DAMTP-2012-58
ICCUB-12-319
MAD-TH-12-03
Prepared for submission to JHEP
arXiv:1208.2672v2 [hep-th] 25 Jan 2013
Quarkonium dissociation by anisotropy
Mariano Chernicoff,a Daniel Fernández,b David Mateos,b,c and Diego Trancanellid,e
a
Department of Applied Mathematics and Theoretical Physics, Centre for Mathematical Sciences,
Wilberforce Road, Cambridge, CB3 0WA, UK
b
Departament de Fı́sica Fonamental & Institut de Ciències del Cosmos (ICC), Universitat de
Barcelona (UB), Martı́ i Franquès 1, E-08028 Barcelona, Spain
c
Institució Catalana de Recerca i Estudis Avançats (ICREA), Passeig Lluı́s Companys 23, E08010, Barcelona, Spain
d
Instituto de Fı́sica, Universidade de São Paulo, 05314-970 São Paulo, Brazil
e
Department of Physics, University of Wisconsin, Madison, WI 53706, USA
E-mail:
[email protected],
[email protected],
[email protected],
[email protected]
Abstract: We compute the screening length for quarkonium mesons moving through an
anisotropic, strongly coupled N = 4 super Yang-Mills plasma by means of its gravity
dual. We present the results for arbitrary velocities and orientations of the mesons, as well
as for arbitrary values of the anisotropy. The anisotropic screening length can be larger
or smaller than the isotropic one, and this depends on whether the comparison is made
at equal temperatures or at equal entropy densities. For generic motion we find that: (i)
mesons dissociate above a certain critical value of the anisotropy, even at zero temperature;
(ii) there is a limiting velocity for mesons in the plasma, even at zero temperature; (iii) in
the ultra-relativistic limit the screening length scales as (1 − v 2 )ǫ with ǫ = 1/2, in contrast
with the isotropic result ǫ = 1/4.
Keywords: Gauge-gravity correspondence, Holography and quark-gluon plasmas
Contents
1 Introduction
1
2 Gravity solution
4
3 Preliminaries
6
4 Static dipole in an anisotropic plasma
9
5 Dipole in an anisotropic plasma wind
5.1 Unbound quark-antiquark pair
5.1.1 Ultra-relativistic motion outside the transverse plane
5.1.2 Ultra-relativistic motion within the transverse plane
5.2 Bound quark-antiquark pair
5.2.1 Ultra-relativistic motion outside the transverse plane
5.2.2 Ultra-relativistic motion within the transverse plane
5.3 Isotropic limit
5.4 Numerical results for generic velocities
11
14
15
16
17
18
19
21
21
6 Dissociation temperature and dissociation anisotropy
24
7 Discussion
34
1
Introduction
A remarkable conclusion from the experiments at the Relativistic Heavy Ion Collider
(RHIC) [1, 2] and at the Large Hadron Collider (LHC) [3] is that the quark-gluon plasma
(QGP) does not behave as a weakly coupled gas of quarks and gluons, but rather as a
strongly coupled fluid [4, 5]. This places limitations on the applicability of perturbative
methods. The lattice formulation of Quantum Chromodynamics (QCD) is also of limited
utility, since for example it is not well suited for studying real-time phenomena. This has
provided a strong motivation for understanding the dynamics of strongly coupled nonAbelian plasmas through the gauge/string duality [6–8] (see [9] for a recent review of
applications to the QGP). In general, a necessary requirement for the string description to
2 N → ∞. Of course,
be tractable is that the plasma be infinitely strongly coupled, λ = gYM
c
the real-world QGP is not infinitely strongly coupled, and its dynamics involves a complex
combination of both weak and strong coupling physics that depend on the possibly multiple
scales that characterize the process of interest. The motivation for studying string models
is that they provide examples in which explicit calculations can be performed from first
–1–
principles at strong coupling, in particular in the real-time domain. The hope is then that,
by understanding the weak and the strong coupling limits, one may be able to bracket the
dynamics of the real-world QGP, which lies somewhere in between.
During the initial stage after the collision the plasma is far from equilibrium, and after
a certain time a hydrodynamic description becomes applicable. If one thinks of hydrodynamics as a gradient expansion around a locally isotropic system, it is somewhat surprising
that the hydrodynamic description actually becomes applicable when the longitudinal and
transverse pressures are still significantly different. This can be explicitly seen, for example, in holographic descriptions [10–13] in which gravity provides a valid description all
the way from the far-from-equilibrium phase to the locally isotropic phase, across the intermediate hydrodynamic-but-still-anisotropic phase. Thus, during most of the time that
viscous hydrodynamics is applied, the plasma created in a heavy ion collision is anisotropic,
with the level of anisotropy in fact increasing as one approaches the edge of the system.
The fact that the range of time and space over which the QGP is anisotropic is larger
than traditionally assumed has provided additional motivation for the study of anisotropic
plasmas.
In this paper we will investigate the effect of an intrinsic anisotropy on the screening
length between a quark-antiquark pair in a strongly coupled plasma. As we will review
below, the plasma is static because it is held in anisotropic equilibrium by an external
force [14, 15]. We will discuss all the caveats in more detail below, but we emphasize from
the beginning that there are several reasons why, in terms of potential extrapolations to
the real-world QGP, our results must be interpreted with caution. First, the sources of
anisotropy in the QGP created in a heavy ion collision and in our system are different. In
the QGP the anisotropy is dynamical in the sense that it is due to the initial distribution
of particles in momentum space, which will evolve in time and eventually become isotropic.
In contrast, in our case the anisotropy is due to an external source that keeps the system
in an equilibrium anisotropic state that will not evolve in time. Nevertheless, we hope that
our system might provide a good toy model for processes whose characteristic time scale
is sufficiently shorter than the time scale controlling the evolution of a dynamical plasma.
The second caveat concerns the fact that, even in an static situation, different external
sources can be chosen to hold the plasma in equilibrium, so one may wonder to what extent
the results depend on this choice. We will provide a partial answer to this question in Sec. 7,
where we will explain that our qualitative results, for example the ultrarelativistic limit, do
not depend on the details of our solution but only on a few general features. Nevertheless,
it would still be very interesting to compute the same observables in other strongly coupled,
static, anisotropic plasmas. Only then a general picture would emerge that would allow
one, for example, to understand which observables are robust, in the sense that they are
truly insensitive to the way in which the plasma is held in anisotropic equilibrium, and
which ones are model-dependent. Obviously it is the first type of observables that have a
better chance of being relevant for the real-world QGP. Our paper should be regarded as
a first step in this general program.
–2–
We will consider the screening length in the case in which the quark-antiquark pair is
at rest in the plasma as well as the case in which it is moving through the plasma. For this
purpose we will examine a string with both endpoints on the boundary of an asymptotically
AdS spacetime [14, 15] that is dual to an anisotropic N = 4 super Yang-Mills plasma. The
gravity solution possesses an anisotropic horizon, it is completely regular on and outside the
horizon, and it is solidly embedded in type IIB string theory. For these reasons it provides
an ideal toy model in which questions about anisotropic effects at strong coupling can be
addressed from first principles. For the particular case of a quark-antiquark pair at rest,
the screening length has also been computed [16] in a different model [17] of a strongly
coupled, anisotropic plasma. The results exhibit some differences with respect to those
presented here. While this may indicate some model dependence of the screening length,
it is important to note that the solution of [17] possesses a naked singularity. Although
this is a rather benign singularity, its presence introduces a certain amount of ambiguity in
the calculations, which can only be performed by prescribing somewhat ad hoc boundary
conditions at the singularity. In any case, this discussion is another indication that it
would be interesting to compute the screening length in a larger class of models in order
to ascertain which of its features are model-independent.
To avoid any possible confusion, we clarify from the beginning that the quarks and
antiquarks that we will consider are infinitely massive, i.e. the bound states that we will
consider are the analogue of heavy quarkonium mesons in QCD. Thus, the reader should
always have the word ‘quarkonium’ in mind despite the fact that we will often refer to these
states simply as ‘mesons’, ‘heavy mesons’, ‘quark-antiquark bound states’, ‘dipoles’, etc.
