
[
166,
167], where
is the
x-dependent saturation scale representing a fixed point of the parton density evolution in
x or, equivalently, the emergent “close packing” scale [
167]—see the right panel of
Figure 5 where the saturation line
,
is also displayed. Such gluons form a highly coherent configuration called Color Glass Condensate (CGC) [
142,
167,
168], or glasma [
169], which due to the high occupation number
has properties of QCD in the classical regime [
166].
At high temperatures, one usually expects that quantum effects become less important. To show that, for the CGC, we follow Ref. [
166] and write the gluonic part of the QCD action in terms of the gauge field potential
and field strength
which are obtained by rescaling of their equivalents
and
used in a more traditional approach when the coupling constant
multiplies the interaction terms in the Lagrangian,
where
with
are the SU(3) group structure constants. For a classical configuration of gluon fields, by definition,
does not depend on the coupling, and the action is large,
. The number of quanta in such a configuration is then
where we rewrote Equation (
18) as a product of four-dimensional gluon condensate density
and spacetime volume
. The number of quanta
in such a configuration depends only on the product of the Planck constant
ℏ and the strong coupling squared
. The classical limit
is indistinguishable from the weak coupling limit
[
166]. Thus, the weak coupling limit of small
corresponds to the semi–classical regime.
An equivalent argument employs the fact that the path integral formulation of the quantum theory in Minkowski space sums over all field configurations weighted with
. Since
appears in the exponential in the same place as
ℏ, cf. Equation (
18), this already suggests that for
, the path integral is dominated by the classical configurations. Such configurations are believed to describe the matter inside incident nuclei during the initial stage of relativistic HICs at RHIC and the LHC [
142,
168].
Although in its original formulation, the QCD saturation is used for partons with fixed
Q in case of the macroscopic bodies, it is more relevant to consider the partons at a fixed temperature
, avoiding at the same time the quantum entanglement problem [
170], which is inevitably present in the description of microscopic objects. In this generalized setting, an object
A filled with gluons may represent not only a fast-moving proton or nucleus but also the interior of the expanding early universe. Moreover, as follows from Equation (
16), the saturation phenomenon is not necessary related to the growth of the gluon density at small
x. For a big fast-moving domain of space filled with deconfined quarks and gluons, with radius
fm, and hence also for the fast-expanding early universe itself characterized by the Hubble horizon
fm, the saturation limit can be reached even at
[
162]
2. In the extreme case, when the gluonic part of QCD matter completely decouples from the QCD fermionic fields and forms the vacuum condensate, the first term in Equation (
9) can be neglected, and we obtain
, making the CGC a prevailing form of matter during both the QCD and EW epochs.
Thus, in the periods of cosmological evolution when
, including a very hot QCD era, EW era and beyond, it is perfectly conceivable that the universe was dominated by the fully saturated gluonic matter with occupation number
. If during its subsequent cooling, the universe followed a trajectory in the
plane staying still above the
line, see the right panel of
Figure 5, the CGC phase would be a prevailing form of matter down to the temperatures
. For lower temperatures, the glasma is expected to fragment into a strongly interacting QGP.
The issue of emergence of classical behavior in the cosmological history has drawn recently a great deal of attention because of its conceptual as well as practical importance, see e.g., Ref. [
171] and references therein. Although the origin of the observed anisotropies in the cosmic microwave background (CMB) radiation is traced back to vacuum fluctuations of quantum fields in the very early universe [
84,
120,
172], there is a general expectation that the main characteristics of the universe can be described in classical terms even in its early history [
171,
173,
174]. This is consistent with the fact that the initial conditions of the Hot Big Bang were determined by cosmic inflation driven by the so-called inflaton field [
84,
86,
175,
176]. Apart from the quantum fluctuations of this field, its very emergence may well be a quantum phenomenon, e.g., the axion condensation (for a more thorough discussion, see below).
Let us note that the word “glass” appearing in the acronym CGC is used in condense matter physics to describe a non-equilibrium, disordered state of matter acting like solids on short time scales but liquids on long time scales [
177]. In the glasma case, there are two scales present—the light cone time
of low
x (or wee) partons and the light cone time
of primary (or valence) partons. For partons of transverse momentum
, we obtain
suggesting that the valence parton modes are static over the times scales over which the wee modes are probed [
167]. It is quite tempting to identify
with the quantum break-time discussed in Refs. [
178,
179] defined as the time-scale after which true quantum evolution of the parton densities departs from the classical mean field evolution.
