You are on page 1of 19

Shear viscosity of a hadronic gas mixture

Kazunori Itakura,1 Osamu Morimatsu,1, 2 and Hiroshi Otomo2


1
High Energy Accelerator Research Organization (KEK), Oho 1-1, Tsukuba, Ibaraki, 305-0801, Japan
2
Department of Physics, University of Tokyo, 7-3-1 Hongo Bunkyo-ku Tokyo 113-0033, Japan
(Dated: November 7, 2007)
We discuss in detail the shear viscosity coefficient η and the viscosity to entropy density ratio η/s
of a hadronic gas comprised of pions and nucleons. In particular, we study the effects of baryon
chemical potential on η and η/s. We solve the relativistic quantum Boltzmann equations with binary
arXiv:0711.1034v2 [hep-ph] 16 Nov 2007

collisions (ππ, πN, and N N ) for a state slightly deviated from thermal equilibrium at temperature
T and baryon chemical potential µ. The use of phenomenological amplitudes in the collision terms,
which are constructed to reproduce experimental data, greatly helps to extend the validity region
in the T -µ plane. The total viscosity coefficient η(T, µ) = η π + η N increases as a function of T and
µ, indirectly reflecting energy dependences of binary cross sections. The increase in µ direction is
due to enhancement of the nucleon contribution η N while the pion contribution η π diminishes with
increasing µ. On the other hand, due to rapid growth of entropy density, the ratio η/s becomes
a decreasing function of T and µ in a wide region of the T -µ plane. In the kinematical region we
investigated T < 180 MeV, µ < 1 GeV, the smallest value of η/s is about 0.3. Thus, it never violates
the conjectured lower bound η/s = 1/4π ∼ 0.1. The smallness of η/s in the hadronic phase and its
continuity at T ≃ Tc (at least for crossover at small µ) implies that the ratio will be small enough
in the deconfined phase T > ∼ Tc . There is a nontrivial structure at low temperature and at around
normal nuclear density. We examine its possible interpretation as the liquid-gas phase transition.

I. INTRODUCTION is not found, but interestingly enough, the values of η/s


extracted from RHIC experiments [3] and from lattice
simulations [4, 5] seem to be small enough and close to
Shear viscosity of a hot QCD matter has been attract- the lower bound.
ing much attention in recent years. The major reason
for that is the intriguing experimental discovery that On the other hand, there is an important empirical ob-
the matter created in heavy-ion collisions at Relativis- servation which can be seen in many substances such as
tic Heavy Ion Collider (RHIC) in Brookhaven National helium, nitrogen, and water: The ratio η/s has a min-
Laboratory could be close to a perfect fluid [1]. This un- imum at or near the critical temperature [6] (see also
expected result has driven people to think about strongly- [7]). More precisely, the ratio shows a cusp at Tc for the
interacting quark-gluon plasma (abbreviated as “sQGP”) first order transition, while it has a convex shape for the
which may be realized at temperature just above the crit- crossover with its bottom around the (pseudo) critical
ical temperature T > ∼ Tc . As a result of strong-coupling temperature. Since this behavior is observed in many
nature, sQGP is thought to have very small shear viscos- substances, it is expected to be universal. Recall that
ity, which is however not directly confirmed yet in a sat- the phase transition in QCD is most probably crossover
isfactory way. In fact, only a few things are understood at least for low densities. Therefore, what we naturally
about sQGP because we will not be able to investigate expect for the QCD matter from the two observations
it within standard perturbative QCD techniques. The mentioned above is that the ratio η/s will have the min-
only technique available now (except for lattice simula- imum at T ∼ Tc , and the numerical value at that point
tions which are not analytic methods) is the one based will be close to the KSS bound η/s ∼ 0.1. It is of primary
on the AdS/CFT correspondence, which relates strongly- importance to check whether this expectation is indeed
coupled supersymmetric Yang-Mills theories to weakly- the case or not, and to understand the properties of the
coupled gravity theories. QCD matter around Tc not only for T > ∼ Tc . These con-
siderations motivated us to investigate the shear viscosity
There is an interesting outcome from the AdS/CFT
in QCD from the hadronic phase T < ∼ Tc . Notice that we
analysis in relation to the shear viscosity: It has been con-
can indirectly study the properties of sQGP from below
jectured that there would be a lower bound in the “shear
Tc because physical quantities such as the ratio η/s will
viscosity coefficient to the entropy density ratio” (or sim-
be continuous at Tc for the crossover transition.
ply the “viscosity to entropy ratio”) η/s ≥ 1/4π [2]. We
call this “the KSS bound” after authors’ names of Ref. [2]. Transport properties of a meson gas have been stud-
The lowest value η/s = 1/4π is satisfied by several super ied by several people. Many of the calculations are
Yang-Mills theories in the large Nc limit (strong coupling based on the Boltzmann equations with the Chapman-
limit), which suggests that the bound could be universal. Enskog method which is a standard approach for weak
Of course, there is no guarantee for this bound to hold dissipative phenomena, especially for computing trans-
in real (non-supersymmetric) QCD whose gravity dual port coefficients [8]. Differences among several papers
2

[9, 10, 11, 12, 13] include kinetic or statistical proper- we should do is rather evident: for the purpose of study-
ties of particles (relativistic or nonrelativistic, quantum ing the µ dependence of the shear viscosity η and the
or classical), species of mesons (pions, kaons, etc.), and ratio η/s in a pion-nucleon gas, we treat the relativistic
the cross sections in the collision terms. For example, quantum Boltzmann equations with binary cross sections
Ref. [12] treated a nonrelativistic quantum pion gas with which are determined to reproduce experimental data.
the binary cross section given by the leading order chiral We are very careful about the range of validity of our
perturbation theory (LO-ChPT). Recently, similar prob- framework. We also check whether the valley structure
lems have been revisited in relation to the KSS bound found in Ref. [16] indeed exists even with the phenomeno-
[14, 15]. In Ref. [14], the ratio η/s computed with the logical cross sections. It is also important to compare our
cross section in LO-ChPT turned out to violate the KSS results with those from hadron cascade simulations. For
bound for temperature beyond Tc ∼ 170 MeV, and it was example, the shear viscosity coefficient is computed for
speculated that such violation could be related to the ex- a meson gas in Ref. [17] and for a meson-baryon gas in
istence of phase transition. However, soon after that, it Ref. [18].
was shown in Ref. [15] that the KSS bound is not violated The paper is organized as follows: in the next section,
in a pionic gas if one computes the shear viscosity with we explain the relativistic quantum Boltzmann equations
a phenomenological cross section using the experimental for a dilute pion-nucleon gas, and define the cross sec-
phase shifts. What we have learned from these papers tions we use in the collision terms. We treat the small
is the following: (i) we have to be careful when we use deviation from the thermal equilibrium to the linear or-
the cross section from effective field theories, and (ii) the der (the Chapman-Enskog method), and give the shear
ratio η/s of a relativistic pion gas is small enough at rel- viscosity coefficient η through the solutions to the Boltz-
atively large temperature T < ∼ Tc , but does not violate mann equations. In Sect. III, we present our numerical
the KSS bound η/s ≥ 1/4π. results for a pure pion gas and a pion-nucleon gas mix-
At this point, there comes a natural question: how ture. We introduce a criterion which measures the valid-
does η/s change if one adds nucleons to the pure pion ity region of the calculations. The use of phenomenolog-
gas? Naively, we expect that the ratio η/s will decrease ical cross sections is very important to enlarge the range
as number of nucleons is increased because the pion cross of validity. We discuss in detail the effects of chemical
section will effectively enhance in the presence of nucle- potential on η and η/s and examine the interpretation
ons, yielding smaller shear viscosity, while the entropy of the valley structure as the liquid-gas phase transition.
will increase. Thus, there is a chance that the ratio could Summary is given in the last section.
violate the KSS bound if the bound does not change. No-
tice that the pion-nucleon gas is the minimum requisite
which allows us to study the effects of baryon chemical II. KINETIC THEORY OF A HADRONIC GAS
potential µ. Therefore, it is quite interesting and impor- MIXTURE
tant to investigate the µ dependence of η/s in the pion-
nucleon gas. Such investigation will also urge people to A. Quantum Boltzmann equations
study the (possible) µ-dependence of the KSS bound. In
fact, the “universality” of the KSS bound has not been
We first explain our theoretical framework which is
tested at finite baryon chemical potential.
necessary for computing the shear viscosity coefficient.
Most recently, the authors of Ref. [14] have applied
Consider a dilute gas of pions (π) and nucleons (N ) in
their framework to the pion-nucleon gas to study the be-
which particles interact with each other through binary
havior of η/s in the T -µ plane [16]. However, their fo-
collisions. Nonequilibrium processes such as relaxation
cus was not on the KSS bound but on the new finding:
to thermal equilibrium can be described by kinetic equa-
a valley structure in η/s at low temperature and large
tions for one particle distribution functions f π (x, p, t)
chemical potential which they argued would correspond
and f N (x, p, t) (below, we suppress x-dependence). For
to the nuclear liquid-gas phase transition. Although this
simplicity, we assume that the gas is isospin symmetric,
is a very interesting suggestion, their results should be
and thus f π (p, t) and f N (p, t) are isospin averaged distri-
critically checked since they are based on the effective
butions. The relativistic quantum Boltzmann equations
field theories whose validity region is severely limited.
(more precisely, the Uehling-Uhlenbeck equations) of this
On the other hand, a realistic calculation of the viscos-
hadronic gas mixture are then given by
ity in the pion-nucleon gas was performed some time ago
by Prakash et al. [10]. Remarkably, they used the bi- pµ
∂µ f π (p) = C ππ [f π , f π ] + C πN f π , f N ,
 
nary cross sections in the collision terms, which roughly π
(1)
Ep
reproduce experimental data. However, unfortunately,
the dependence on baryon chemical potential was not in- pµ
∂µ f N (p) = C N N f N , f N + C N π f N , f π , (2)
   
N
Ep
vestigated in detail. Besides, this calculation is based
on classical Boltzmann equations, and thus cannot be q
applied to relatively large chemical potential where the where Epπ,N = m2π,N + p2 and the collision terms are
effects of Fermi statistics is expected to be large.
defined as
In view of the present situation mentioned above, what
3


Z n o
ππ πN
dΓππ f1π f2π (1 + f3π ) 1 + fpπ − (1 + f1π ) (1 + f2π ) f3π fpπ

C +C =
2
Z n o
dΓπN f1N f2π 1 − f3N 1 + fpπ − 1 − f1N (1 + f2π ) f3N fpπ ,
  
+ gN (3)
gN
Z n o
CNN + CNπ = dΓN N f1N f2N 1 − f3N 1 − fpN − 1 − f1N 1 − f2N f3N fpN
   