This is specially relevant in the ultra-relativistic limit of the screening length, to which
we will pay particular attention since it can be determined analytically. We emphasize
that our results correspond to sending the quark and antiquark masses to infinity first,
and then sending v → 1. In particular, this means that in any future attempt to connect
our results to the phenomenology of the QGP, this connection can only be made to the
phenomenology of heavy quarkonium moving through the plasma.
The screening length for quarkonium mesons at rest in the anisotropic plasma of [14, 15]
has been previously studied in [16, 18]. Our Sec. 4 has some overlap with these references
and, wherever they overlap, our results agree with theirs. Other physical properties of the
anisotropic plasma that have been calculated include its shear viscosity [19, 20], the drag
force on a heavy quark [18, 21], the jet quenching parameter [16, 18, 22], and the energy
lost by a rotating quark [23]. The phase diagram of the zero-coupling version of the model
considered in [14, 15] has been studied in [24]. Dissociation of baryons in the isotropic
N = 4 plasma has been analyzed in [25].
–3–
H
1.2
1
B
0.8
FB
0.6
0.4
F
1.2
6
1
5
0.8
4
0.6
3
0.4
0.2
0.2
0
φ
0.2
2
FB
1
B
0
-1
-0.2
0.4
0.6
0.8
1
-2
6
5
4
0
-0.2
0
H
0
0.2
u/uH
0.4
3
F
2
1
0
φ
0.6
0.8
-1
1
-2
u/uH
Figure 1. Metric functions for a/T ≃ 4.4 (left) and a/T ≃ 86 (right).
2
Gravity solution
The type IIB supergravity solution of [14, 15] in the string frame takes the form
1
du2
L2
2
2
2
2
2
+ L2 e 2 φ dΩ25 ,
ds = 2 −FB dt + dx + dy + Hdz +
u
F
χ = az ,
φ = φ(u) ,
(2.1)
(2.2)
where χ and φ are the axion and the dilaton, respectively, and (t, x, y, z) are the gauge
theory coordinates. Since there is rotational invariance in the xy-directions, we will refer
to these as the transverse directions, and to z as the longitudinal direction. F, B and H
are functions of the holographic radial coordinate u that were determined numerically in
[14, 15]. Their form for two values of a/T is plotted in Fig. 1. The horizon lies at u = uH ,
where F = 0, and the boundary at u = 0, where F = B = H = 1 and φ = 0. The
metric near the boundary asymptotes to AdS5 × S 5 . Note that the axion is linear in the
z-coordinate. The proportionality constant a has dimensions of mass and is a measure of
the anisotropy. The axion profile is dual in the gauge theory to a position-dependent theta
parameter of the form θ ∝ z. This acts as an isotropy-breaking external source that forces
the system into an anisotropic equilibrium state.
If a = 0 then the solution reduces to the isotropic black D3-brane solution dual to the
isotropic N = 4 theory at finite temperature. In this case
B = H = 1,
F =1−
χ = φ = 0,
u4
,
u4H
uH =
1
πT
(2.3)
and the entropy density takes the form
siso =
π2 2 3
N T .
2 c
(2.4)
Fig. 2 shows the entropy density per unit 3-volume in the xyz-directions of the anisotropic
plasma as a function of the dimensionless ratio a/T , normalized to the entropy density of
–4–
1.2
log(s/siso )
1.0
0.8
0.6
0.4
0.2
0.0
-4
-2
0
2
4
log(a/T )
Figure 2. Log-log plot of the entropy density per unit 3-volume in the xyz-directions as a function
of a/T , with siso defined as in eqn. (2.4). The dashed blue line is a straight line with slope 1/3.
the isotropic plasma at the same temperature. At small a/T the entropy density scales as
in the isotropic case, whereas at large a/T it scales as [14, 15, 26]
s = cent Nc2 a1/3 T 8/3 ,
[a/T ≫ 1]
(2.5)
where cent is a constant that can be determined numerically. The transition between the
two asymptotic behaviors of the entropy density takes place at a/T ≃ 3.7.
For later use we list here the near-boundary behavior of the different functions that
determine the solution (2.2):
7 4
11 2 2
F = 1 + a u + F4 + a log u u4 + O(u6 ) ,
24
12
11 2 2
7 4
B = 1 − a u + B4 − a log u u4 + O(u6 ) ,
24
12
5 4 1 4
2
1 2 2
(2.6)
B4 −
a − a log u u4 + O(u6 ) .
H = 1+ a u −
4
7
4032
6
The coefficients F4 and B4 depend on a and T and are known analytically in the limits of
low, and high temperature and numerically for intermediate regimes [15].
A feature of the solution (2.2) that played an important role in the analysis of [14, 15]
is the presence of a conformal anomaly. Its origin lies in the fact that diffeomorphism
invariance in the radial direction u gets broken in the process of renormalization of the onshell supergravity action. In the gauge theory this means that scale invariance is broken
by the renormalization process. One manifestation of the anomaly is the fact that, unlike
the entropy density, other thermodynamic quantities do not depend solely on the ratio a/T
but on a and T separately. Fortunately, this will not be the case for the screening length,
as we will see below.
To facilitate a (rough) comparison of the anisotropy in our system to that in other
anisotropic plasmas it is useful to consider the ratio
α=
4E + P⊥ − PL
,
3T s
–5–
(2.7)
1
5
10
15
20
a/T
-2
-4
α
-6
-8
-10
-12
Figure 3. Ratio (2.7) as a function of a/T . The blue dots are the actual values of the ratio, and
the red curve is the fit (2.8).
where E is the energy density and P⊥ , PL are the transverse and longitudinal pressures,
respectively. In addition to being dimensionless, this ratio has the virtue that it does not
depend on a and T separately, but only on the combination a/T . For the isotropic N = 4
super Yang-Mills plasma α = 1, whereas for 0 < a/T . 20 the ratio is well approximated
by the expression
a 4
a 2
− 0.000072
,
(2.8)
α ≃ 1 − 0.0036
T
T
as shown in Fig. 3.
At various points we will refer to the limit T = 0 of the anisotropic plasma. The zerotemperature version of the solution (2.2) was found in [26]. In this case the string-frame
metric exhibits a naked curvature singularity deep in the infra-red, and the Einstein-frame
metric exhibits infinite tidal forces [27, 28]. However, we emphasize that, for any finite
temperature, the singularity is hidden behind the horizon and the solution is completely
regular on and outside the horizon, exhibiting no pathologies of any type. Thus we will
think of the T = 0 results as those obtained by taking the limit T → 0 of the finitetemperature results. Moreover, regulating the infra-red geometry in this or any other way
is actually unnecessary for most of the physics of quarkonium dissociation. The reason is
that, as we will see, in the limit in which a/T becomes large the penetration depth into
the AdS bulk of the string that is dual to the quarkonium meson becomes very small. As
a result, the dissociation is entirely controlled by the metric near the boundary, which is
insensitive to the infra-red behavior described above.
3
Preliminaries
In this paper we define the screening length Ls as the separation between a quark and
an antiquark such that for ℓ < Ls (ℓ > Ls ) it is energetically favorable for the quarkantiquark pair to be bound (unbound) [29, 30]. Obviously this satisfies Ls ≤ Lmax , where
Lmax is the maximum separation Lmax for which a bound quark-antiquark solution exists.
We will determine Ls by comparing the action S(ℓ) of the bound pair, which is a function
–6–
0.0
√
∆Edipole /T λ
-0.5
-1.0
-1.5
-2.0
-2.5
-3.0
-3.5
0.00 0.05 0.10 0.15 0.20 0.25
Tℓ
Figure 4. Energy difference, as defined in (3.2), between a bound and an unbound quark-antiquark
pair moving through the isotropic plasma (2.3) with velocities (from the rightmost curve to the
leftmost curve) v = 0, 0.35, 0.85, 0.996. The dipole is oriented orthogonally to its velocity. For
v < vtrans one has Ls < Lmax , whereas for v > vtrans one finds L = Lmax , where vtrans ≃ 0.45 is the
transition velocity between the two behaviors. At v = 0 the screening length and the maximum
separation are Ls ≃ 0.24/T and Lmax ≃ 0.27/T , respectively.
of the quark-antiquark separation ℓ, to the action Sunbound of the unbound system, i.e. by
computing:
∆S(ℓ) = S(ℓ) − Sunbound .