Glasses are formed when liquids are cooled too fast to form the crystalline equilibrium state. The fast cooling leads to an enormous number of possible configurations
into which the glasses can freeze and consequently to their large entropy
, which does not vanish, even at
, see e.g., Refs. [
177,
180]. In case of the CGC, the fast cooling is expected to take place in the Grand Unified Theory (GUT) era (see
Section 2.5) when about one-third of all gauge bosons are gluons. By the end of that period, the excess of effective entropy DoF
is almost completely absorbed by the saturated gluonic matter.
The gluon condensation into (many domains of) the saturated phase was also facilitated by the fact that the wee partons “see” the color charge of other gluons over very large distances given by their transverse wavelength
. Since the glasma domains were formed in separate and completely different gluonic configurations, the saturated gluonic matter occupying the early universe had a substantial excess in the entropy DoF,
, over the effective number of DoF in energy,
. Consequently, the value of the generalized EoS parameter
w was higher than that of the ideal massless gas; see Equation (
6).
One possibility of how the EoS of a non-equilibrium matter comprising weakly interacting gluons can be approximated by the ideal massless gas of quasi-particles with
follows from Equation (
48) discussed later in
Section 3.1. The glasma with
may be looked upon as either a two-dimensional sheet (
), a one-dimensional string (
) or, more generally, a fractal with the Hausdorff dimension
. Recent investigations of the dynamics of expanding glasma show that the spatial asymmetry introduced by the initial geometry is effectively transmitted to the azimuthal distribution of the gluon momentum field, even at very early times [
181,
182].
2.5. The Running Couplings of the Standard Model and Their Unification
The importance of QCD interactions in the EW era of the universe evolution can be also seen by comparing the corresponding couplings—the strong
, electromagnetic
and weak
ones. This is illustrated on the left panel of
Figure 6 showing the RG flow in the scale
of the electromagnetic, weak and strong coupling parameters above
GeV. Note that due to the fact that the gauge group of SM interactions SU
× SU
× U
is not a simple Lie group, the theory has not one but three coupling parameters, which are often denoted as
,
and
[
183].
To see how the values of the coupling parameters influence the state of early universe, let us compare the collision time among its constituents
(
is the effective cross-section,
n is the particle number density and
v is their relative velocity) with the characteristic time-scale of the universe expansion
[
84]. Let us first restrict ourselves to the temperatures
when all particles of the SM are ultra-relativistic and the gauge bosons are massless. Then, the cross-sections for strong and EW interactions have a similar energy dependence
, where
are the corresponding dimensionless running couplings
varying logarithmically with
T; see
Figure 6. Taking into account Equations (
39) and (
42) for
, we find
Thus, for the temperatures GeV , the local equilibrium in the fluid is established before expansion of the universe becomes relevant.
At the lower temperatures down to
MeV, the strong interactions prevail over the EW ones, and the local equilibrium is controlled by the interactions among the quarks and gluons in the strongly interacting QGP liquid. Even though the number density of the medium may be smaller
3, the big effective cross-section among the particles forming the medium guarantees that the condition for local equilibrium
is satisfied. Thus, over that whole range of the temperatures, the early universe is in the local equilibrium. It develops along the maximal possible entropy path, making it amenable to the hydrodynamical description.
In spite of its success, the SM cannot be the ultimate theory of particle physics. Such long-standing problems as the absence of a suitable DM candidate, no explanation of the observed baryon asymmetry in the universe, as well as various hierarchy problems in the underlined mass spectra (such as the flavor problem, the neutrino mass problem and the Higgs hierarchy problem) call for various bottom-up extensions of the SM framework as well as continuous attempts to derive the SM structure from a top-down perspective.
As was earlier discussed, e.g., in Ref. [
184], the Higgs boson quartic coupling in the SM turns negative at scales ≳
GeV, rendering the vacuum state of the theory unstable at high energies. The current theoretical developments and experimental measurements suggest that the metastability of the Higgs vacuum is favored. This means a vacuum decay may occur with possibly catastrophic consequences for cosmology, since there are many catalysts that could trigger such a decay in the early universe. For a comprehensive review on cosmological implications of the Higgs vacuum metastability, see, e.g., Ref. [
185].