2
Z n o
+ gπ dΓN π f1π f2N (1 + f3π ) 1 − fpN − (1 + f1π ) 1 − f2N f3π fpN .
 
(4)

We have used shorthand notation fiπ,N ≡ for π and N ,


f π,N (ki ), fpπ,N ≡ f π,N (p) and gπ , gN are the de-
d3 p
Z
generacy factors gπ = 3, gN = 2. For the collisions Eπ
 
ln Zπ = −V gπ 3
ln 1 − e− T , (8)
between the same species (ππ, N N ), we have added a (2π)
factor 1/2. The factors (1 + f π ) and (1 − f N ) represent Z 3
d p  E N −µ

the Bose-Einstein and Fermi statistics of particles, ln ZN = V gN 3
ln 1 + e− T , (9)
(2π)
respectively. Finally, dΓ in the integrants are invariant
measures with the scattering amplitudes squared: For one obtains the total entropy density s = sπ + sN as
example, follows:
1 ∂
(2π)4 δ (4) (k1 + k2 − k3 − p) Y d3 ki sπ = T ln Zπ
dΓπN ≡ |MπN |2 , V ∂T
(2E1N )(2E2π )(2E3N )(2Epπ ) i (2π)3 ( )
dp 2 Eπ
Z

 
(5) = −gπ p ln 1 − e T − Eπ
,
2π 2 T (e T − 1)
where MπN is the elastic scattering amplitude for
N (k1 ) + π(k2 ) → N (k3 ) + π(p). Explicit form of the 1 ∂
scattering amplitudes will be shown at the end of this sN = T ln ZN
V ∂T
section. These are the basic ingredients of the kinetic
( )
dp 2 EN − µ
Z
E N −µ
 
theory. =gN p ln 1 + e T + .
2π 2 E N −µ
T (e T + 1)
Before we discuss how to solve the Boltzmann equa-
tions (1) and (2), we need to know the equilibrium states.
We define them without solving the full Boltzmann equa-
B. Shear viscosity coefficient
tions: They are given by the distributions f0π , f0N which
make the collision terms vanish. Namely, C ππ [f0π , f0π ] +
C πN [f0π , f0N ] = C N π [f0N , f0π ] + C N N [f0N , f0N ] = 0. These The shear viscosity coefficient is defined through the
conditions are easily satisfied by the following Bose- deviation of spatial components of the energy momentum
Einstein and Fermi-Dirac distributions if the common tensor in the linear order with respect to fluctuation from
temperature T = 1/β and hydrodynamic velocity V µ the equilibrium. Consider a nonequilibrium state which
are used is slightly deviated from the global equilibrium. Small
deviation of the space components of energy momentum
1 tensor (T ij = T0ij + δT ij ) can be divided into traceful
f0π (p) = µ , (6) and traceless parts:
eβVµ p − 1
1 δT ij ≡ ζ (δ ij ∇ · V) − 2η ∇i V j trl ,

(10)
f0N (p) = β(V pµ −µ) , (7)
e µ +1 ij
1   δ
∇i V j trl ≡ ∇i V j + ∇j V i −

∇ · V, (11)
2 3
where µ is the baryon chemical potential. The parame-
ters T , V µ and µ can, in principle, depend on the coor- where V i is a space component of the hydrodynamic four
dinates (local equilibrium), but when we compute quan- velocity (which is common for π and N )
tities in thermal equilibrium, we simply select the rest
µ
frame V µ = (1, 0, 0, 0) so that Vµ pµ = E. d3 p Epπ,N f π,N (p)
R
Vµ = .
In evaluating the entropy density, we use the expres-
R
d3 pf π,N (p)
sion in the equilibrium state as is done in the literature
because the deviation from the equilibrium is assumed to Eq. (10) is the definition of the shear and bulk viscosity
be small. Namely, by using the grand partition functions coefficients η and ζ. The flow vector V i , as well as T
4

π,N
and µ, is in principle arbitrary and can depend on spa- where W(n) (p) is a polynomial of the order n, and bπ,N
(n) is
tial coordinates. But below we consider the case where the coefficient independent of p. Notice that W(n) π
(p) and
only the flow vector V i depends on the coordinates, and N
W(n) (p) are not equivalent to each other. Indeed, they
in particular, its divergence is vanishing: ∇i V i = 0 and
are defined so that they satisfy the following different
∇i V j 6= 0 (i 6= j). This situation corresponds to consid-
ering only the shear viscosity coefficient η and ignoring orthogonal conditions:
all the other transport coefficients such as bulk viscosity Z
d3 p f0π (1 + f0π ) 2 π π
or heat conductivity. p W(n) (p)W(m) (p) = δnm Lπ(n) ,
(2π)3 Epπ
If one knows distributions f π,N (p) = f0π,N (p) + (17)
δf π,N (p) as the solutions to the Boltzmann equations
(1) and (2), one can explicitly compute the (deviation d3 p f0N (1 − f0N ) 2 N
Z
N
of) energy momentum tensor: p W(n) (p)W(m) (p) = δnm LN(n) .
(2π)3 EpN
Z
d3 p pi pj π (18)
ij
δT = gπ δf (p) The normalization factors Lπ(n) and LN are not chosen
(2π)3 Epπ (n)
to be 1. Instead, we choose the polynomial so that the
d3 p pi pj N
Z
+ gN δf (p). (12) coefficient of the term with the highest degree is unity
(2π)3 EpN (such polynomials are called ”monic”). Namely, the first
three polynomials have the following form:
As mentioned above, we consider only the deviation
δf π,N that originates from the shear. Then, it is quite W(0) (p) = 1 ,
convenient to parametrize δf π and δf N as follows (p̂i = W(1) (p) = p + c1 ,
pi /p):
W(2) (p) = p2 + d1 p + d2 .
δf π ≡ −f0π (1 + f0π ) β B π (p)

δij
 Parameters c1 , d1 and d2 are uniquely determined by
∇i V j trl ,

× pˆi pˆj − (13) the orthogonal conditions (thus independent of the dy-
3
namics). In fact, W(n) has n unknown parameters, which
δf N ≡ −f0N (1 − f0N ) β B N (p) are uniquely determined by n orthogonal conditions with
lower polynomials (W(m) , m = 0, · · · , n − 1). Thus, Lπ(n)
 
δij
∇i V j trl ,

× pˆi pˆj − (14)
3 and LN (n) in Eqs. (17) and (18) are known after we com-
pletely determine W(n) . Practically, the expansion (16)
where we have introduced new quantities B π,N (p) to be
is well approximated by the first few terms. Thus, in the
determined by the Boltzmann equations.
present analysis, we take only the first three terms:
Substituting Eqs. (13) and (14) into Eq. (12) and com-
paring the result with Eq. (10), one finds the shear vis- π
B π (p) ≃ bπ(0) + bπ(1) W(1) π
(p) + bπ(2) W(2) (p) , (19)
cosity coefficient η as a function of unknown functions N
B π,N (p): B (p) ≃ bN
(0) + bN N
(1) W(1) (p) + bN N
(2) W(2) (p) , (20)
π,N
gπ β d3 p f0π (1 + f0π ) 2 π
Z
where we have used the definition W(0) = 1. The coef-
η = p B (p)
15 (2π)3 Epπ ficients bπ,N
(n) are numerically determined. Once we know
gN β d3 p f0N (1 − f0N ) 2 N these coefficients, we can compute the shear viscosity co-
Z
+ p B (p) . (15) efficient η from Eq. (15). We have checked that the re-
15 (2π)3 EpN
sults do not change even if we take up to the fourth terms
The unknown functions B π,N (p) (or equivalently, the de- (n ≤ 3).
viations δf π,N ) are numerically determined by solving Two comments are in order about our formulation.
the Boltzmann equations that are linearized with respect Firstly, we recall that the expansion of B π,N (p) in
to δf π,N . This procedure corresponds to the lowest order Eq. (16) is a familiar technique in solving the Boltzmann
Chapman-Enskog method. Here we discuss only the out- equation by the Chapman-Enskog method. If one treated
line of the procedure to solve the Boltzmann equations. a classical Boltzmann equation, measure of the orthogo-
More details are discussed in Appendix B. nal condition would be given by a much simpler distribu-
After the linearization, the Boltzmann equations (1), tion, the Maxwell-Boltzmann distribution. In this case,
(2) become a coupled linear equations for B π (p) and the polynomials that satisfy the orthogonal condition are
B N (p). Following Ref. [12], we solve these equations in given by famous functions, the Sonine polynomials [19].
the functional space spanned by the orthogonal polyno- For the quantum Boltzmann equation, however, the mea-
mials {W(n) (p), n = 0, 1, 2, ...}. Let us expand B π,N (p) sure in the orthogonal condition is given by either the
by these bases: Fermi-Dirac or Bose-Einstein distribution, as shown in
∞ Eqs. (17) and (18). In this case, the polynomials satisfy-
bπ,N π,N ing them are not known, and we have to find them order
X
B π,N (p) = (n) W(n) (p) , (16)
n=0 by order.
5

Secondly, note that the shear viscosity coefficient η de- The isospin averaged πN scattering amplitude in the
fined by Eq. (15) may be formally written as Center-of-Mass (CM) frame is
n o
η = ηπ + ηN . (21) |MπN |2 = (2mN )2 4|g− |2 + 2qcm
4
sin2 θcm |h+ |2 , (23)