(3.1)
The screening length is the maximum value of ℓ for which ∆S is positive (since we will
work in Lorentzian signature). This may correspond to the value of ℓ at which ∆S crosses
zero, in which case Ls < Lmax , or the maximum value of ℓ for which a bound state exists,
in which case Ls = Lmax . In the Euclidean version of our calculations, this criterion corresponds to determining which configuration has the lowest free energy, which is therefore
the configuration that is thermodynamically preferred. As shown in Fig. 4, for a meson
moving through the isotropic plasma (2.3) one has Ls < Lmax for v < vtrans , whereas for
v > vtrans one finds that Ls = Lmax , where vtrans ≃ 0.45 is the transition velocity between
the two behaviors [31–33]. These qualitative features extend to the anisotropic case, as we
have illustrated in Fig. 5. The transition velocity decreases with the anisotropy, so for large
a/T one has Ls = Lmax except for very low velocities. Similarly, if the ultra-relativistic
limit v → 1 is taken at fixed a and T , then obviously v > vtrans and again Ls = Lmax .
All our calculations will be done in the rest frame of the quark-antiquark pair, to
which we will refer as the dipole rest frame. Since any observable can be easily translated
between this frame and the plasma rest frame, we will speak interchangeably of ‘mesons in
a plasma wind’ and of ‘mesons in motion in the plasma’. We emphasize however that all
the physical quantities that we will present, e.g. the screening length, are computed in the
dipole rest frame.
The actions are scalar quantities, so ∆Sdipole = ∆Splasma . Moreover, in the dipole rest
–7–
0.1
-2
0.0
-4
-0.1
√
∆Edipole /T λ
√
∆Edipole /T λ
0
-6
-8
-0.2
-0.3
-10
-0.4
-12
0.00 0.05 0.10 0.15 0.20 0.25
Tℓ
-0.5
0.00 0.05 0.10 0.15 0.20 0.25
Tℓ
Figure 5. Energy difference in an anisotropic plasma, as defined in (3.2), between a bound and
an unbound quark-antiquark pair oriented along the transverse direction x and moving along the
anisotropic direction z. All the curves on the left correspond to a/T = 12.2 and different velocities
(from the rightmost curve to the leftmost curve) v = 0, 0.35, 0.85, 0.996. All the curves on the right
correspond to the same velocity v = 0.25 and different anisotropies (from the rightmost curve to the
leftmost curve) a/T = 0, 6.5, 43, 744. For these anisotropies the corresponding transition velocities
are respectively given by vtrans = 0.45, 0.29, 0.19, 0.11.
frame we have
∆Sdipole = −T ∆Edipole ,
(3.2)
since the dipole is static in its own rest frame. In this expression Edipole is the energy
R
(as opposed to the free energy) of the configuration and T = dt is the length of the
integration region in time. Thus we see that our criterion, which is based on comparing
the actions, can also be thought of as a comparison between the energies of the bound and
the unbound configurations in the dipole rest frame.
We will see that the ultraviolet divergences in the string action associated to integrating
all the way to the boundary of AdS cancel out in the difference (3.1), and neither the bound
nor the unbound actions possess infrared divergences associated to integrating all the way
down to the horizon. This can be verified explicitly and it also follows from their relation
to the energy in the rest frame of the dipole: While the energy of the unbound string
pair possesses an infrared logarithmic divergence in the plasma rest frame [34], no such
divergence is present in the dipole rest frame (see e.g. the discussion in [33]).
–8–
4
Static dipole in an anisotropic plasma
In an anisotropic plasma the screening length depends on the relative orientation between
the dipole and the anisotropic direction z. Given the rotational symmetry in the xy-plane
we assume without loss of generality that the dipole lies in the xz-plane, at an angle θ with
the z-axis. We thus choose the static gauge t = τ, σ = u and specify the string embedding
as
x → sin θ x(u) ,
z → cos θ z(u) .
(4.1)
The string action takes the form
Z
Z umax
q
L2
1
B 1 + FH cos2 θ z ′2 + F sin2 θ x′2 ,
S=−
du
2
dt
′
2
2πα
u
0
(4.2)
where the 2 comes from the two branches of the string and umax will be determined below.
The conserved momenta associated to translation invariance in the x, z directions are given
by
Πx =
Πz =
1 ∂L
BF sin θ x′
q
=
,
sin θ ∂x′
u2 B 1 + FH cos2 θ z ′2 + F sin2 θ x′2
BFH cos θ z ′
1 ∂L
q
=
.
cos θ ∂z ′
u2 B 1 + FH cos2 θ z ′2 + F sin2 θ x′2
Inverting these relations we find
√
H csc θ u2 Πx
′
x =√ p
,
F BFH − u4 (Π2z + H Π2x )
z′ = √
Substituting back in the action we arrive at
L2
2
S=−
2πα′
Z
dt
Z
umax
0
sec θ u2 Πz
p
.
FH BFH − u4 (Π2z + H Π2x )
√
1
B FH
du 2 p
.
u
BFH − u4 (Π2z + H Π2x )
(4.3)
(4.4)
(4.5)
(4.6)
For a U-shaped string describing a bound quark-antiquark pair the turning point umax
is determined in terms of the momenta by the condition that x′ (umax ) = z ′ (umax ) → ∞.
This happens if umax = uH , in which case F(umax ) = 0, or if
BFH − u4 Π2z + H Π2x
umax
= 0.
(4.7)
The first possibility is not physically relevant because the second possibility is always
realized first, meaning that the string turns around at umax < uH , before reaching the
horizon. The only exception is the case Πx = Πz = 0, but this corresponds to x′ = z ′ = 0,
namely to an unbound pair of strings that descend from the boundary straight down to
the horizon.
–9–
The momenta are determined by the boundary conditions that require the string endpoints to lie a distance ℓ apart from each other:
Z umax
Z umax
ℓ
′
du z ′ .
(4.8)
du x =
=
2
0
0
These two equations, together with (4.7), can be solved numerically to express the momenta
and umax in terms of ℓ. In this way the on-shell action (4.6) for a bound pair becomes a
function of ℓ alone. In order to determine Ls we subtract from this action the action of a
static, unbound quark-antiquark pair, which is described by two straight strings hanging
down from the boundary to the horizon. The action of this unbound pair is equal to (4.6)
with the momenta set to zero and the range of integration extended down to the horizon:
Z
Z uH √
L2
B
Sunbound = −
2 dt
(4.9)
du 2 .
′
2πα
u
0
We obtain the screening length by numerically determining the value of ℓ at which the
difference S(ℓ) − Sunbound crosses zero, since in the static case we always have Ls < Lmax .
The result for this difference as a function of ℓ in the isotropic plasma [29, 30] described
by eqn. (2.3) is plotted in Fig. 4, from which we see that the screening length is
0.24
Liso (T ) ≃
[static dipole] .
(4.10)
T
The scaling with the temperature is expected on dimensional grounds. In the isotropic case
the temperature and the entropy density are related simply through (2.4), so this result
can be recast as
2 2 1/3
π Nc
[static dipole] ,
(4.11)
Liso (s) ≃ 0.24
2s
which will be useful later.
The results in the anisotropic case are plotted in Figs. 6 and 7. Fig. 6 shows the
screening length, for several orientations of the dipole, as a function of the anisotropy
measured in units of the temperature (left) and the entropy density (right). The reason
for working with both normalizations is that we wish to compare the screening length
in the anisotropic plasma to that in the isotropic plasma, and this can be done at least
in two different ways: the two plasmas can be taken to have the same temperatures but
different entropy densities, or the same entropy densities but different temperatures. Fig. 7
shows the screening length as a function of the dipole orientation for several values of the
anisotropy.
We see from Fig. 6(left) that Ls decreases monotonically as a increases, for any dipole
orientation, if the temperature is kept fixed. We also see from Fig. 7(left) that this effect
is more pronounced for a dipole oriented along the anisotropic direction. In contrast,
the behavior of the screening length at constant entropy density depends on the dipole’s
orientation, as shown in Figs. 6(right) and 7(right). For dipole’s aligned sufficiently close
to the anisotropic direction the screening length decreases with the anisotropy, whereas for
orientations sufficiently close to the transverse plane the screening length increases with
the anisotropy.