Incidentally, almost immediately after the discovery of the asymptotic freedom, it was suggested [
186] that at very high energies, the three gauge interactions of the SM are merged into a single force. The model, a first example of the Grand Unified Theory (GUT), is based on the smallest simple Lie group which contains the SM gauge groups SU(5) ⊃
× SU
× U
. Among its 24 gauge bosons, there are in addition to eight gluons of QCD and four EW gauge bosons
and
also 12 new ones called
X and
Y. Their emission or absorption makes it possible to transform a lepton into a quark or vice versa. Hence, the SU(5) GUT does not conserve baryon and lepton numbers separately, making it the first theory providing an explicit mechanism for the proton decay
, with the half-time
years, where
is the mass of SU(5) gauge boson at the scale of Grand Unification and
is the mass of the proton. Although it was later found to disagree with experimental lower limit of
years [
187] SU(5), unification is still considered an important example and a reference point of GUT model-building.
The basic property of the SU(5) theory and its later GUT successors [
188,
189,
190,
191,
192] is that by virtue of the unification into a single (simple Lie) gauge group at very high energies, strict unification of the SM gauge couplings must take place. This is hinted at but not really achieved in the SM; see the left panel of
Figure 6. First,
and
cross each other at
GeV; then,
crosses
at
GeV, and finally,
and
cross at
GeV, providing a hint of unification close to the Planck scale. Thus, as ultraviolet (UV) completions of the SM describing the physics at very high energies, GUTs are sometimes connected to theories of gravity such as string theory.
Let us now try to speculate on how the Grand Unification might have influenced the dynamics of the early universe at the temperatures when gluons decouple from quarks, forming a condensate. In SU(5) GUT, with increasing temperatures , the gluon exchange between quarks (antiquarks) becomes overshadowed by the exchange of EW massless gauge bosons to be finally superseded by the GUT super-heavy X and Y gauge bosons. The latter also facilitate the conversion of quarks into leptons and vice versa. Since the transformations occur in the thermal and chemical equilibrium, the number of quarks and leptons remains constant on average.
One of the fundamental issues with the GUT models, which remains a challenge today, is the large hierarchy between the mass scale of Grand Unification and the EW scale. The latter is a source of large loop corrections to the Higgs mass. The problem is usually solved by extending the GUT with supersymmetry, the hypothetical symmetry between fermions and bosons (for a recent review on the concepts of supersymmetric GUTs, see e.g., Ref. [
192]), which is broken at lower energies. Some of its minimal realizations, such as the Minimal Supersymmetric Standard Model (MSSM) [
193], predict the unification of all three gauge couplings at the same scale; see the right panel of
Figure 6.
Notwithstanding the drawbacks of GUTs, their cosmological signatures look quite promising. With the critical temperature
of the GUT phase transition approximately of the same order of magnitude as the particle masses at that temperature, the phase transitions that take place in GUTs at
GeV, as a rule, prove to be FOPTs [
194]. Such transitions proceed via bubble nucleation [
195]. While an isolated spherical bubble may produce GWs through sound waves in the plasma and magneto-hydrodynamics turbulence effects (see for example Refs. [
196,
197,
198,
199] and references therein), the process of bubble collision contributes to the GWs spectrum in the quadrupole approximation [
86,
200]. This contrasts with the SM phase transitions where the crossovers do not lead to a strong enhancement over the primordial GW spectrum. Moreover, the FOPTs also generate a primordial magnetic field during the turbulence phase of the plasma and bubble collision [
86,
201], and in some instances, they may generate topological defects such as domain walls and strings [
195,
202].
Let us add that an FOPT in the EW sector though precluded in the SM is possible in many of its scalar sector extensions [
203]. In the most exotic scenario, a very peculiar history of the universe may occur: a first-order QCD phase transition (with six massless quarks) triggers an EW FOPT, which is eventually followed by a low-scale reheating of the universe where hadrons (likely) deconfine again, before a final, conventional crossover QCD transition to the current vacuum [
204].