It apparently looks ‘additive’ with respect to each con- where qcm and θcm are, respectively, the magnitude of
tribution. Indeed, if one considers a pion gas system pion momentum, and the scattering angle in the CM
without nucleons, one finds exactly the same expression frame. Two functions g− and h+ are
as the first term of Eq. (15). One might then be tempted 2
to conclude from the expression (21) that the inclusion gA 1 ω
2ω 2 − 2m2π + t + 2 ,

g− = − 2
of nucleons always contributes to enhance the value of fπ 4ω 2fπ
shear viscosity. However, such argument does not make 2
gA 1
sense because the function B π (p) itself will change by the h+ = − 2 ,
fπ 2ω
inclusion of nucleons. We will see later that η π indeed de-
creases as the effects of nucleons become large (i.e., with where gA = 1.26 is the nucleon axial charge, and ω is
increasing chemical potential). the pion energy in the CM frame. Note that the overall
factor in Eq. (23) is not the same as in Ref. [16] though
both are based on Ref. [21]. This is because we have
C. Scattering amplitudes of binary collisions used the standard normalization factor for the spinors
ūr (p)us (p) = 2mN δ rs instead of ūr (p)us (p) = {(E N +
Let us show the explicit expression of the scattering mN )/2mN }δ rs which was adopted in Ref. [21].
amplitudes in the collision terms (3), (4), and (5). Note Lastly, let us consider the N N scattering. In the low
that we have assumed factorization of the scattering am- energy EFT, scattering amplitudes in the CM frame for
plitudes from the products of one-particle distributions in fixed spin and isospin can be parametrized in terms of
the collision terms. This is physically natural in a dilute the scattering length a and the effective range r as
gas where each collision is simply treated as an indepen- 1
dent binary collision (ignoring higher order multiparticle |MN N ′ |2 = 64π 2 s · , (24)
| − a−1 + 21 rqcm
2 − iq
cm |
2
correlations). Therefore, as long as we consider a dilute
gas system where the Boltzmann equation is applicable, where qcm is the magnitude of nucleon momentum in the
it is reasonable to use the amplitudes for two particle CM frame. For each process, parameters a and r are
scatterings in the vacuum (free space). This is a great determined to fit the low energy experimental data with
merit in computing transport coefficients. If one follows the contribution from Coulomb force removed. Their nu-
the microscopic Kubo formula to compute the transport merical values are shown in table I. They are taken from
coefficients, it is quite nontrivial how to include the ef- Ref. [22]:
fects of physical processes. On the other hand, if one
uses the Boltzmann equation, one can easily incorporate
the physical cross sections in the collision terms, which TABLE I: Scattering length a and effective range r for N N
is however less justified from the first principle. In the collisions.
present paper, we examine two different parametrizations system parameter S = 0, I = 1 S = 1, I = 0
pp a −17.1 (fm) —–
for the scattering amplitudes. One is based on theoretical
r 2.79 (fm) —–
calculations, while the other is constructed from the ex-
nn a −16.6 (fm) —–
perimental data. More precisely, we use the amplitudes r 2.84 (fm) —–
from the low energy effective field theory (EFT) on the np a −23.7 (fm) 5.42 (fm)
one hand, and the phenomenological amplitudes designed r 2.73 (fm) 1.73 (fm)
to reproduce experimental data of the elastic scatterings
on the other hand.
Scattering amplitudes based on EFT are the following.
First of all, the isospin-averaged ππ scattering amplitude In low energy scattering, the dominant contribution
is given by the leading order Chiral Perturbation Theory to the amplitude is given by the s-wave (orbital angular
(LO-ChPT) [20]: momentum ℓ = 0). Thus, we construct the spin-isospin
averaged scattering amplitude from the expression (24)
1 n o for the s-wave with appropriate weight factors of spin and
|Mππ |2 = 21m 4
π + 9s2
− 24m 2
π s + 3(t − u)2
, (22) isospin. Details of the spin-isospin average are explained
9fπ4
in Appendix A.
where fπ is the pion decay constant (fπ = 93 MeV), s, t In summary, the scattering amplitudes based on EFT
and u are the Mandelstam variables for the scattering are given by Eqs. (22), (23), and (24). It should be em-
π(k1 ) + π(k2 ) → π(k3 ) + π(p). phasized that all these expressions of the scattering am-
Next, for the πN scattering, we use the results of low plitudes are valid only in a limited region of kinemat-
energy effective theory (LO heavy baryon ChPT) [21]. ics. LO-ChPT (for ππ case) is usually considered to be
6

90 Boltzmann equations will also be restricted to a small


80 π-π region in the T -µ plane, as we will discuss later.
With this limitation in mind, it is quite important to
70
consider physical scattering amplitudes in order to check
cross section (mb)

60
the usefulness of, or more importantly to go beyond,
50 the low energy EFT. To this end, we have constructed
40 phenomenological amplitudes from the experimental data
30 (differential cross sections) by fitting the coefficients of
20
the partial-wave
√ expansion. The fit was performed to the
data up to √ s = 1.15 GeV (qcm = 550 MeV) in the ππ
10
scattering, s =√2.00 GeV (qcm = 770 MeV) in the πN
0
200 300 400 500 600 700 800 900 1000 1100 scattering, and s = 2.04 GeV (qcm = 405 MeV) in the
s1/2 (MeV) N N scattering. Details of the fitting procedure are dis-
160
cussed in Appendix A. In Fig. 1, we compare elastic cross
sections of the two different parametrizations. The elastic
140 π-N cross sections from the phenomenological amplitudes are
120 almost identical with the experimental data (not shown)
cross section(mb)

in the energy regions shown in the figures. Peaks in ππ


100
and πN cross sections are ρ-meson and ∆ resonances, re-
80 spectively. As mentioned above, the range of validity of
60 LO-ChPT (for ππ, πN cases) is p ≪ 1 GeV for the pion
momentum. Thus, if one applies this limitation to the
40
momentum of colliding particles √ in the
√ CM frame,
√ one
20 finds for the scattering energy √ s ≪ s0 with s0 ∼ 2
0 GeV for the ππ scattering, and s0 ∼ 2.4 GeV for the πN
1000 1200 1400 1600 1800 2000 scattering. Indeed, as evident from the figure, deviation
s1/2(MeV) of LO-ChPT from the phenomenological cross sections is
10000 already sizable well below the upper limits (mainly be-
cause of the ρ-meson and ∆ resonances). For example,
N-N in the ππ case, LO-ChPT gives the same √ tendency as
1000 the phenomenological cross section up to s ∼ 600 MeV
cross section (mb)

(qcm ∼ 270 MeV), but beyond that the deviation is not


small. Therefore, it would be safe to consider the validity
100 region of LO-ChPT to be
√ q
s < EFT
∼ smax ≡ 600 MeV (ππ scattering), (25)
10
√ √
which is of course within s ≪ s0 ∼ 2 GeV. Later we
will use this limit to evaluate the maximum temperature
1
1.86 1.88 1.9 1.92 1.94 1.96 1.98 2 2.02 2.04 up to which LO-ChPT is applicable. From the figures,
s1/2 (GeV) it is obvious that the difference of two parametrizations
becomes larger and larger with increasing energies. The
FIG. 1: (Color online) Comparison of the elastic cross sec- ππ and πN cross sections from LO-ChPT monotonically
tions in ππ, πN , and N N scatterings. Square and triangle increase and become too large compared to the actual
points are the results of the low energy EFT and the phe- physical cross sections. We will see in the next section
nomenological amplitudes, respectively. The actual fit to the that this difference greatly affects the numerical value of
experimental data was done for the differential cross sections. the shear viscosity.
The N N elastic cross section is shown in logarithmic scale.

III. NUMERICAL RESULTS


valid for p ≪ 4πfπ ∼ 1 GeV with p being the magni-
tude of pion momentum. In the heavy baryon ChPT for A. Pion gas
the πN case, there is an additional expansion parameter
p/mN ≪ 1, which gives almost the same limitation as Let us first discuss the case with only pions. This is
for the ππ case. Lastly, for the N N case, the differential important for properly understanding the effects of nu-
cross section (24) is valid only for small qcm , or equiva- cleons in the next subsection and, at the same time, for
lently, small scattering energy near the threshold. This checking the validity of our calculation compared with
fact implies that the validity of results derived from the the existing results [14, 15]. As we mentioned before, our
7

formalism for the πN gas can be easily reduced to the 1.4


<s> (GeV2)
case with only pions (by setting the nucleon degeneracy <s>+Σ (GeV2)
1.2
factor gN → 0, for example). Then, the shear viscosity
coefficient is given by the first term of Eq. (15), and we 1
determine B π (p) by solving the Boltzmann equation (1)
0.8
with the collision term given by the first term of Eq. (3).
The expansion of B π (p) with respect to the orthogonal 0.6
polynomials was taken up to the third order to ensure
0.4
the convergence of the result.
0.2

0
1. Range of validity 0 20 40 60 80 100 120 140 160 180
T(MeV)
Before presenting our numerical results, we clarify the
range of validity in temperature for two parametrizations FIG. 2: (Color online) Temperature dependence of the aver-
of the scattering amplitudes. As we already specified in age scattering energy squared hsi and the measure of highest
Eq. (25), LO-ChPTpis valid only in a limited kinemat- scattering energy squared smax = hsi + Σ.

ical regime: s < EFT
∼ smax = 600 MeV, while the phe-
nomenological
√ amplitude is by construction valid up to 10000
s = 1.15 GeV. These conditions can be translated to
the limitation in temperature in the following way. Con-
1000
sider the binary collisions in thermal equilibrium. Since
each collision takes place between particles in thermal
equilibrium, the scattering energy squared s will fluctu-
λ/d

100
ate around its mean value hsi with a width Σ. The av-
erage hsi and the width (standard deviation) Σ may be
defined by 10

d3 p1 d3 p2 π π
R R
(2π)3 (2π)3 s(p1 , p2 ) f0 (p1 )f0 (p2 ) 1
hsi ≡ d3 p1
R d3 p2 π , (26) 20 40 60 80 100 120 140 160 180
π
R
(2π)3 (2π)3 f0 (p1 )f0 (p2 ) T (MeV)
p
Σ≡ hs2 i − hsi2 . (27)
FIG. 3: (Color online) The ratio λ/d as a function of temper-
ature T . The description based on the Boltzmann equation
Because the scattering energy squared of most of the
is valid when λ/d ≫ 1.
collisions are below smax (T ) ≡ hsi + Σ, one can regard
smax (T ) as (a measure of) the highest energy squared at
temperature T . So, one may interpret the validity con-
where λ is the mean-free path
dition of LO-ChPT as smax (T ) < sEFT max = 0.36 GeV ,
2

and thus obtain the validity condition for temperature 1


T < TmaxEFT EFT
with Tmax EFT
given by smax (Tmax ) = 0.36 GeV2 . λ= (29)

In Fig. 2, we show hsi and smax as functions of temper-
EFT with σ being the cross section, and d is the interac-
ature. From the figure, one can read Tmax ∼ 70 MeV.
Therefore, we may conclude that the results of LO-ChPT tion range, meaning that each collision takes place in-
are reliable only up to T ∼ 70 MeV. On the other hand, dependently. If we define the interaction range d by the
the phenomenological amplitude is valid up to s ∼ 1.3 Compton length of pions: d ∼ 1/mπ , then the valid-
GeV2 . Therefore, we expect from the figure that the re- ity condition for the Boltzmann equation reads λ/d ∼
sults of the phenomenological amplitude will be reliable mπ /nσ ≫ 1. Alternatively, if we define d through the
up to temperature close to Tc ∼ 170 MeV. cross section
p as σ ∼ πd2 , we obtain another expression
3
λ/d ∼ π/σ /n ≫ 1 . Both expressions give similar re-
There is another important check for the validity of our
framework. It is the applicability of the Boltzmann equa- striction on temperature. Figure 3 shows the ratio λ/d
tions. Recall that in deriving the Boltzmann equations, as a function of temperature, where the (phenomenolog-
one assumes that a two-point correlation function can be ical) cross section σ is estimated by its thermal average.
factorized into a product of two one-point functions (one One finds λ/d > ∼ 3 at T = 140 MeV, which manages to
particle distributions f (x, p, t)). This is physically ac- satisfy the inequalities. At this temperature, the mean-
ceptable when the density of particles n is small enough. free path of pions is estimated as λ ∼ 4 fm, which is
This condition is normally expressed as consistent with the literature [23]. As the temperature is
decreased, the mean-free path becomes longer, and the
λ≫d (28) Boltzmann description gets better and better.
8