– 10 –
0.9
Ls /Liso (s)
Ls /Liso (T )
1.0
0.8
0.7
0.6
0.5
0
5
10
15
20
25
30
1.2
1.1
1.0
0.9
0.8
0.7
0.6
0.5
0
5
10
15
20
25
30
aNc2/3 /s1/3
a/T
Figure 6. Screening length as a function of the anisotropy for a static quark-antiquark dipole
lying at an angle with the z-direction (from top to bottom on the right-hand side of the plot)
θ = π/2, π/3, π/4, π/6, 0. The screening length is plotted in the appropriate units to facilitate
comparison with the isotropic result for a plasma at the same temperature (left), or at the same
entropy density (right). The isotropic result is given in eqs. (4.10) and (4.11).
1
0.6
Ls /Liso (s)
Ls /Liso (T )
0.8
0.4
0.2
0
0
Π
8
Π
4
3Π
8
Π
2
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0
θ
Π
8
Π
4
3Π
8
Π
2
θ
Figure 7. Screening length for a quark-antiquark dipole lying at an angle θ with the z-direction
for anisotropies a/T = 12.2 (red, solid), 42.6 (maroon, coarsely dashed), 86 (violet, dashed), 744
(orange, dot-dashed). The corresponding values in units of the entropy density are (in the same
order) aNc2/3 /s1/3 = 6.2, 19, 35, 242. The screening length is plotted in the appropriate units to
facilitate comparison with the isotropic result for a plasma at the same temperature (left), or at
the same entropy density (right). The isotropic result is given in eqs. (4.10) and (4.11).
5
Dipole in an anisotropic plasma wind
In this section we will consider a static quark-antiquark pair in an anisotropic plasma that
is moving with constant velocity with respect to the dipole — a dipole in an ‘anisotropic
plasma wind’. We will pay particular attention to the ultra-relativistic limit, which can be
understood analytically.1 This limit, together with the static results from Sec. 4, will allow
us to understand qualitatively the results at any velocity 0 < v < 1.
We will first rewrite the solution (2.2) in a boosted frame, and then place a dipole in
1
We recall that we first send the quark mass to infinity and then v → 1 (see Sec. 1).
– 11 –
z
θv
θ
ϕ
y
x
Figure 8. Orientation of the dipole in an anisotropic plasma wind. The wind’s velocity lies in the
original xz-plane (before the boost (5.2)) at an angle θv with respect to the anisotropic direction z.
The quark lies at angles ~q = (x, y, z) = 2ℓ (sin θ sin ϕ, sin θ cos ϕ, cos θ) with respect to the relabeled
directions (after the boost (5.2)), and the antiquark lies at −~q.
it — see Fig. 8. Given the rotational symmetry in the xy-plane we assume that the boost
velocity is contained in the xz-plane, and that it lies at an angle θv with the z-axis. Thus
we first rotate to a new coordinate system defined through
t = t̃ ,
x = z̃ sin θv + x̃ cos θv ,
y = ỹ ,
z = z̃ cos θv − x̃ sin θv ,
(5.1)
and then perform a boost along the z̃-direction by setting
t̃ = γ t′ − v z ′ ,
x̃ = x′ ,
ỹ = y ′ ,
z̃ = γ −v t′ + z ′ ,
(5.2)
√
where γ = 1/ 1 − v 2 is the usual Lorentz factor. Below we will consider a dipole with an
arbitrary orientation with respect to both the velocity of the plasma and the anisotropic
direction z — see Fig. 8. We parametrize the orientation of the dipole by two angles θ, ϕ
so that the quark lies at
ℓ
~q = (x′ , y ′ , z ′ ) = (sin θ sin ϕ, sin θ cos ϕ, cos θ)
2
– 12 –
(5.3)
and the antiquark lies at −~q.
For notational simplicity, below we will drop the primes in the final set of coordinates.
To avoid confusion, we emphasize that the direction θv of the plasma wind is always
measured with respect to the original (x, y, z) axes, i.e. before the rotation and the boost
above. In particular, motion within (outside) the transverse plane refers to a dipole in
a plasma wind with θv = π/2 (θv 6= π/2). In contrast, the orientation of the dipole
is measured with respect to the final set of coordinates (x′ , y ′ , z ′ ). However, if instead
of specifying the dipole’s orientation through a pair (θ, ϕ) we specify it by saying that
the dipole is aligned with the x-, y- or z-directions, then we are referring to the original
directions. Just as an illustration, consider the case of a plasma wind blowing along the
original x-direction, i.e. a plasma wind with θv = π/2. Then we see from (5.1) and (5.2)
that (x, z) ∼ (z ′ , x′ ). Thus in this case by ‘a dipole oriented along the x-direction’ we mean
a dipole with θ = 0.
After dropping the primes from the final set of coordinates in (5.2) the five-dimensional
part of the metric (2.2) takes the form
du2
L2
2
2
2
2
2
ds = 2 −gtt dt + gxx dx + dy + gzz dz + gtx dt dx + gtz dt dz + gxz dx dz +
,
u
F
(5.4)
where
gtt =
BF − v 2 (sin2 θv + H cos2 θv )
,
1 − v2
gxx = cos2 θv + H sin2 θv ,
gzz =
sin2 θv + H cos2 θv − v 2 BF
,
1 − v2
(5.5)
(5.6)
(5.7)
(H − 1)v
sin(2θv ) ,
gtx = √
1 − v2
(5.8)
gtz =
(5.9)
2v(BF − sin2 θv − H cos2 θv )
,
1 − v2
1−H
gxz = √
sin(2θv ) .
1 − v2
(5.10)
In order to determine the screening length for a generic velocity we need to compare
the actions of a bound and an unbound quark-antiquark pair, as in the static case of
Sec. 4. However, in the ultra-relativistic this is not strictly necessary because Ls = Lmax
(see Sec. 3). In other words, in this limit we only need to determine the maximum possible
quark-antiquark separation for which a bound state exists. Nevertheless, for completeness
we will briefly present the analysis of the unbound configuration. Each of the strings in the
unbound pair is one of the trailing strings studied in [21], so the reader is referred to this
reference for additional details. Note, however, that [21] worked in the plasma rest frame.
Here we will work in the dipole’s rest and focus on the ultra-relativistic limit.
– 13 –
5.1
Unbound quark-antiquark pair
As in Sec. 4 we fix the static gauge t = τ , σ = u, and specify the embedding of the unbound
string as
x → x(u) ,
z → z(u) .
(5.11)
The embedding in the y-direction is simply y = 0 because of rotational symmetry in the
xy-plane and because the string is unbound. As we will see below, in the case of a bound
string (dipole) the boundary conditions will generically imply a non-trivial embedding y(u).
The action for the unbound string reads
Z
Z uH
L2
1p
Sunbound = −
2
dt
du 2 F −1 K0 + Kxx x′2 + Kzz z ′2 + Kxz x′ z ′ ,
′
2πα
u
0
(5.12)
where
K0 = gtt ,
BF(cos2 θv + H sin2 θv ) − Hv 2
,
1 − v2
= BF sin2 θv + H cos2 θv ,
Kxx =
Kzz
Kxz =
BF(1 − H)
√
sin(2θv ) .
1 − v2
(5.13)
Introducing the conjugate momenta
∂Lunbound
,
∂x′
Πz =
∂Lunbound
∂z ′
(5.14)
N
u2
√
√x ,
F BH D
z′ =
N
u2
√
√z ,
F BH D
(5.15)
Πx =
and solving for x′ , z ′ we find
x′ =
where
1
Nx = Kzz Πx − Kxz Πz ,
2
1
Nz = − Kxz Πx + Kxx Πz ,
2
D = BHFK0 − u4 Kzz Π2x + Kxx Π2z − Kxz Πx Πz .