2.6. Axions
A promising avenue connecting the strong interactions with physics beyond the SM having at the same time far-reaching consequences in cosmology is provided by hypothetical ultra-light particles—the axions [
183,
205,
206,
207,
208]. The QCD, unlike the EW interaction, is symmetric under time reversal and hence under a combined charge conjugation C and parity P operation CP. In principle, one can add to the QCD Lagrangian the term:
where
is the gluon field strength tensor (see Equation (
17)), and
is its dual. For a non-zero value of the parameter
, called the vacuum angle [
209], the strong coupling permits violation of the CP symmetry. However, since
can be written as a total derivative of
, the Chern–Simons current,
, the new term does not produce any effects in perturbation theory and is therefore usually neglected. Nevertheless, classical configurations, topological in nature, do exist, one example being the instantons [
210], for which this term cannot be ignored. For instance, in the semi-classical dilute instanton gas approximation (DIGA) [
211], the QCD vacuum energy density
depends on
as
Here,
C is a positive constant and
is the QCD instanton action [
207,
208], cf. Equation (
18). Moreover, since
in Equation (
22) preserves the charge conjugation C, it contributes directly to the neutron electric dipole moment
, where
e is the proton charge,
denotes the mass of
quarks, and
is the neutron mass. Current measurements [
212] provide an upper bound on the CP-violation parameter,
.
Within the SM, the smallness of the
parameter becomes a true fine-tuning problem. Since
could acquire an
contribution from the observed CP-violation in the EW sector (via the common quark mass phase, arg det
, where
is the quark mass matrix), its not obvious why it becomes cancelled to a high precision by the (unrelated) gluon term [
207]. To solve this problem, the SM is augmented with an extra pseudo-scalar particle called axion
A, whose only non-derivative coupling is to the CP-violating topological gluon density
that is suppressed by a large scale
. With
, where
is the angular DoF of spin-zero complex field,
the minimum of the vacuum energy occurs when the coefficient
in front of
vanishes.
It is worth noting that interactions of the scalar field with ordinary matter is controlled by the factor
. Thus, even at its originally suggested value of
GeV at the EW breaking scale [
213], the axions interact so weakly that they emerge without attenuation from reactor cores or stellar interiors. Present astrophysical constraints push
to substantially higher values, which are somewhere between a few
GeV and a few
GeV.
In cosmology, since their introduction, the axion-like particles were considered to be potentially important candidates if not for all than at least for the main component of the DM—a form of matter accounting for about one-quarter of its total energy density [
72]. In order to fulfill their mission, the axions must contribute a non-negligible amount to the energy density of the universe and should have not been in thermal equilibrium with the cosmological plasma at any time in the history of the universe. This, together with the smallness of axion mass, implies their large occupation numbers [
207]—the situation already encountered when we have discussed the properties of saturated gluon matter, cf. Equation (
19). This implies that the axions over the whole history of the universe can be modeled by solving the classical field equations of a scalar condensate [
214].
Starting from the Peccei–Quinn (PQ) scalar field
introduced in Equation (
24), the Lagrangian invariant under the global
transformation reads [
215]:
Focusing on the early universe, the evolution of the field
passes the following milestones. At high temperatures
, the effective potential
depicted on the left panel of
Figure 7 has the
symmetric minimum at
. With increasing time
t, the universe cools, and the vacuum with
becomes unstable. Due to the misalignment mechanism, the field starts to roll down from
, and the potential becomes tilted. At
, the PQ symmetry is spontaneously broken—the field acquires the vacuum expectation value
. Then, the axion—the Nambu-Goldstone boson of the spontaneously broken
symmetry—becomes a massless angular DoF at the minimum of the potential. The detailed results depend on whether the PQ phase transition occurs before or after inflation. While in the former case, only one
angle contributes (all other values are inflated away), in a post-inflationary scenario, the initial value of the angle
takes all values in the interval
. Eventually, when the universe cools to the temperatures of a few GeV, the axion obtains a mass through the QCD non-perturbative instanton effect known as the axial anomaly [
183,
216,
217].
At the leading order in
, the axion mass
at some temperature
T can be extracted from the QCD-generating functional
in the presence of a theta term [
218],
where
is the QCD topological susceptibility. This quantity is typically computed using the lattice methods developed by several groups, see e.g., Refs. [
89,
219,
220,
221,
222,
223,
224,
225,
226,
227,
228], but also in the framework of analytical approaches [
218,
229]. Its temperature dependence can be extracted either from the DIGA or from the LQCD calculations; see the right panel of
Figure 7. The lattice simulations performed, for instance, in Ref. [
89] have revealed that for
MeV, the susceptibility falls as
with
, extending thus the previous DIGA result
up to
GeV. At the temperatures of
MeV in the vicinity of the QCD chiral phase transition temperature
, see
Section 2.1,
flattens. Further analysis performed in Ref. [
89] exploiting the QCD EoS obtained therein has revealed that in the post-inflationary scenario, depending on the fraction of DM consisting of axions,
50–1500
eV. In particular, for the 50% axion content of the DM, one obtains the axion mass scale of
eV. Since at
, the chiral perturbation theory [
183] becomes applicable, it is worth making a comparison with the original formula for the axion mass,
where
is the pion decay constant, and
and
are the down- and up-quark masses appearing in the QCD Lagrangian [
183,
208]. Let us note that if
eV, the axions decay faster than the age of the universe.