Combining these two results, we may conclude that 0.004


the phenomenological analysis of a pion gas based on Boltzmann [ pheno
0.0035 EFT
the Boltzmann equation will be valid up to temperature pheno
-
η=p/3σ [ EFT
T ∼ 140 MeV which is much higher than the LO-ChPT

shear viscosity(GeV3)
0.003
limit T ∼ 70 MeV, but lower than the critical temper-
0.0025
ature Tc ∼ 170 MeV. However, we notice that there is
a caveat to this conclusion. In fact, even though we 0.002
have satisfactory description of ππ scatterings in wider
0.0015
kinematical regime, we have to worry about at least two
other effects as temperature increases. The first one is 0.001
the effects of other (heavier) degrees of freedom such as 0.0005
kaons, and the second is the possible modification of pi-
ons in thermal enviornment. Both are however beyond 0
0 20 40 60 80 100 120 140 160
the scope of the paper and we leave them for future prob-
T(MeV)
lems.
FIG. 4: (Color online) Shear viscosity coefficient η of a pion
gas as a function of temperature. Numerical results from the
2. Shear viscosity coefficient η Boltzmann equations are compared with the rough estimate
(30).
Figure 4 shows the shear viscosity coefficient η as a
function of temperature T . Open diamonds (⋄) and open
triangles (△) are based on numerical solutions to the
Boltzmann equations with the phenomenological ampli-
tudes and LO-ChPT, respectively. Remarkably, the two nomenological amplitude does not grow, we expect that
results show quite different behavior. While the result of the shear viscosity decreases in the ChPT case while the
LO-ChPT decreases with increasing T , that of the phe- opposite happens in the phenomenological case.
nomenological amplitude shows the opposite behavior.
Now let us come back to Fig. 4, where we also show the
Qualitative behavior of two different results can be
rough estimate (30) with two different parametrizations.
easily understood by using rough estimate of the vis-
For the (average) cross section σ in the rough estimate
cosity coefficient. In classical transport theory for a di-
(30), we evaluate it at the average momentum p, i.e.,
lute gas of one component, the shear viscosity coefficient
hσi ≡ σ(p(T )).[32] Comparing the results of the Boltz-
is expressed in terms of the mean-free path λ = 1/nσ,
mann equations and of the rough estimate (30), we find
Eq. (29), as follows:
that qualitative agreement of the results in two different
1 ways of computation. Nontrivial behavior of the rough
η∼ npλ , (30) estimate for the phenomenological amplitude (filled dia-
3
mond points) is due to the resonance shape of the elas-
where n is the particle number density, p is the aver- tic cross section in Fig. 1. Indeed, the valley around
age momentum, and σ is the binary cross section. Thus, T ∼ 110 MeV corresponds to the ρ meson peak around
the shear viscosity coefficient is inversely proportional to s ∼ 0.5 GeV2 in the ππ cross section (According to Fig. 2,
the cross section. As we already saw in Fig. 2, the typ- s ∼ 0.5 GeV2 is translated into √T ∼ 150 MeV, which is
ical energy in the ππ scattering increases with tempera- further diminished due to p ∼ T in the numerator of
ture. This means that average cross section hσi indirectly η). In the Boltzmann equations, such structure is further
depends on temperature. NoticeR also that the average washed out by thermal average. From these analyses, we
momentum p =√ d3 p |p| f0π (p)/ d3 pf0π (p) roughly in-
R
now understand that the behavior of the numerical re-
creases like p ∝ T (because (p)2 /2m ∼ 3kT /2). There- sults of the Boltzmann equations is largely due to the
fore, temperature dependence of the shear viscosity coef- energy dependence of the cross sections used in the colli-
ficient is essentially determined by the interplay between sion terms. In particular, the decreasing η of LO-ChPT
those of p and hσi. For example, if the average cross at relatively high temperature T > ∼ 80 MeV is an ar-
section
√ hσi increases rapidly as a function of T (faster tifact of too large cross section outside of the validity
than T ), η ∼ p̄/hσi is a decreasing function of T . But region of LO-ChPT. On the other hand, our most reli-
if hσi is almost constant, η is an increasing function of able result (open diamonds ⋄ in Fig. 4) shows linear in-
T . Based on these considerations and Fig. 1, we can crease with temperature for T > 120 MeV. This behavior
easily deduce the followings: As for the temperature de- is consistent with that of hadronic resonance gas mod-
pendence of the shear viscosity, we expect that ηpheno is els [9, 24]. Moreover, our result is also consistent with
slightly smaller than ηChPT at low temperature (because that of Ultrarelativistic Quantum Molecular Dynamics
σpheno >∼ σChPT ), while ηpheno > ηChPT at high temper- (UrQMD) for a meson gas [17], which also gives linear de-
ature. More precisely, since hσi of the ChPT monoton- pendence on T , and η ∼ 0.1 GeV·fm−2 = 0.0039 GeV−3
ically increases as hσi ∼ T 2 /fπ4 while that of the phe- at T = 150 − 160 MeV.
9

6 0.025
EFT extrapolation
phenomenological
5 KSS bound
0.02

|d(η/s)/dT| (MeV-1)
4
0.015
η/s

3
0.01
2

1 0.005

0 0
80 100 120 140 160 180 100 110 120 130 140 150 160 170 180
T(MeV) T(MeV)

FIG. 5: (Color online) The ratio η/s as a function of tem- FIG. 6: (Color online) The absolute value of the slope
perature T in a pion gas system. Solid (△) and dashed (⋄) |d(η/s)/dT | for the result of phenomenological amplitude.
curves are, respectively, the results with LO-ChPT and the Straight line is the linear extrapolation at T = 140 MeV.
phenomenological amplitude. Dotted line corresponds to the
conjectured lower bound η/s = 1/4π.
the other property to catch the indication of phase tran-
sition. As we mentioned in Introduction, Ref. [6] suggests
3. The ratio η/s that the ratio has a minimum around the (pseudo-) crit-
ical temperature Tc . If this is indeed the case, it implies
Figure 5 shows our numerical results of the ratio η/s as that (for crossover transition) the curve of η/s as a func-
a function of temperature T . Dashed and solid lines cor- tion of temperature will have a convex form around Tc ,
respond to the results of LO-ChPT and the phenomeno- and the slope will decrease as temperature approaches Tc
logical amplitude, respectively. In both cases, η/s is a from the left. (This kind of argument will hold as far as
monotonically decreasing function of T . Notice that the the temperature is not too far away from the critical end
result of LO-ChPT violates the conjectured bound 1/4π point.) What we observed in Fig. 5 for the phenomeno-
(the KSS bound, shown as the dotted line) at around logical amplitude is indeed the decrease of the slope with
the critical temperature Tc ∼ 170 MeV. On the other increasing temperature. Thus, even if the ratio is still
hand, η/s from the phenomenological amplitude keeps well above the conjectured bound, we can anticipate the
well above the KSS bound up to temperature ∼ Tc . existence of phase transition.
These are consistent with the results of Refs. [14, 15]. In Fig. 6, we show the (absolute value of) slope of
Since the entropy density is common in both cases s ∝ the ratio η/s for the phenomenological amplitude. Re-
T 3 , behavior of η/s can be understood by that of η itself. call that we have estimated the border of validity region
For example, we saw that η decreases in LO-ChPT, while of our calculation to be T ∼ 140 MeV. Thus, we use
it increases in the phenomenological case. Such difference the results around T ∼ 140 MeV for linear extrapolation
affects on the ratio η/s: it decreases faster in the LO- towards higher temperature. More precisely, we approxi-
ChPT case than in the phenomenological one. Also there mate the curve R(T ) ≡ η/s around at some temperature
is a crossing point for the two results of η/s at T ∼ T = T0 < Tc as
80 MeV (see Fig. 5), and this point coincides with that R′′ (T0 )
of the shear viscosity η (see the open diamonds and open R(T ) ≃ R(T0 )+R′ (T0 )(T −T0 )+ (T −T0 )2 . (31)
2
triangles in Fig. 4).
It has been argued in Ref. [14] that violation of the KSS Then the slope dR(T )/dT is approximated by a lin-
bound η/s ≥ 1/4π suggests the existence of phase transi- ear function of T . The critical temperature Tc may
tion (or crossover transition) in order for the KSS bound be defined by the temperature where the slope is zero:
to remain valid. This kind of argument is of course dan- dR(T )/dT = 0. Namely,
gerous because the precise value of η/s depends on the R′ (T0 )
amplitude in the collision term, and the result of Ref. [14] Tc ≃ T0 − , (32)
R′′ (T0 )
is based on LO-ChPT, whose applicability is limited to
T < ∼ 70 MeV as we already discussed in detail. In fact, where R′ (T0 ) < 0 and R′′ (T0 ) > 0 on the left-hand side
more reliable result with the phenomenological amplitude of a convex function R(T ). The result of linear extrapo-
does not violate the KSS bound even around the critical lation at T = T0 = 140 MeV is shown on the same figure.
temperature. Therefore, “violation of the KSS bound” Remarkably, the temperature at which the straight line
cannot be the signature of phase transition. Instead of cuts the horizontal axis is 173 MeV, which is quite a rea-
seeing violation of the KSS bound, we propose to check sonable result as the critical temperature. In addition to
10