(5.16)
Substituting into the action we arrive at
Sunbound
L2
2
=−
2πα′
Z
dt
Z
uH
du
0
√
BHK0
√
.
u2 D
(5.17)
The momenta are determined by the condition that (5.15) remain real for a string that
extends all the way from the boundary to the horizon. Following [21] we analyze this
condition by noting that D can be rewritten as
h
ih
i
2u4
D=
Nx Nz − b Πx Πz − c BF − v 2 (sin2 θv + H cos2 θv )
(5.18)
Kxz
– 14 –
where
b=
Hu4
,
(1 − H) 1 − v 2 sin θv cos θv
√
c=
BF(1 − H) sin θv cos θv
√
.
u4 1 − v 2
(5.19)
As in [21] we must require that the zeros of the second summand in (5.18) coincide with
one another and with those of Nx and Nz . One of the zeros of the second summand occurs
at a critical value u = uc such that
Bc Fc − dc v 2 = 0 ,
dc ≡ Hc cos2 θv + sin2 θv ,
(5.20)
where Bc = B(uc ), etc. At this point we have
N x N z | uc =
h
i2
(Hc − 1) cos θv sin θv
v 4 cos θv sin θv
√
√
(Hc − 1) dc dc Πx +
Πz . (5.21)
1 − v2
1 − v2
Noting that Hc > 1 and that Kxz < 0, we see that D would be negative at uc unless the
momenta are related through
Πx =
(1 − Hc ) cos θv sin θv
√
Πz .
dc 1 − v 2
(5.22)
Assuming this relation and requiring that the other zero in the second summand of (5.18)
coincide with uc yields
Π2z =
Bc Fc dc
,
u4c
Π2x =
Bc Fc (Hc − 1)2 cos2 θv sin2 θv
.
u4c (1 − v 2 )dc
(5.23)
Note that Πz does not vanish for any value of θv , whereas Πx vanishes if θv = 0, π/2.
The reason is that for these two particular orientations the plasma wind blows along the
original z- or x-directions and the string orients itself with the corresponding axis [21]. As
a consequence, the momentum along the orthogonal axis vanishes. However, the changes
of coordinates (5.1) and (5.2) always relabel the direction of motion as z, so after these
changes the non-vanishing momentum is labelled Πz irrespectively of whether θv = 0 or
θv = π/2.
We will analyze in detail the ultra-relativistic limit. This is facilitated by explicitly
distinguishing the case of motion outside the transverse plane (θv 6= π/2) and motion
within the transverse plane (θv = π/2).
5.1.1
Ultra-relativistic motion outside the transverse plane
In the ultra-relativistic limit uc approaches the boundary, i.e. uc → 0, and we can use the
near-boundary expansion (2.6) to determine it. The condition (5.20) yields in this limit
[21]
4(1 − v 2 )
u2c ≃ 2
[θv 6= π/2] ,
(5.24)
a cos2 θv
which when substituted in (5.23) gives the momenta
Π2z ≃
a4 cos4 θv
,
16(1 − v 2 )2
Π2x ≃
– 15 –
a4 cos2 θv sin2 θv
.
16(1 − v 2 )
(5.25)
In these expressions we have ignored subleading terms in an expansion in 1−v 2 , for example
we have set v ≃ 1, Hc ≃ 1, etc. Note that in this expansion Πx is subleading with respect
to Πz .
For later use we must evaluate how Sunbound scales with 1 − v 2 in the limit v → 1. For
this purpose we split the integration region, and hence the action (5.17), as
(1)
(2)
Sunbound = Sunbound + Sunbound ,
(1)
(5.26)
(2)
where Sunbound is the action with the integral in u ranging between 0 and uc , and Sunbound is
the action with the integral in u ranging between uc and uH . The reason for this separation
is that in the first interval u is small and hence we will be able to use the near-boundary
(1)
expressions (2.6), (5.24) and (5.25). In order to exhibit the dependence on 1 − v 2 of Sunbound
explicitly, it is convenient to work with a rescaled variable r which remains finite in the
v → 1 limit, defined though
p
p
uc = r c 1 − v 2 .
(5.27)
u = r 1 − v2 ,
In terms of this variable we get
Z
Z rc
1 − 14 a2 r2 cos2 θv + . . .
2
L2
(1)
q
√
.
dr
dt
Sunbound = −
2πα′ 1 − v 2
1 4 4
0
a r cos4 θv + . . .
r2 1 − 41 a2 r2 cos2 θv − 16
(5.28)
The divergence near r = 0 will cancel out with that in the action for the bound string. The
integrand is smooth across r = rc . The crucial point is that the result is O (1 − v 2 )−1/2
in the counting in powers of 1 − v 2 , and we will find this same scaling in the bound string
(2)
action (see below). In contrast, Sunbound scales as 1 − v 2 in the ultra-relativistic limit. The
reason is that u is not small in units of 1 − v 2 in the corresponding region of integration,
so all the dependence comes from the fact that the action (5.17) scales as 1/Πz ∼ 1 − v 2
in this region.
5.1.2
Ultra-relativistic motion within the transverse plane
In this case θv = π/2 and hence we see from (5.23) that Πx = 0. The condition (5.20) now
gives [21]
r
1 − v2
2
,
(5.29)
uc ≃
C
where
121 4
C=
a − F4 − B4 ,
(5.30)
576
and we recall that F4 , B4 are the coefficients that enter the near-boundary expansion (2.6).
Substituting (5.29) into (5.23) and dropping subleading terms as before we obtain the
momentum in the z-direction (recall that this corresponds to the original x-direction):
r
C
1
Πz ≃ 2 =
.
(5.31)
uc
1 − v2
– 16 –
It is now convenient to work with a rescaled radial coordinate r defined through
u = r(1 − v 2 )1/4 .
(5.32)
Splitting the unbound string action as before, we find
Z
Z rc
1 − Cr4 + . . .
2
L2
(1)
√
Sunbound = −
dr
dt
.
2πα′ (1 − v 2 )1/4
r2 1 − 2Cr4 + . . .
0
(5.33)
Again, the divergence near r = 0 will cancel out with that in the action for the bound string,
which will also be of O (1 − v 2 )−1/4 in the counting in powers of 1 − v 2 (see below). In
√
(2)
contrast, Sunbound scales as 1/Πx ∼ 1 − v 2 in the ultra-relativistic limit, and is therefore
subleading.
5.2
In summary, we find that in the ultra-relativistic limit
2 −1/2 if θ 6= π/2
[outside the transverse plane]
v
O (1 − v )
(5.34)
Sunbound =
O (1 − v 2 )−1/4 if θ = π/2
[within the transverse plane] .
v
Bound quark-antiquark pair
We now consider a dipole with an arbitrary orientation with respect to both the velocity
of the plasma and the anisotropic direction z — see Fig. 8. As before we fix the static
gauge τ = t, σ = u and specify the string embedding via three functions (x(u), y(u), z(u))
subject to the boundary conditions
Z umax
ℓ
x′ du ,
sin θ sin ϕ =
2
Z0 umax
ℓ
y ′ du ,
sin θ cos ϕ =
2
Z0 umax
ℓ
cos θ =
z ′ du ,
(5.35)
2
0
where umax is the turning point of the U-shaped string. The integral in the action of the
bound string extends only up to this point and now includes a term proportional to y ′2 :
Z
Z umax
q
1
L2
du
2
dt
S =−
F −1 K0 + Kxx x′2 + Kyy y ′2 + Kzz z ′2 + Kxz x′ z ′ .
2πα′
u2
0
(5.36)
All the K’s were defined in (5.13) except for Kyy , which is given by
Kyy =
BF − v 2 (sin2 θv + H cos2 θv )
.
1 − v2
(5.37)
The momenta are defined as
Πx =
∂L
,
∂x′
Πy =
∂L
,
∂y ′
– 17 –
Πz =
∂L
.
∂z ′
(5.38)
Inverting these equations we get
u2
1
√
x = √
Kzz Πx − Kxz Πz ,
2
F BH D
√
u2 BH
y′ = √
Πy ,
D
1
u2
′
√
− Kxz Πx + Kxx Πz ,
z = √
2
F BH D
′
where
D = BHFK0 − u4 Kzz Π2x + BFH Π2y + Kxx Π2z − Kxz Πx Πz .
Substituting these expressions into the action (5.36) we get
√
Z
Z umax
BHK0
L2
√
du
2 dt
S=−
.
2πα′
u2 D
0
(5.39)
(5.40)
(5.41)
As in the case of the unbound string, we will now distinguish between the cases of motion
outside and within the transverse plane, focusing on the ultra-relativistic limit.
5.2.1
Ultra-relativistic motion outside the transverse plane
The turn-around point umax is defined by the condition D(umax ) = 0. In the ultra-relativistic
limit we expect that this point approaches the boundary for the string solution of interest,
as in the isotropic case. Thus in this limit umax can be determined by using the nearboundary expansions of the metric functions (2.6).