For temperatures above the chiral phase transition, the axion potential computed in DIGA reads [
211]:
Expanding Equation (
28) around
, we obtain
at a finite
T. Assuming the Friedmann-Lemaître-Robertson-Walker (FLRW) metric and classical axion field with this potential, the corresponding unperturbed energy density
and pressure
due to the axion field read
respectively. Substituting
and
from Equation (
29) into the fluid Equation (
39), we obtain
At early times
, we can neglect
in Equation (
30) to obtain the solution
—the axion field is frozen at a constant value
.
With increasing time
t, eventually, the oscillating term proportional to
in the equation of motion of the axion field (
28) begins to contribute. At the time
defined implicitly as
, the universe is sufficiently large to host a sizeable fraction of one oscillation period—the axion field starts to oscillate with an amplitude damped by the expansion rate. A solution of Equation (
30) then reads [
120],
where
is the time at which
, i.e., when the temperature drops below the QCD chiral phase transition temperature
,
is the constant and
is the phase. In particular, for
and a radiation-dominated universe, one obtains
and
[
120].
For the initial conditions at the onset of oscillations
,
, where
is called the initial misalignment angle, we obtain from Equation (
30)
where
,
,
and
are the values at the PQ symmetry breaking time
. In the second equation, we have also used
and
[
208].
While in the pre-inflationary scenario, inflation selects one patch of the universe within which the spontaneous breaking of the PQ symmetry leads to a homogeneous value of the initial misalignment angle
, in the post-inflationary scenario, the PQ symmetry breaks with
, taking different values in patches that are initially out of causal contact; see the left panel of
Figure 7. However, today, they populate the volume enclosed by our Hubble horizon. In the post-inflationary scenario, the initial misalignment angle
takes all possible values on the unit circle. For a quadratic potential,
, this is equivalent to an assumption that the initial condition reads
, where the angle brackets represent the value averaged over
[
208].
Starting from
, the number of axions in a comoving frame becomes frozen, and their number density evolves as
For isotropic evolution, the ratio of the number density
to the entropy density
s in the comoving frame is conserved, i.e.,
leading for
to the expression for the axion energy density,
where the effective number of DoFs of the entropy
is defined in Equation (
3). Let us note that in contrast to Ref. [
208] in Equation (
34), the constancy of the axion mass for
is already taken into account.
After the spontaneous
symmetry breaking the axion field,
, being an angular variable, takes values in the interval
, cf. Equation (
24). Consequently, the axion potential
given by Equation (
28) is periodic in
with period
. In the other words,
has an exact
discrete symmetry. The axion acquires a periodic potential with
equivalent minima. The Kibble mechanism [
195,
230] then dictates that, depending on the homotopy group
of the manifold
of degenerate vacua, the topological defects—domain walls, strings or monopoles—form each time the symmetry is broken [
195,
202]. With
and
, the production of axionic strings, which are vortex-like topological defects that form as soon as the symmetry is spontaneously broken, is possible. Those are not important when the PQ symmetry is broken before inflation—they are inflated away—but they play an important role in the post-inflationary scenario. When the Hubble parameter
H becomes comparable to the axion mass
, the axion starts to roll down to one of the minima. Since the axion field settles into different minima in different places of the universe, domain walls are formed between the different vacua; see the left panel of
Figure 7. It is worth mentioning that this phenomenon is similar to ice formation on the surface of a pond or a puddle when the water begins to freeze in many places independently, and the growing plates of ice join up in random fashion, leaving zigzag boundaries between them [
195].
As an example, let us consider a planar wall orthogonal to the
z-axis
. The solution of the classical field equation with potential given by Equation (
28) reads [
231]:
This configuration interpolates between the two allowed vacua, at and , which are separated by the wall of thickness .
Astrophysical signatures of the axion can be broadly divided into its couplings to elementary/composite particles, i.e., photons, electrons, protons or neutrons, and to the macroscopic objects in the universe such as BHs [
208]. In the latter case, when the Compton length of the axions becomes of order of the BH size, they form gravitational bound states around it. The phenomenon of superradiance [
232,
233] causes the axion occupation numbers to grow exponentially, providing a way to extract very efficiently energy and angular momentum from the BH. The presence of axions could be inferred by observations of BH masses and angular momenta. Current measurements exclude the region of
GeV.