this, one can guess the value of η/s at T = Tc by using 180 7


the approximation (31): 160 6.5
6
140
(R′ (T0 ))2 5.5
R(Tc ) ≃ R(T0 ) − . (33) 120
5
2R′′ (T0 ) 100 4.5
T (MeV)
80 4
If one substitutes numerical values at T0 = 140 MeV, 3.5
60
namely, R(T0 ) ≃ 1.0, R′ (T0 ) ≃ −0.005 MeV−1 and
40
R′′ (T0 ) ≃ 1.38×10−4 MeV−2 which are read from Figs. 5
20
and 6, one finds R = η/s ≃ 0.9 at T = Tc . These re-
0
sults should be understood with reservation at least for 0 200 400 600 800 1000
two reasons. First of all, there is ambiguity in the choice
of T0 , which will affect the values of Tc and η/s. How- µ (MeV)
ever, in fact, T0 cannot be taken arbitrary because we
need to take T0 as close to Tc as possible for the linear FIG. 7: A measure of the highest scattering energy squared
extrapolation to be accurate. To obtain reasonable val- sNN
max = hs
NN
i + ΣNN for the N N scattering is plotted in
ues of Tc and η/s, the largest possible value for T0 is the T -µ plane. Gradation represents the values in the range
preferable. Since the choice T0 =140 MeV is the upper 3.5 GeV2 < sNN 2
max < 7 GeV . Solid curve corresponds to the

limit of our validity region, we expect that the estimated borderline sNN
max = 5.29 GeV 2
.
value Tc = 173 MeV is the best value of our calcula-
tion. Therefore, even if there might be some ambiguity
in selecting T0 , we can say that we have chosen the best a pion gas presented in the previous subsection change
value. The second source which might change the values under the influence of the chemical potential.
of Tc and η/s is the possible contributions from heavier
mesons. Since such contributions become more impor-
1. Range of validity
tant as T → Tc , our results with only pions become bet-
ter as we depart from Tc (which is however not desirable
for determination of Tc ). Still, we expect our results are We start again by the discussion on the range of valid-
not so bad because at T ∼ 100 MeV such heavier parti- ity of our framework. Since we have to take the πN and
cles can be ignored, and even at T = 140 MeV, number N N collisions into account in the πN gas, we introduce
of kaons amounts to only 20% of total particles. the average scattering energy squared between particles
For the crossover transition, the ratio η/s will be con- i and j (i, j = π or N ):
tinuous at the (pseudo) critical temperature Tc . This dp31 dp32 j
ij i
R R
immediately implies that the numerical value of η/s de- ij (2π)3 (2π)3 s (p1 , p2 ) f0 (p1 )f0 (p2 )
termined above is relevant even in the deconfined phase. hs i ≡ dp31 R dp32 j
. (34)
i
R
Our result η/s ∼ 0.9 is well above the KSS bound, but (2π)3 (2π)3 f0 (p1 )f0 (p2 )

is small enough compared to the weak-coupling QCD re-


For the ππ scattering, this is of course equivalent to
sult (see for example, Fig. 4 of Ref. [6]). In this sense,
Eq. (26) and depends only on T . But for the πN and
the QCD matter around Tc could be “strongly interact-
N N scatterings, the average values depend on both T
ing”. However, we should be careful when we draw such
and µ. Similarly as before, p we further define the stan-
a conclusion from the value of η/s. In fact, the small-
dard deviation by Σij ≡ h(sij )2 i − hsij i2 , and regard
ness of the ratio η/s is not a direct consequence of large
sij ij
max (T, µ) ≡ hs i + Σ
ij
as a measure of the highest en-
cross section which may be realized by a strongly inter-
ergy of the i, j scattering at temperature T and baryon
acting matter. In our calculation with the phenomeno-
chemical potential µ.
logical amplitude, the cross section does not grow a lot
In Fig. 7, we show the values of sN N
max on the T -µ
(in contrast with the LO-ChPT) and the viscosity η even
plane. Recall that the fit to experimental data in the
increases as T → Tc . Still, since the entropy increases
N N scatterings is by construction valid up to sN N <
faster than η, the ratio η/s becomes a decreasing func-
(2.04)2 = 4.16 GeV2 . However, in fact, our parametriza-
tion of temperature. Therefore, the smallness of the ratio
tion works well up to slightly higher value sN N ∼ (2.3)2 =
is realized in a nontrivial way.
5.29 GeV2 . Therefore, we define the borderline of the va-
lidity region by temperature and chemical potential that
satisfy sN N 2
max (T, µ) = 5.29 GeV . The result is shown on
B. Pion-nucleon gas the same figure 7 by a thick solid curve. The maximum
temperature is about 130 MeV or slightly smaller than
Let us now present the numerical results for a dilute that, while the chemical potential is not restricted up to
πN gas. As advocated in Introduction, addition of nu- µ = 1 GeV. If we perform the same analysis for the πN
cleons to a pion gas enables us to study the effects of scattering, the borderline defined by sπN 2
max = 4 GeV lo-
baryon chemical potential. We discuss how the results of cates outside of the region of our interest T < 180 MeV,
11

160 0.005
λπ/d= 3 EFT/300MeV
140 λπ/d =10 0.0045 EFT/500MEV
λN/d= 1 EFT/700MeV
λN/d= 3 0.004

shear viscosity (GeV3)


120 pheno/300MeV
λN/d=10 pheno/500MeV
0.0035 pheno/700MeV
100
T (MeV)

0.003
80
0.0025
60
0.002
40 0.0015
20 0.001
0 0.0005
0 200 400 600 800 1000 60 80 100 120 140 160 180
µ (MeV) T (MeV)

FIG. 8: (Color online) The lines for the ratio λπ,N /d in a πN FIG. 9: (Color online) Temperature dependence of the shear
gas. In the region of low density and high temperature, we viscosity coefficient η of a πN gas at different values of chemi-
show only λπ /d, while in the region of low temperature and cal potential µ = 300, 500, 700 MeV. Comparison is made be-
high density, λN /d. We may regard the line for λ/d = 3 as tween the results of the low energy EFT and the phenomeno-
the border of the validity region of the Boltzmann equations. logical amplitudes.
The line for λπ /d = 1 is far outside of the T -µ region shown
here.

µ < 1 GeV. On the other hand, the validity region of (twice of the saturating value at high energy, see Fig. 1)
the low energy EFT is severely restricted.
√ For example, and uses d ∼ 1/mπ again for the interaction range, then
if we take the maximum value of sN N to be 1.88 GeV the condition for nucleons λN ≫ d can be estimated as
(beyond which the phenomenological fit starts to devi- nN ≪ mπ /σN N ≃ n0N with n0N = 0.157 fm−3 being the
ate from the low energy EFT fit, see Appendix A), then normal nuclear density. This condition is physically quite
the borderline defined by sN N 2
max = (1.88) = 3.53 GeV
2
reasonable since we do not expect standard Boltzmann
is very close to the horizontal axis (see Fig. 7). Even if description useful at normal nuclear matter density. To
we relax the condition to higher value sN N 2
max = (1.90) = obtain more precise restriction depending on T and µ,
2
3.61 GeV , allowed region is still very narrow. we need to use Eqs. (36) and (37) and estimate each
Let us also examine the validity condition for the Boltz- condition by replacing σ and n by their thermal aver-
mann equations. The criterion is again given by Eq. (28), ages. In Fig. 8, we show the lines for several values of
but now pion’s mean-free path should be modified in the the ratio λ/d in the T -µ plane. For simplicity, we used
presence of nucleons, and we have to separately consider d = 1/mπ for both pions and nucleons. In the region of
nucleon’s mean-free path, too. According to classical low baryon density where pions are dominant degrees of
transport theories, the mean-free path of i-th component freedom for transport phenomena, we show the ratio for
in a mixed gas is modified as [25] pions λπ /d. The line for λπ /d = 3 (solid line) is consis-
tent with the previous result shown in Fig. 3. We may
λ0 regard this line as the border of the validity of the Boltz-
λi = q i , (35)
P 1+mi /mj σij nj mann equations at low densities. On the other hand, at
1+ j6=i 2 σii ni large chemical potential and low temperature, transport
phenomena is dominated by nucleons. The lines for the
where λ0i = 1/ni σi is the mean-free path of a pure gas of ratio λN /d = 1, 3, 10 are shown in the figure. We may
i-th component, and σij is the cross section between i-th again regard the line λN /d = 3 as the borderline, which
and j-th components. Therefore, the mean-free paths of reaches at (T, µ) = (0, 950MeV). This condition is more
pions and nucleons are respectively given by restrictive than that of Fig. 7.
1
λπ = q , (36) Combining these two analyses, we may conclude that
1+mπ /mN
nπ σππ + nN σπN 2 our framework is valid in a wide region of the T -µ plane,
1 whose boundary is roughly given by (a quarter of) the
λN = q . (37) elliptic curve connecting (T, µ) ∼ (130MeV, 0) and (0,
1+mN /mπ
nN σN N + nπ σπN 2 950MeV). The first point (130MeV, 0) is specified by the
measure of the highest scattering energy squared, and the
Consider the condition for nucleons. If one takes last point (0, 950MeV) is from the validity limit of the
σN N ∼ 40 mb as a typical value of the N N cross section Boltzmann equations.
12

T = 50 MeV T = 100 MeV


0.004 0.004
η η
0.0035 π 0.0035
η η
π
N
η η
N
0.003 0.003

0.0025 0.0025

0.002 0.002

0.0015 0.0015

0.001 0.001

0.0005 0.0005

0 0
100 200 300 400 500 600 700 800 900 1000 100 200 300 400 500 600 700 800 900 1000
µ (MeV) µ (MeV)

FIG. 10: (Color online) µ dependence of η at T = 50MeV (left) and T = 100MeV (right), and its decomposition η = η π + η N .