In the limit u → 0 we find the following expansions:
a2 u2 cos2 θv
+ ··· ,
4
a2 u2 sin θv cos θv
√
≃ 0−
+ ··· ,
2 1 − v2
Kzz ≃ 1 +
(5.42)
Kxz
(5.43)
Kxx ≃ 1 −
from which it follows that
D ≃1−
a2 u2 cos2 θv
+ ··· ,
4(1 − v 2 )
a2 u2 cos2 θv
− u4 (Π2x + Π2y + Π2z ) + · · · .
4(1 − v 2 )
Similarly, the boundary conditions (5.35) take the form
Z umax
ℓ
u2
sin θ sin ϕ ≃
du √ Πx + · · · ,
2
D
0
Z umax
ℓ
u2
sin θ cos ϕ ≃
du √ Πy + · · · ,
2
D
0
Z umax
ℓ
a2 u2 cos2 θv
u2
√
1−
du
cos θ ≃
Πz + · · · ,
2
4(1 − v 2 )
D
0
– 18 –
(5.44)
(5.45)
(5.46)
In the ultra-relativistic limit, all the terms that we have omitted in the equations above,
in particular in (5.45) and (5.47), are subleading with respect to the terms that we have
retained provided the radial coordinate and the momenta scale as
p
pi
Πi =
,
(5.47)
u = r 1 − v2 ,
1 − v2
where r and pi are kept fixed in the limit v → 1. In terms of these rescaled variables (5.47)
the boundary conditions (5.47) take the form
p
ℓ
sin θ sin ϕ ≃ 1 − v 2 px I2 (p, θv ) ,
2
p
ℓ
sin θ cos ϕ ≃ 1 − v 2 py I2 (p, θv ) ,
2
p
a2 cos2 θv
ℓ
cos θ ≃ 1 − v 2 pz I2 (p, θv ) −
I4 (p, θv ) ,
(5.48)
2
4
where the integral
In (p, θv ) ≡
Z
rmax
0
dr q
1−
rn
a2 r 2
4
(5.49)
cos2 θv − r4 (p2x + p2y + p2z )
is of O(1) in the counting in powers in (1 − v 2 ), and is finite if n ≥ 0. Further noting that
K0 = 1 −
a2 u2 cos2 θv
a2 r2 cos2 θv
4
+
O(u
)
≃
1
−
,
4(1 − v 2 )
4
we see that the bound action scales as
Z
a2 cos2 θv
L2
2
√
I
(p,
θ
)
−
S ≃−
I
(p,
θ
)
dt .
−2
v
0
v
2πα′ 1 − v 2
4
(5.50)
(5.51)
Since both this bound action and the unbound action (5.28) scale as (1 − v 2 )−1/2 , the
divergence at r = 0 in the bound action coming from the I−2 (p, θv ) integral would exactly
cancel that in the unbound action in the difference (3.1). Moreover, by comparing the two
actions we would conclude that the momenta pi introduced in (5.47) are indeed of O(1) in
the counting in powers of (1 − v 2 ) in the ultra-relativistic limit. It would then follow that
the integrals In (p, θv ) are also of O(1), and therefore that the screening length scales as
Ls ∼ (1 − v 2 )1/2 in the ultra-relativistic limit. However, as explained below (5.10), in the
ultra-relativistic Ls = Lmax is simply the maximum possible separation between a bound
quark-antiquark pair, so it can be determined by maximizing ℓ in (5.48) with respect to
the momenta. Since the integrals are bounded from above for any value of the pi , and the
maximum is v-independent, it follows that Ls = Lmax ∼ (1 − v 2 )1/2 .
5.2.2
Ultra-relativistic motion within the transverse plane
In this case θv = π/2 and the expansions of D and of the boundary conditions (5.35)
become
Cu4
− u4 (Π2x + Π2y + Π2z ) + · · ·
(5.52)
D ≃1−
1 − v2
– 19 –
and
ℓ
sin θ sin ϕ ≃
2
Z
ℓ
sin θ cos ϕ ≃
2
Z
ℓ
cos θ ≃
2
Z
umax
0
umax
0
umax
0
du u2 q
du u2 q
du u2 q
Πx
1−
Cu4
1−v 2
− u4 (Π2x + Π2y + Π2z )
Πy
1−
Cu4
1−v 2
1−
Cu4
1−v 2
− u4 (Π2x + Π2y + Π2z )
Cu4
Πz
1 − 1−v
2
−
u4 (Π2x
+
Π2y
+
Π2z )
+ ··· ,
+ ··· ,
+ ··· ,
where C was defined in (5.30). As in the previous section, in the ultra-relativistic limit all
the terms that we have omitted in the equations above are subleading with respect to the
terms that we have retained provided the radial coordinate and the momenta scale in this
case as
u = r(1 − v 2 )1/4 ,
Πi = √
pi
,
1 − v2
(5.53)
where r and pi are kept fixed in the limit v → 1. In terms of the rescaled variables the
boundary conditions (5.53) become
ℓ
sin θ sin ϕ ≃ (1 − v 2 )1/4 px J2 (p) ,
2
ℓ
sin θ cos ϕ ≃ (1 − v 2 )1/4 py J2 (p) ,
2
ℓ
cos θ ≃ (1 − v 2 )1/4 pz (J2 (p) − CJ6 (p)) ,
2
(5.54)
where the integral
Jn (p) =
Z
rmax
0
rn
dr q
1 − r4 (C + p2x + p2y + p2z )
(5.55)
is of O(1) in the counting in powers in (1 − v 2 ), and is finite if n ≥ 0. Further noting that
K0 = 1 −
C
u4 + O(u6 ) ≃ 1 − Cr4 ,
1 − v2
(5.56)
we see that the bound action becomes
L2
2
S≃−
J
(p)
−
CJ
(p)
−2
2
2πα′ (1 − v 2 )1/4
Z
dt .
(5.57)
Since both this bound action and the unbound action (5.33) scale as (1 − v 2 )−1/4 , the
divergence at r = 0 in the bound action coming from the J−2 (p) integral would exactly
cancel that in the unbound action in the difference (3.1). Moreover, by comparing the two
actions we would conclude that the momenta pi introduced in (5.53) are indeed of O(1)
in the counting in powers of (1 − v 2 ) in the ultra-relativistic limit. It would then follow
– 20 –
that the integrals Jn (p) are also of O(1), and therefore that the screening length scales as
Ls ∼ (1 − v 2 )1/4 in the ultra-relativistic limit. However, as explained below (5.10), in the
ultra-relativistic Ls = Lmax is simply the maximum possible separation between a bound
quark-antiquark pair, so it can be determined by maximizing ℓ in (5.54) with respect to
the momenta. Since the integrals are bounded from above for any value of the pi , and the
maximum is v-independent, it follows that Ls = Lmax ∼ (1 − v 2 )1/4 .
In summary, we conclude that in the dipole rest frame the screening length scales in
the ultra-relativistic limit as
2 1/2 if θ 6= π/2
[motion outside the transverse plane]
v
(1 − v )
(5.58)
Ls ∼
(1 − v 2 )1/4 if θ = π/2
[motion
within
the
transverse
plane]
v
irrespectively of the dipole orientation.
5.3
Isotropic limit
The results above reduce to the isotropic result of Ref. [31, 32] in the limit a → 0. This
limit is most easily recovered from the results for motion within the transverse plane, since
some of the terms in the expansions in Section 5.2.1 vanish if a = 0, thus invalidating the
analysis. In contrast, setting a = 0 in Section 5.2.2 boils down to simply setting C to its
isotropic value, which from (5.30) and (2.3) is
C = −F4 =
1
= π4T 4 .
u4H
(5.59)
Since the value of C does not affect the ultra-relativistic scaling of the screening length,
we recover the scaling
Liso ∼ (1 − v 2 )1/4
[isotropic plasma]
(5.60)
found in the isotropic case by the authors of [31, 32]. As in the anisotropic case, the
ultra-relativistic scaling of the screening length is independent of the dipole’s orientation.
In fact, even for v < 1, the isotropic screening length depends only mildly on the dipole’s
orientation, as shown in Fig. 9.