In particle physics signatures, the most important is the decay channel of axion into two photons with the decay width
in terms of the model-dependent coupling constant
. The main uncertainty is due to the electromagnetic and color anomalies of the axial current associated with the axion. The most relevant process induced by the
is the Primakoff process—the conversion of thermal photons in the electrostatic field of electrons and nuclei into axions,
A strong bound on exotic cooling processes in the sun is provided by the helioseismological considerations [
234] giving the following constraint:
. For more details, consult Refs. [
205,
206,
207,
208,
215]. Another option is the inverse Primakoff scattering, which allows solar axions to coherently scatter off the atomic electric field and back-convert into photons in the detector volume,
, proceeding through a
t-channel photon exchange [
235].
The axion besides being a well-recognized DM candidate may provide also an interesting explanation of a wider range of phenomena related to the early universe dynamics. As a first example, let us mention the SMASH model—a minimal extension of the SM with additional particle content comprising three sterile right-handed neutrinos
, a color triplet
Q and a complex SM-singlet scalar
, whose VEV of
GeV breaks the lepton number and the PQ symmetry simultaneously [
236]. At low energies, the model reduces to the SM, which is augmented by seesaw-generated neutrino masses and mixing, plus the axion. In this scenario, the inflaton, a scalar field driving cosmic inflation in the very early universe, is a mixture of
and the SM Higgs fields. The reheating of the universe after inflation occurs via a mechanism known as a Higgs portal [
237]—by the DM particles, which interact only through their couplings with the Higgs sector of the theory. The model provides a consistent picture of particle physics from the EW scale to the Planck scale
and of cosmology from inflation until today. In particular, in the SMASH model framework, the PQ symmetry is first broken and then restored non-thermally during preheating for
GeV.
The second example, the Axion Quark Nugget (AQN) DM model, see, e.g., Ref. [
238] and references therein, replaces the commonly accepted baryogenesis scenario with a charge separation process in which the global baryon number of the universe remains zero at all times. Similarly to Witten’s idea of stranglets [
5], the AQN DM is composed of quarks and anti-quarks but now in a new high-density CSC phase. Initially, nuggets of both matter and antimatter are formed with equal probability as a result of the dynamics of the axion domain walls which at the same time provide the extra pressure needed to stabilize the CSC phase. Later on, due to the global CP violating processes associated with the initial misalignment angle
during the early formation stage, the populations of the nuggets with the positive and negative baryon number become different. The unobserved antibaryons hidden inside the DM would not participate in nucleosynthesis and, therefore, according to the usual definition would not contribute to the visible matter. However, since antimatter nuggets can interact with regular matter via annihilation leading to electromagnetic radiation, their existence has observational consequences [
238].
It is worthwhile mentioning here an interesting generalization of the PQ mechanism where in addition to
angle, also the strong coupling
is promoted to a dynamical quantity. The latter evolves through the VEV of a singlet scalar field that mixes with the Higgs field [
239,
240]. In the resulting cosmic history, the QCD confinement and EW symmetry breaking initially occur simultaneously close to the weak scale.
To conclude this section, as we have already noticed in several examples above, the non-perturbative dynamics of QCD often exhibits very non-trivial and rather unexpected consequences at cosmological scales. An attractive mechanism proposing that the QCD axion may emerge as a composite state has been discussed very recently in Ref. [
241]. In particular, it was suggested that Majorana neutrinos, that combine into Cooper pairs, can form collective low-energy degrees of freedom. This motivates the existence of the QCD axion as a collective excitation of the neutrino condensate. Such a condensate can be produced after the QCD phase transition epoch in a cold coherent state by means of a misalignment mechanism, thus providing an alternative DM candidate. In this case, a QCD anomalous portal provides the necessary means for a tiny mass gap generation by neutrinos. Furthermore, the Cosmological Constant emerges as a result of the spontaneously broken mirror symmetry of the QCD ground state triggered by the quantum gravity effects as suggested by Refs. [
65,
66] (see also below). Hence, one concludes that QCD may be responsible for the dynamical generation of both the DM and DE components of the universe such that a complete knowledge of the QCD in the infrared regime may be absolutely critical for understanding of the cosmological evolution since the latest QCD transition epoch and for eventual formation of the current state of the universe.