2. Shear viscosity coefficient η formula interpolates a pure pion gas at low µ and a pure
nucleon gas at high µ. Notice that the pion contribution
In Fig. 9, we show temperature dependence of η at η π is always smaller than η0π due to the presence of πN
different values of chemical potential µ = 300, 500, 700 interaction, as we alluded before in relation to Eq. (21).
MeV. Comparison is made between the results of the This is exactly what we expected. On the other hand,
low energy EFT and the phenomenological amplitudes. it is not straightforward to predict the behavior of the
These temperature dependence is qualitatively consistent total shear viscosity. If one plots the rough estimate (38)
with the previous results of the pion gas (see Fig. 4). As as a function of nN /nπ assuming that the cross sections
for the µ dependence, however, the shear viscosity coef- are constant and are of the same order, one finds that η
ficient shows nontrivial behavior. In particular, in the decreases at small values of nN /nπ (small µ), but turns
window 80 MeV < T <130 MeV, it increases with in- into increase at large nN /nπ (large µ). If one changes the
creasing µ. This is not quite understandable at first be- numerical value of cross sections a little, then the curve
cause we expect that the cross section of ππ scattering turns into monotonic increase. In fact, both can happen
will effectively enhance in the presence of nucleons while in reality depending on temperature. In Fig. 10, we have
the effects of nucleon viscosity may be ignored at lower plotted η at T = 50 MeV and 100 MeV as a function of
density, meaning that the shear viscosity will decrease. µ. While the total viscosity behaves differently in these
To understand what really happens when µ 6= 0, let us two panels, one can see monotonic decrease (increase)
again consider a rough estimate of the shear viscosity of of η π (η N ), namely, the interplay between η π and η N
a mixed gas. Let the shear viscosity coefficient of a pure at both temperatures. Therefore, the increase of η with
gas of particle species i be η0i = ni pi λ0i /3 where ni is the increasing µ observed in Fig. 9 can be understood as a
number density, pi is the average momentum, and λ0i is result of enhancement of η N .
the mean-free path. Then the shear viscosity coefficient
for n component gas is given by the sum of each viscosity
η0i with modified mean-free path λi given in Eq. (35) [25]: 3. The ratio η/s

X λi Figure 11 shows temperature dependence of the ratio


ηmix = η0i .
i
λ0i η/s of the πN gas at different values of baryon chemi-
cal potential µ = 300, 500, 700 MeV. The left-hand side
Thus for the πN gas mixture, we obtain is the results of the low energy EFT, while the right-
hand side, the phenomenological amplitudes. The ratio
η = ηπ + ηN (38) η/s shows qualitatively the same behavior as in the pion
ηπ η0N gas system (Fig. 5): (i) Both the results (EFT and phe-
≃  0  + q   , nomenological amplitudes) are monotonically decreasing
√1 σπN nN mN σπN nπ
1+ 2 σππ nπ 1+ 2mπ σN N nN functions of T , (ii) (η/s)EFT > (η/s)pheno at lower T
while opposite at higher T , and (iii) the ratio of the
where we have used the approximation mπ /mN ≪ 1. In EFT violates the KSS bound at around T ∼ Tc while
the two extreme limits nN /nπ → 0 and ∞, the formula that of the phenomenological amplitudes does not. No-
(38) reduces to η0π and η0N , respectively. Therefore, this tice that the inclusion of chemical potential works to re-
13

EFT Phenomenological
4 4
µ=300MeV µ=300MeV
3.5 µ=500MeV 3.5 µ=500MeV
µ=700MeV µ=700MeV
3 KSS bound 3 KSS bound
2.5 2.5
η/s

η/s
2 2
1.5 1.5
1 1
0.5 0.5
0 0
80 100 120 140 160 180 80 100 120 140 160 180
T(MeV) T(MeV)

FIG. 11: (Color online) The ratio η/s of a πN system plotted as a function of temperature T at different values of baryon
chemical potential µ = 300, 500, 700 MeV. (Left) results of low energy EFT, (Right) results of phenomenological amplitudes.

duce the value of η/s. As a result, for the ratio of the figure shows a new structure: There is a valley at large
EFT, the temperature at which the curve cuts the KSS µ ∼ 950 MeV. We will discuss later the physical implica-
bound becomes smaller. On the other hand, for the re- tion of this structure.
sult of the phenomenological amplitudes, the flattening Let us comment again the point made in the last para-
of the curves seems to occur at lower temperature with graph of the previous subsection. The ratio η/s becomes
increasing µ. If this is indeed the symptom of crossover less than 0.3 at T > ∼ 140 MeV and µ = 700 MeV.
transition as we discussed before, one can say that the This is small enough and is close to the KSS bound
(pseudo) critical temperature Tc will decrease with in- η/s ∼ 0.1 compared to other systems such as water.
creasing µ, which is consistent with what we know from However, as shown in Figs. 9 and 10, the shear viscos-
lattice simulations or effective models. [33] ity itself grows as the system approaches phase boundary
Our result (η/s)pheno ∼ 0.5 − 0.4 at µ = 700MeV (T → Tc , µ → µc ). Therefore, even if the ratio is small
and T ∼ 100 − 140 MeV is consistent with that of enough, the πN system cannot be treated, strictly speak-
“URASiMA” (Ultra-Relativistic AA collision Simulator ing, as a perfect fluid with η = 0. Sometimes ideal hydro-
based on Multiple scattering Algorithm)[18] which is a dynamics is used to describe the matter after hadroniza-
Monte-Carlo event generator of hadronic collisions. It in- tion in heavy-ion collisions, but one will have to take into
cludes both elastic and inelastic scatterings whose cross account the effects of viscosity for a realistic simulation.
sections are given by experimental data. The numerical
coincidence of the ratio from different frameworks is very
interesting. In fact, the essential difference of URASiMA 4. Valley structure at large µ and low T
from our framework is the presence of inelastic collisions.
But as far as we consider small deviation from thermal In the right panel of Fig. 12 (low temperature
equilibrium, inelastic collisions which will change particle T = 10 MeV), we pointed out a valley structure at
numbers (such as ππ → πππ or πN → ππN ) would not high baryon chemical potential, which was not seen in
be so important, and it seems reasonable to obtain the higher temperature (left panel). Let us look at this
same result from two different frameworks. new structure in more detail and examine its possible
As we already saw in Fig. 9, the shear viscosity coef- interpretation. We recall again Ref. [6], where it was
ficient increases in the window 80 MeV< T <130 MeV suggested that the ratio η/s would give a minimum
with increasing µ. However, in the same window, the ra- at the phase transition temperature. If this is true
tio does decrease. This clearly implies that the reduction for other control parameters, in particular, chemical
of η/s at T ∼ 80 − 130 MeV is due to the increase of potential, and if there is phase transition under the
entropy. change of chemical potential, it is natural to expect
In Fig. 12, we show the µ dependence of the ratio η/s that the ratio would exhibit a valley structure with
at temperature T = 50, 100 MeV (left) and at low tem- its minimum at the critical chemical potential µc . In
perature T = 10 MeV (right). Recall that the ratio de- other words, if one finds a valley structure in the µ
creases with increasing µ in Fig. 11 where T ≥ 70 MeV direction, one can expect some kind of phase transition
was shown. This is consistent with the left figure, and around the minimum. This is what we observed in
in agreement with our original expectation as we men- Fig. 12. Then, what kind of phase transition could be
tioned in Introduction. On the other hand, the right related to this valley structure? The valley locates at
14

Phenomenological T=10MeV
100 1e+008
T=50MeV
T=100MeV 1e+007
1e+006
10
100000

η/s
η/s

10000
1000
1
100
10 phenomenological
EFT
0.1 1
200 300 400 500 600 700 800 900 1000 200 300 400 500 600 700 800 900 1000
µ(MeV) µ(MeV)

FIG. 12: (Color online) µ dependence of η/s: (left): the results of the phenomenological amplitudes at higher temperature
T = 50, 100 MeV, (right): the results at lower temperature T = 10 MeV.

low temperature T < 20 MeV and at high chemical transition correctly, we will have to include the effects
potential µ ∼ 950 MeV (which is however not enough for of higher-order correlations to the Boltzmann equations,
the quark-hadron phase transition). This is the region or start from different models (such as the σ-ω model)
where we can see the liquid-gas phase transition. [34] which are more appropriate for nuclear matter, both
At low T < 20 MeV and around normal nuclear density, of which, however, are beyond the scope of the present
there is a critical line (possibly first order) separating a paper.
nucleon gas phase and a nuclear matter (liquid). This
line terminates at around T ∼ 15 MeV [27], and above • As mentioned in Introduction, a similar valley
that temperature, there is no distinction between a gas structure was already reported in the calculation with
and a liquid. Therefore, if the valley indeed corresponds the low energy EFT [16] and the authors of Ref. [16]
to the liquid-gas phase transition, it should disappear claim that it is related to the liquid-gas phase tran-
when the temperature go far beyond T ∼ 15 MeV. In sition. We have done the same calculation with our
Fig. 13, we show the transition of the valley structure own parametrizations of low energy EFT, and obtained
from T = 5 MeV up to T = 20 MeV. The right panel consistent results with Ref. [16] as shown in the left panel
is the results of the phenomenological amplitudes. of Fig. 13. Unlike the results of the phenomenological
Clearly, with increasing temperature, the valley becomes amplitudes (the right panel), the valley persists even
shallow, which supports the interpretation that the at higher temperature. As we repeated several times,
valley structure indeed corresponds to the liquid-gas since the range of validity of the low energy EFT is
phase transition. There are several comments about this. severely restricted (see the discussion about Fig. 7), it
is quite dangerous to draw any conclusions about the
• Although the liquid-gas phase transition is a phe- physics outside of the validity region. However, a similar
nomenon in the hadronic phase, it is not obvious valley structure is observed even in our calculation with
whether the Boltzmann equations (valid for a dilute the phenomenological amplitudes, and thus our cal-
gas) correctly describe the transition to liquid phase. culation partially supports the results of low energy EFT.
Notice that the region where liquid-gas phase transition
takes place is close to the border of the validity region of • Since the entropy in our calculation is evaluated by
the Boltzmann equations, and thus it is not surprising using free particle distributions f0π,N (p), there is no in-
that there might exist another phase (i.e., liquid phase) formation about phase transition in the denominator of
outside the region of validity. Still, it would be safe η/s. However, if there is a real phase transition, entropy
to reserve that reliable information from our analysis will of course show nontrivial change around Tc or µc .
should be only the tendency towards phase transition. For example, in the liquid-gas transition of water, both
We do not expect we can describe precise structure of the shear viscosity and the entropy have nontrivial struc-
phase transition such as the order of transition and ture around Tc , and contribute to give a convex shape of
the temperature dependence of the critical chemical the ratio η/s. We emphasize that, in our calculation, the
potential µc (T ) (Our result is that the minimum of the valley structure of η/s is not the result of entropy. If
valley moves to the right with increasing T , as opposed we treated actual entropy, the structure would emerge in
to common expectation). In order to describe the more pronounced way.
15

EFT Phenomenological
25 25
T=5MeV
T=10MeV
T=15MeV
20 20 T=20MeV

15 15
η/s

η/s
10 10

5 T=5MeV 5
T=10MeV
T=15MeV
T=20MeV
0 0
920 930 940 950 960 970 980 990 1000 920 930 940 950 960 970 980 990 1000
µ(MeV) µ(MeV)

FIG. 13: (Color online) Temperature dependence of the valley structure of η/s at relatively high chemical potential. Only the
results at low temperature T = 5, 10, 15, 20 MeV are shown. (Left): Low energy EFT, (Right): Phenomenological amplitudes.