5.4
Numerical results for generic velocities
Away from the ultra-relativistic limit the screening length must be obtained numerically.
For this reason we have focused on a few representative cases, namely those in which both
the direction of the plasma wind and the dipole’s orientation are aligned with one of the
original x, y, or z axes. Given the rotational symmetry in the xy-plane, there are only five
inequivalent cases to consider, because if the wind ‘blows’ in the z-direction then orienting
the dipole along x or y gives identical physics. In each case, we plot the screening length
– 21 –
T Liso
0.25
0.20
0.15
0.10
0.05
0.00
0.0
0.2
0.4
v
0.6
0.8
1.0
Figure 9. Screening length for a dipole moving through an isotropic plasma in a direction orthogonal (top, blue curve) or parallel (bottom, orange curve) to its orientation.
both as a function of the velocity v for different degrees of anisotropy a, and also as a
function of the degree of anisotropy for different values of the velocity. In each case the
result can be qualitatively understood combining the static results from Sec. 4 and the
ultra-relativistic behavior derived analytically in Section 5. We recall that in all cases
below, by ‘a dipole oriented along x, y or z’ we are referring to the original directions
before the rotation (5.1) and the boost (5.2).
Wind along z and dipole along z. The numerical results are shown in Figs. 10 and 11.
The curves in Fig. 10 start at v = 0 with the same value as the θ = 0 static result shown in
Fig. 7, and that they vanish as (1 − v 2 )1/4 in the limit v → 1, in agreement with (5.58)(top
line) and (5.60). The screening length decreases with the anisotropy, irrespectively of
whether T or s are kept fixed.
Wind along z and dipole along x. The numerical results are shown in Figs. 12 and
13. We see that the curves in Fig. 12 start at v = 0 with the same value as the θ = π/2
static result shown in Fig. 7, and that they vanish as (1 − v 2 )1/4 in the limit v → 1, in
agreement with (5.58)(top line) and (5.60). In this case the screening length decreases with
the anisotropy for any velocity provided the temperature is kept fixed. The same behavior
is found at constant entropy density for high enough velocities, whereas for low velocities
the screening length at constant s actually increases with a.
Wind along x and dipole along x. The numerical results are shown in Figs. 14 and
15. The curves in Fig. 14 start at v = 0 with the same value as the θ = π/2 static result
shown in Fig. 7, and that they approach a finite, non-zero value as v → 1, in agreement
with (5.58)(bottom line) and (5.60). As in previous cases, the screening length decreases
with the anisotropy for any velocity provided the temperature is kept fixed. The opposite
behavior is found at constant s.
Wind along x and dipole along y. The numerical results are shown in Figs. 16 and
17. We see that the curves in Fig. 16 start at v = 0 with the same value as the θ = π/2
– 22 –
1.0
0.8
0.8
Ls /Liso (s)
Ls /Liso (T )
1.0
0.6
0.4
0.2
0.0
0.0
0.6
0.4
0.2
0.2
0.4
0.6
0.8
0.0
0.0
1.0
0.2
0.4
v
0.6
0.8
1.0
v
1.0
1.0
0.8
0.8
Ls /Liso (s)
Ls /Liso (T )
Figure 10. Screening length for a plasma wind along the z-direction and a dipole oriented
along the z-direction, for four different values of the anisotropy (from top to bottom) a/T =
12.2, 42.6, 86, 744. The corresponding values in units of the entropy density are (in the same
order) aNc2/3 /s1/3 = 6.2, 19, 35, 242. The screening length is plotted in the appropriate units to
facilitate comparison with the isotropic result for a plasma at the same temperature (left), or at
the same entropy density (right). The isotropic result is plotted in Fig. 9, and its ultra-relativistic
behavior is given in eq. (5.60). At v = 0 the curves agree with the θ = 0 values in Fig. 7. As v → 1
they vanish as (1 − v 2 )1/4 , in agreement with (5.58)(top line) and (5.60).
0.6
0.4
0.2
0.0
0
0.6
0.4
0.2
5
10
15
20
25
30
0.0
0
5
10
15
20
25
30
aNc2/3 /s1/3
a/T
Figure 11. Screening length for a plasma wind along the z-direction and a dipole oriented along
the z-direction, at five different velocities (from top to bottom) v = 0.25, 0.5, 0.7, 0.9, 0.9995. The
screening length is plotted in the appropriate units to facilitate comparison with the isotropic result
for a plasma at the same temperature (left), or at the same entropy density (right). The isotropic
result is plotted in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60).
static result shown in Fig. 7, and that they approach a finite, non-zero value as v → 1, in
agreement with (5.58)(bottom line) and (5.60). The qualitative behavior in as in the case
of motion and orientation along x.
Wind along x and dipole along z. The numerical results are shown in Figs. 18 and
19. We see that the curves in Fig. 18 start at v = 0 with the same value as the θ = 0
static result shown in Fig. 7, and that they approach a finite, non-zero value as v → 1, in
agreement with (5.58)(bottom line) and (5.60). The screening length decreases with the
– 23 –
1.0
1.5
Lani /Liso (s)
Lani /Liso (T )
0.8
0.6
0.4
1.0
0.5
0.2
0.0
0.0
0.2
0.4
0.6
0.8
1.0
0.0
0.0
0.2
0.4
v
0.6
0.8
1.0
v
Figure 12. Screening length for a plasma wind along the z-direction and a dipole oriented along
the x-direction, for four different values of the anisotropy a/T = 12.2 (red, solid), 42.6 (maroon,
coarsely dashed), 86 (violet, dashed), 744 (orange, dot-dashed). The corresponding values in units
of the entropy density are (in the same order) aNc2/3 /s1/3 = 6.2, 19, 35, 242. The screening length
is plotted in the appropriate units to facilitate comparison with the isotropic result for a plasma at
the same temperature (left), or at the same entropy density (right). The isotropic result is plotted
in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60). At v = 0 the curves agree with
the θ = π/2 values in Fig. 7. As v → 1 they vanish as (1 − v 2 )1/4 , in agreement with (5.58)(top
line) and (5.60).
0.8
Lani /Liso (s)
Lani /Liso (T )
1.0
0.6
0.4
0.2
0.0
0
5
10
15
20
25
30
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0
5
10
15
20
25
30
a/s1/3
a/T
Figure 13. Screening length for a plasma wind along the z-direction and a dipole oriented along
the x-direction, at five different velocities (from top to bottom) v = 0.25, 0.5, 0.7, 0.9, 0.9995. The
screening length is plotted in the appropriate units to facilitate comparison with the isotropic result
for a plasma at the same temperature (left), or at the same entropy density (right). The isotropic
result is plotted in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60).
anisotropy for any velocity provided the temperature is kept fixed. The same is true at
large anisotropies if the entropy density is kept fixed.
6
Dissociation temperature and dissociation anisotropy
In previous sections we have focused on computing the screening length in an anisotropic
plasma, Ls (T, a), and on comparing it to its isotropic counterpart Liso = Ls (T, 0). The
– 24 –
1.6
0.9
1.4
Ls /Liso (s)
Ls /Liso (T )
1.0
0.8
0.7
0.6
0.0
1.2
1.0
0.8
0.6
0.2
0.4
0.6
0.8
1.0
0.0
0.2
0.4
v
0.6
0.8
1.0
v
1.00
1.20
0.95
1.15
Ls /Liso (s)
Ls /Liso (T )
Figure 14. Screening length for a plasma wind along the x-direction and a dipole oriented
along the x-direction, for four different values of the anisotropy (from top to bottom) a/T =
12.2, 42.6, 86, 744. The corresponding values in units of the entropy density are (in the same
order) aNc2/3 /s1/3 = 6.2, 19, 35, 242. The screening length is plotted in the appropriate units to
facilitate comparison with the isotropic result for a plasma at the same temperature (left), or at
the same entropy density (right). The isotropic result is plotted in Fig. 9, and its ultra-relativistic
behavior is given in eq. (5.60). At v = 0 the curves agree with the θ = π/2 values in Fig. 7. As
v → 1 they approach a finite, non-zero value, in agreement with (5.58)(bottom line) and (5.60).