IV. SUMMARY the Chapman-Enskog method. Two of the authors (KI


and HO) are thankful to S. Muroya for his encourage-
We have performed a detailed calculation of the shear ments and constructive comments. Lastly, one of the
viscosity coefficient η and the viscosity to entropy ratio authors (HO) thanks A. Dote for useful advises on the
η/s in a wide region of the hadronic T -µ plane. Our for- fitting method of experimental data.
malism is based on the relativistic quantum Boltzmann
equations, and we found that it is very important to use APPENDIX A: PHENOMENOLOGICAL
phenomenological amplitudes for the scattering ampli- AMPLITUDES
tudes in the collision terms in order to obtain reliable
results. On the other hand, the validity region of the low In this Appendix, we explain the phenomenological
energy effective field theories is severely restricted, and amplitudes we used in the kinetic equations. Since the
reliable results based on them are also limited to a small scattering amplitude (squared) is related to the differen-
region of T -µ plane. We found that the ratio η/s de- tial cross section in the CM frame as
creases (for T >∼ 20 MeV) under the inclusion of nucleon
degrees of freedom, but still keeps above the conjectured  
KSS bound η/s = 1/4π in the region we investigated dσ 1
= |M |2 , (A1)
(T < 180 MeV, µ < 1000 MeV). Since the shear viscosity dΩ 64π 2 s
coefficient itself increases with increasing temperature,
the behavior of the ratio is largely due to the entropy.
with s being the Mandelstam variable (scattering energy
At low temperature T < ∼ 15 MeV and high baryon chem- squared), we discuss only the differential cross sections.
ical potential µ ∼ 950 MeV, we found a valley structure
in the ratio. There is some argument about the relation-
ship between such structure and phase transition, and we
expect that the valley structure found in our calculation
would correspond to the liquid-gas phase transition.
1. ππ scattering

Acknowledgments Consider the partial-wave expansion of the isospin av-


eraged ππ elastic differential cross section. Taking up to
The authors are grateful to E. Nakano for discussion p-wave scattering (ℓ = 1) yields a very nice description
and to H. Hayakawa for explaining them the essence of of the experimental data:
16

 ππ 2 2
dσ 1 X 4 X I
= P ′ + 1)
(2 I + 1) 2
(2 ℓ + 1) eiδℓ sin δℓI Pℓ (cos θcm )
dΩ I′ (2I qcm ℓ=0,1
I=0
 
4 1 2 0 5 2 2 3 2 1 2
= 2 sin δ0 + sin δ0 + · 9 sin δ1 cos θcm , (A2)
qcm 9 9 9

where 4 in the overall factor comes from the identical fac- ential cross section:
tor, θcm and qcm are the scattering angle and the magni-  πN 2ℓmax  
tude of momentum in the CM frame, and Pℓ is the ℓ-th dσ 1 X 2 I=1/2 4 I=3/2
order Legendre polynomial (P0 (x) = 1, P1 (x) = x). The = 2 Q (qcm ) + Qℓ (qcm )
dΩ qcm 6 ℓ 6
ℓ=0
numerical factors in the last line are from isospin I and
orbital angular momentum ℓ. For example, the last term ×Pℓ (cos θ) . (A4)
corresponds to I = ℓ = 1 scatteringPand thus (3/9) · 9
is from the isospin weight (2I + 1)/ I ′ (2I ′ + 1) = 3/9 The coefficients
for I = 1 and (2ℓ + 1)2 = 9 for ℓ = 1. The phase shift 2 1/2 4 3/2
δℓI depends on ℓ and I. Since the total wavefunction of a Cℓ (qcm ) ≡ Q (qcm ) + Qℓ (qcm )
6 ℓ 6
ππ system must be symmetric under the exchange, only
three phase shifts (I, ℓ) = (0, 0), (1, 1), (2, 0) are possible. are functions of qcm and are determined from the ex-
It is known that the energy dependence of the phase shifts perimental data. More precisely, each function Cℓ is ex-
are well parametrized by the following function [28]: pressed by a superposition of 30 Gaussians: For example,
for ℓ = 0, we use
tan δℓI
15
(  2 )
r X qcm − 100MeV
4m2π 2ℓ  I 
Cℓ=0 (qcm ) = An exp −
= 1− qcm Aℓ + BℓI qcm
2
+ CℓI qcm
4
+ DℓI qcm
6
60n MeV
s n=1
 2
4mπ − sIℓ 30
 (  2 )
× , (A3)
X qcm − 800MeV
s − sIℓ + An exp − ,
n=16
60n MeV

where
√ the parameters are determined to fit the data up where positions and widths of Gaussians are found by
to s = 1.15 GeV (qcm ∼ 550 MeV) and are shown in trial and error. We use different values of positions and
table II: widths for different ℓ. This kind of technique is some-
times used in describing nuclear many body wavefunc-
TABLE II: tions. By using these functions and taking the maximum
(I, ℓ) = (0, 0) (I, ℓ) = (1, 1) (I, ℓ) = (2, 0) angular momentum√2ℓmax = 8, we can fit the experimen-
A00 2.25 × 10−1 A11 3.63 × 10−2 A20 −3.71 × 10−2 tal data [29] up to s = 2 GeV (qcm ∼ 770 MeV). Notice
B00 2.46 × 10−1 B11 1.34 × 10−4 B02 −8.55 × 10−2 that this parametrization works very well at relatively
C00 −1.67 × 10−2 C11 −6.98 × 10−5 C02 −7.54 × 10−3 high scattering energies, but in fact its quality becomes
D00 −6.40 × 10−4 D11 1.41 × 10−6 D02 −1.99 × 10−4 worse at small scattering energies. This is due to the fac-
2
s00 36.7 s11 30.7 s20 −11.9 tor 1/qcm in Eq. (A4). If one absorbed this factor into
the coefficients and performed the same Gaussian fitting,
quality of the fit would be better even at low scattering
energies. However, we decided to start from the conven-
In this table, dimensionful parameters are redefined so tional expression shown in Eq. (A4), and to find another
that the mass dimension is provided by mπ . For ex- parametrization for the low energy data. We have inter-
ample, since B00 has mass dimension −2, we define polated the parametrization proposed in Ref. [30] which
B00 ≡ b00 /m2π and b00 = 2.46 × 10−1 . This phenomeno- is compactly represented for the scattering amplitude:
logical parametrization describes the experimental data
very well. We use this for the ππ scattering amplitude.
 
MπN = b0 + b1 (~t · ~τ ) + c0 + c1 ~t · ~τ (~q · q~′ )
 
+i d0 + d1 ~t · ~τ ~σ · (~q × q~′ ) ,
2. πN scattering
where ~t and ~τ /2 are the isospin vectors of a pion and a
Let us now turn to the πN scattering. Consider the nucleon respectively, ~q and q~′ are the momenta of incom-
partial-wave expansion of the isospin averaged πN differ- ing particles in the CM frame, and lastly ~σ is the Pauli
17
P P
where 16 = I=0,1 (2I + 1) S=0,1 (2S + 1) and
TABLE III:
b0 −0.010 b1 −0.091 (dσ/dΩ)I,S in the right-hand side are the differential
c0 0.208 c1 0.175 cross sections with isospin I and spin S specified. One
d0 −0.190 d1 −0.069 can further decompose each contribution depending on
the value of orbital angular momentum ℓ. (Notice that
the total N N system must be antisymmetric under the
exchange of two particles: (−1)ℓ+S+I = −1.)
matrix. The parameters are determined as shown in ta-
ble III (dimensionful parameters are again redefined by
using mπ so that they become dimensionless).
These two √ different parametrizations are smoothly 

0,0
1
Oℓ0,0 (qcm )Pℓ (cos θ),
X
matched at s = 1101 MeV (qcm =79 MeV), giving a = 2
(A6)
very nice parametrization of the experimental data for a dΩ qcm
ℓ=odd
wide region of scattering energies. 

1,0
1
Eℓ1,0 (qcm )Pℓ (cos θ), (A7)
X
= 2
dΩ qcm
ℓ=even
3. N N scattering 

0,1
1
Eℓ0,1 (qcm )Pℓ (cos θ), (A8)
X
= 2
dΩ qcm
Let us finally discuss the N N scattering. In the text ℓ=even
1,1
we discussed only the s-wave scattering, but here we treat

dσ 1
Oℓ1,1 (qcm )Pℓ (cos θ).
X
arbitrary orbital angular momentum. We define the spin- = 2
(A9)
dΩ qcm
ℓ=odd
isospin averaged N N differential cross section
 N N ( 
0,0  1,0
dσ 1 dσ dσ
≡ +3
dΩ averaged 16 dΩ dΩ Inserting these into Eq. (A5), we obtain the spin-isospin
 0,1

 1,1 )
dσ averaged differential cross section separately for even and
+3 +9 ,(A5) odd ℓ:
dΩ dΩ

 N N " #
dσ 1 X  3 1,0 3 0,1
 X  1 0,0 9 1,1

= E + Eℓ Pℓ (cos θ) + O + Oℓ Pℓ (cos θ) . (A10)
dΩ averaged
2
qcm 16 ℓ 16 16 ℓ 16
ℓ=even ℓ=odd

When ℓ = 0, this gives the spin-isospin averaged scatter- partial-wave expansion (A11). For actual fitting of the
ing amplitude used in the parametrization of low energy experimental data [31], we divide the rest of the region
effective theory. into two:
Restriction to the s-wave alone may be a good ap- (i) 6.762 MeV√ < qcm < 48.76 MeV
proximation at low scattering energy, but to obtain a (1876 MeV < s < 1879 MeV)
parametrization which describes the data in much wider (ii) 48.76 MeV
√ < qcm < 405 MeV
range of energies, we have to include larger ℓ. To do this, (1879 MeV < s < 2043 MeV)
we again perform the partial-wave expansion of the spin- In each region, we determine the coefficient Dℓ (qcm ) sim-
isospin averaged differential cross section (A5) and fit the ilarly as in the case for πN scattering, so that the curve is
energy-dependent coefficients by using the Gaussian su- smoothly connected at the matching points qcm = 6.762
perposition. We define the coefficients Dℓ (qcm ) similarly MeV and qcm = 48.76 MeV. In region (i), the highest
to Eq. (A4): value of the angular momentum is taken to be 6, while
 N N in region (ii), much higher√value 16. We performed the
dσ 1 X fit up to qcm = 405 MeV ( s = 2043 MeV).
= 2
Dℓ (qcm )Pℓ (cos θ). (A11)
dΩ averaged qcm

Since the low energy data are already described well by APPENDIX B: SOLVING THE LINEARIZED
the parametrization (24) and the parameters given in ta- BOLTZMANN EQUATIONS
ble I, we adopt them as the low energy part of the global
parametrization.
√ We use them up to qcm = 6.76 MeV In this Appendix, we discuss how to solve the Boltz-
( s = 1.88 GeV), and beyond that, we switch to the mann equations (1) and (2). First of all, recall that the lo-
18

cal equilibrium state is defined by the distributions f0π,N hand side of the Boltzmann equations (which have coor-
which make the collision terms (the right hand side of dinate derivative ∂µ ) do not vanish. With this in mind,
the Boltzmann equations) vanishing [see Eqs. (6), (7)]. the Boltzmann equations linearized with respect to the
However it should be noticed that the local equilibrium deviations δf π,N from the local equilibrium f0π,N (x, p)
is not the solution to the Boltzmann equations: Indeed, can read as follows (leading order of the left hand side
the parameters T, µ and V µ characterizing f0π,N are de- does not contain δf π,N ):
pendent on the coordinates, which implies that the left


∂µ f0π = C ππ [δf π , f0π ] + C ππ [f0π , δf π ] + C πN [δf π , f0N ] + C πN [f0π , δf N ] , (B1)
Epπ

∂µ f0N = C N π [δf N , f0π ] + C N π [f0N , δf π ] + C N N [δf N , f0N ] + C N N [f0N , δf N ] , (B2)
EpN

where the notation for the collision term is for example (see Eq. (3))

C ππ [δf π , f0π ] ≡ C ππ [f0π + δf π , f0π ] − C ππ [f0π , f0π ]



Z n
= dΓππ δf1π f02 π π
(1 + f03 )(1 + f0pπ π π
) + f01 f02 δf3π (1 + f0pπ
)
2
o
−δf1π (1 + f02 π π π
)f03 f0p − (1 + f01π
)(1 + f02π
)δf3π f0p
π
. (B3)

This corresponds to the lowest order Chapman-Enskog method. So far, the deviations δf π,N are in principle arbitrary,
but for the purpose of computing the shear viscosity coefficient η, we can restrict only to the deviations that are directly
from the shear ∂i Vj 6= 0 (i 6= j). Thus we ignore any other effects except the shear.
By using the particle number conservation ∂t n = −n∇i V i and the energy-momentum conservation ∂ν T µν = 0, the
left hand sides of Eqs. (B1) and (B2) can be rewritten as
pµ f0π (1 + f0π )
 
π δij 2
∇i V j trl ,

π
∂µ f 0 (x) = β π
p i p j − p (B4)
Ep Ep 3
pµ f0N (1 − f0N )
 
N δij 2
∇i V j trl ,

N
∂µ f0 (x) = β N
pi pj − p (B5)
Ep Ep 3

where ∇i V j trl is defined in Eq. (11). Notice that this functional form was the motivation for defining new quantities


B π,N (p) in Eqs. (13), (14). Indeed, with respect to B π,N (p), the right hand sides of Eqs. (B1) and (B2) can be expressed
rather compactly. Introducing further the following notation,
 
π,N δij
Bij (p) ≡ p̂i p̂j − B π,N (p) , (B6)
3
π
the linearized Boltzmann equations (B1) and (B2) can be expressed as (f01 ≡ f0π (k1 ), f0p
π
≡ f0π (p), etc.)
π π 
f0p (1 + f0p )

δij 2 gπ
Z  
π
p p
i j − p = dΓππ (1 + f01
π π
)(1 + f02 π π
)f03 f0p Bijπ
(p) + Bijπ π
(k3 ) − Bij π
(k2 ) − Bij (k1 )
Ep 3 2
Z  
+ gN dΓπN (1 − f01 N π
)(1 + f02 N π
)f03 f0p Bij π
(p) + BijN π
(k3 ) − Bij N
(k2 ) − Bij (k1 ) ,(B7)

N N
f0p (1 − f0p )
 
δij 2
Z  
N
pi pj − p = gπ dΓN π (1 + f01π N
)(1 − f02 π N
)f03 N
f0p Bij π
(p) + Bij N
(k3 ) − Bij π
(k2 ) − Bij (k1 )
Ep 3
gN
Z  
+ dΓN N(1 − f01
N N
)(1 − f02 N N
)f03 N
f0p Bij N
(p) + Bij N
(k3 ) − Bij N
(k2 ) − Bij (k1 ) .(B8)
2

These are the equations for B π,N (p). As explained in to a finite dimensional functional space. More precisely,
the text, we solve these equations by restricting B π,N (p)
19

1 δij
we expand B π,N (p) in terms of orthogonal polynomial N
)W(m)
(2π)3 (pˆi pˆj − 3 (p) (m = 0, 1, 2), and integrate them
π,N
functions W(n) (p) [see Eq. (16)] and approximate the over p. Then, by using the orthogonal conditions (17),
series by the first three terms [see Eqs. (19) and (20)]. (18), we will obtain six independent equations for the
The expansion coefficients (bπ,N (n) , n = 0, 1, 2) in Eqs. (19)
coefficients bπ,N
(n) (n = 0, 1, 2). The resulting equations
and (20) are determined as follows: we multiply Eq. (B7) are still complicated but can be solved in a numerical
1
by (2π)
δ π
ˆi pˆj − 3ij )W(m) way.
3 (p (p) (m = 0, 1, 2), Eq. (B8) by

[1] For an overview of the RHIC results, see the special [23] P. Gerber, H. Leutwyler, and J.L. Goity, Phys. Lett.
volume ”First three years of operation of RHIC”, Nucl. B246 (1990) 513.
Phys. A757, issues 1-2 (2005). [24] P. Danielewicz and M. Gyulassy, Phys. Rev. D 31, 53
[2] P. Kovtun, D. T. Son and A. O. Starinets, Phys. Rev. (1985).
Lett. 94, 111601 (2005) [arXiv:hep-th/0405231]. [25] E. H. Kennard, “Kinetic Theory of Gases, with an In-
[3] S. Gavin and M. Abdel-Aziz, Phys. Rev. Lett. 97, 162302 troduction to Statistical Mechanics” (McGraw-Hill, New
(2006) [arXiv:nucl-th/0606061]. York, 1938).
[4] A. Nakamura and S. Sakai, Phys. Rev. Lett. 94, 072305 [26] R. A. Lacey et al., Phys. Rev. Lett. 98 (2007) 092301
(2005) [arXiv:hep-lat/0406009]; Nucl. Phys. A 774, [arXiv:nucl-ex/0609025].
775 (2006) [arXiv:hep-lat/0510039]; PoS LAT2005, 186 [27] J. B. Natowitz et al., Phys. Rev. C 65, 034618
(2006) [arXiv:hep-lat/0510100]. (2002) [arXiv:nucl-ex/0106016], J. B. Elliott et al.
[5] H. B. Meyer, arXiv:0704.1801 [hep-lat]. [EOS Collaboration], Phys. Rev. C 67, 024609 (2003)
[6] L. P. Csernai, J. I. Kapusta and L. D. McLerran, Phys. [arXiv:nucl-ex/0205004], L. G. Moretto, J. B. El-
Rev. Lett. 97, 152303 (2006) [arXiv:nucl-th/0604032]. liott and L. Phair, Phys. Rev. C 72, 064605 (2005)
[7] T. Hirano and M. Gyulassy, Nucl. Phys. A 769, 71 (2006) [arXiv:nucl-ex/0507015].
[arXiv:nucl-th/0506049]. [28] G. Colangelo, J. Gasser and H. Leutwyler, Nucl. Phys. B
[8] S. R. De Groot, W. A. Van Leeuwen and C. G. Van 603, 125 (2001) [arXiv:hep-ph/0103088].
Weert, “Relativistic Kinetic Theory. Principles and Ap- [29] P. Bareyre, C. Bricman and G. Villet, Phys. Rev. 165
plications,” (Amsterdam, Netherlands: North-holland, (1968) 1730, R. Koch and E. Pietarinen, Nucl. Phys. B
1980, 417p) 336 (1980) 331.
[9] S. Gavin, Nucl. Phys. A 435, 826 (1985). [30] T. Ericson and W. Weise, “Pions and Nuclei” (Claren-
[10] M. Prakash, M. Prakash, R. Venugopalan and G. Welke, don, Oxford, 1988)
Phys. Rept. 227, 321 (1993). [31] Experimental data of nucleon-nucleon scattering are
[11] D. Davesne, Phys. Rev. C 53, 3069 (1996). available from online at http://nn-online.org/
[12] A. Dobado and S. N. Santalla, Phys. Rev. D 65, 096011 [32] One can evaluate hσi by its thermal average similarly as
(2002) [arXiv:hep-ph/0112299]. in Eq. (26), but the qualitative behavior is the same as
[13] A. Dobado and F. J. Llanes-Estrada, Phys. Rev. D 69, σ(p(T )).
116004 (2004) [arXiv:hep-ph/0309324]. [33] One can try to determine the values of Tc from the slope
[14] J. W. Chen and E. Nakano, Phys. Lett. B 647, 371 (2007) of each curve as in the pion gas case, but we do not
[arXiv:hep-ph/0604138]. perform such extrapolation for two reasons: First, it is
[15] A. Dobado and F. J. Llanes-Estrada, Eur. Phys. J. C 49, practically very difficult to obtain reliable results since
1011 (2007) [arXiv:hep-ph/0609255]. the highest temperature allowed for the Boltzmann equa-
[16] J. W. Chen, Y. H. Li, Y. F. Liu and E. Nakano, “QCD tions gradually decreases with increasing µ, as shown in
Viscosity to Entropy Density Ratio in the Hadronic Fig. 8. Second, the chiral phase transition at moderate
Phase,” arXiv:hep-ph/0703230. values of µ will be most likely of the first order, for which
[17] A. Muronga, Phys. Rev. C 69, 044901 (2004) the ratio will show a discontinuity and the slope will not
[arXiv:nucl-th/0309056]. necessarily zero at Tc . Position of the critical end point
[18] S. Muroya and N. Sasaki, Prog. Theor. Phys. 113, 457 may be detected from the behavior of the ratio η/s [26],
(2005) [arXiv:nucl-th/0408055]. but we do not expect our Boltzmann equations can de-
[19] S. Chapman and T.G. Cowling, “The Mathematical The- scribe the precise structure of the phase transition. We
ory of Non-uniform Gases, Third edition” (Cambridge, will discuss similar problems in relation to the nuclear
1970) liquid-gas phase transition in Sect.III.B.4.
[20] S. Weinberg, Physica A 96, 327 (1979). [34] Notice that all the examples (except for the QCD phase
[21] N. Fettes, U. G. Meissner and S. Steininger, Nucl. Phys. transition) discussed in Ref. [6] are about the liquid-gas
A 640, 199 (1998) [arXiv:hep-ph/9803266]. phase transitions.
[22] A. Sitenko and V. Tartakovskii, “Theory of Nucleus”
(Kluwer, Netherlands, 1997)

You might also like