0.90
0.85
0.80
0.75
0
1.10
1.05
1.00
5
10
15
20
25
30
0.95
0
5
10
15
20
25
30
aNc2/3 /s1/3
a/T
Figure 15. Screening length for a plasma wind along the x-direction and a dipole oriented along
the x-direction, at five different velocities v =0.25 (yellow, dot-dashed), 0.5 (green, short dashed),
0.7 (brown, medium dashed), 0.9 (cyan, long dashed), 0.9995 (blue, solid). The screening length is
plotted in the appropriate units to facilitate comparison with the isotropic result for a plasma at
the same temperature (left), or at the same entropy density (right). The isotropic result is plotted
in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60).
screening length characterizes the dissociation of a quark-antiquark pair for fixed T and
a: A pair separated a distance ℓ < Ls forms a bound state, but if ℓ is increased above
Ls then the bound state dissociates. Similarly, one may define a dissociation temperature
Tdiss (a, ℓ) that characterizes the dissociation of a quark-antiquark pair of fixed size ℓ in a
plasma with a given degree of anisotropy a: for T < Tdiss the pair forms a bound state,
but if T is increased above Tdiss then the bound state dissociates. Analogously, one may
define a dissociation anisotropy adiss (T, ℓ) such that a bound state forms for a < adiss but
not for a > adiss . It is useful to think of the three-dimensional space parametrized by
– 25 –
1.6
0.9
1.4
Lani /Liso (s)
Lani /Liso (T )
1.0
0.8
0.7
0.6
0.0
0.2
0.4
0.6
0.8
1.2
1.0
0.8
0.6
0.0
1.0
0.2
0.4
v
0.6
0.8
1.0
v
Figure 16. Screening length for a plasma wind along the x-direction and a dipole oriented along
the y-direction, for four different values of the anisotropy a/T = 12.2 (red, solid), 42.6 (maroon,
coarsely dashed), 86 (violet, dashed), 744 (orange, dot-dashed). The corresponding values in units
of the entropy density are (in the same order) aNc2/3 /s1/3 = 6.2, 19, 35, 242. The screening length
is plotted in the appropriate units to facilitate comparison with the isotropic result for a plasma at
the same temperature (left), or at the same entropy density (right). The isotropic result is plotted
in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60). At v = 0 the curves agree with
the θ = π/2 values in Fig. 7. As v → 1 they approach a finite, non-zero value, in agreement with
(5.58)(bottom line) and (5.60).
Lani /Liso (s)
Lani /Liso (T )
1.00
0.95
0.90
0.85
0.80
0.75
0
5
10
15
20
25
30
1.20
1.15
1.10
1.05
1.00
0.95
0
5
10
15
20
25
30
aNc2/3 /s1/3
a/T
Figure 17. Screening length for a plasma wind along the x-direction and a dipole oriented along
the y-direction, at five different velocities v =0.25 (yellow, dot-dashed), 0.5 (green, short dashed),
0.7 (brown, medium dashed), 0.9 (cyan, long dashed), 0.9995 (blue, solid). The screening length is
plotted in the appropriate units to facilitate comparison with the isotropic result for a plasma at
the same temperature (left), or at the same entropy density (right). The isotropic result is plotted
in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60).
(T, a, ℓ) as divided in two disconnected regions by a two-dimensional surface: in one region
quark-antiquark pairs bind together, while in the other one they do not. The functions
Ls (T, a), Tdiss (a, ℓ) and adiss (T, ℓ) are then simply different parametrizations of the dividing
surface. It is therefore clear that if a triplet (T, a, ℓ) lies on the dividing surface then
T Ls (a, T ) = Tdiss (a, ℓ)ℓ ,
aLs (T, a) = adiss (T, ℓ)ℓ ,
etc.
(6.1)
In this section we will focus on the qualitative form of Tdiss and adiss . As we will
– 26 –
1.0
0.8
0.8
Lani /Liso (s)
Lani /Liso (T )
1.0
0.6
0.4
0.2
0.0
0.0
0.6
0.4
0.2
0.2
0.4
0.6
0.8
0.0
0.0
1.0
0.2
0.4
v
0.6
0.8
1.0
v
Figure 18. Screening length for a plasma wind along the x-direction and a dipole oriented
along the z-direction, for four different values of the anisotropy (from top to bottom) a/T =
12.2, 42.6, 86, 744. The corresponding values in units of the entropy density are (in the same
order) aNc2/3 /s1/3 = 6.2, 19, 35, 242. The screening length is plotted in the appropriate units to
facilitate comparison with the isotropic result for a plasma at the same temperature (left), or at
the same entropy density (right). The isotropic result is plotted in Fig. 9, and its ultra-relativistic
behavior is given in eq. (5.60). At v = 0 the curves agree with the θ = 0 values in Fig. 7. As v → 1
they approach a finite, non-zero value, in agreement with (5.58)(bottom line) and (5.60).
1.0
0.9
Lani /Liso (s)
Lani /Liso (T )
1.0
0.8
0.7
0.6
0
0.9
0.8
0.7
0.6
5
10
15
20
25
30
0
5
10
15
20
25
30
a/s1/3
a/T
Figure 19. Screening length for a plasma wind along the x-direction and a dipole oriented along
the z-direction, at five different velocities (from bottom to top) v = 0.25, 0.5, 0.7, 0.9, 0.9995. The
screening length is plotted in the appropriate units to facilitate comparison with the isotropic result
for a plasma at the same temperature (left), or at the same entropy density (right). The isotropic
result is plotted in Fig. 9, and its ultra-relativistic behavior is given in eq. (5.60).
see, most of the analysis follows from the asymptotic behavior of the screening length for
a ≫ T . This means that, at the qualitative level, most of the results that we will obtain
would also apply if we were to replace the temperature by the entropy density as one of our
variables. The reason is that, by virtue of (2.5), the limit a ≫ T corresponds to the limit
a ≫ s1/3 and vice versa. In addition, we will see that for generic dipole’s orientations and
velocities, the large-anisotropy limit is entirely controlled by the near-boundary behavior
of the metric at O(u2 ), which depends solely on a and is therefore completely insensitive
to the values of the temperature or of the entropy density.
– 27 –
2.0
vproper /vz
1.8
1.6
1.4
1.2
1.0
0.0
0.2
0.4
0.6
0.8
1.0
u/uH
Figure 20. Proper velocity in the z-direction at a position u away from the boundary, as defined
in (6.2), for different values of a/T . From right to left, a/T = 1.38, 33, 86, 249.
The key point in the large-a analysis is the requirement that no point on the string
can move faster than the local speed of light in the bulk. Consider a meson moving with a
velocity v that has a non-zero component vz along the z-direction. Then we see from (2.2)
that the proper velocity along this direction of a point on the string sitting at a value u of
the radial coordinate is
s
s
gzz (u)
H(u)
= vz
.
(6.2)
vproper (u) = vz −
gtt (u)
F(u)B(u)
The function H(u) increases monotonically from the boundary to the horizon, and is does
so more steeply as a/T increases, as illustrated in Fig. 1. The combination F(u)B(u) has
the opposite behavior, as expected from the fact that gravity is attractive: it decreases
monotonically from the boundary to the horizon. In the isotropic case H = 1 and FB
decreases more steeply as T increases. This is thus the first hint that increasing the
anisotropy has an effect similar to increasing the temperature: both make vproper (u) a more
steeply increasing function away from the boundary. We have illustrated the effect of the
anisotropy in Fig. 20, where we see that vproper /vz becomes a steeper function of u as a/T
increases.
It follows that, for fixed vz 6= 0, there is a maximum value of umax beyond which
vproper becomes superluminal, so no string solution can penetrate to u > umax . As we
will corroborate numerically, this upper bound on umax translates into an upper bound
on Ls . Moreover, umax decreases as a/T increases. This means that for sufficiently large
anisotropies we can use the near-boundary expansions (2.6) in order to determine Ls , in
analogy to what we did in the ultra-relativistic limit. As in that case, for vz 6= 0 the analysis
is controlled by the O(u2 ) terms in (2.6). The key point is that these terms depend on a
but not on T , so by dimensional analysis it follows that umax ∼ a−1 and Ls ∼ a−1 in the
limit a/T ≫ 1. This limit can be understood as a → ∞ at fixed T , or as T → 0 at fixed
a. We thus conclude that, even at T = 0, a generic meson will dissociate for a sufficiently
large anisotropy adiss .
– 28 –
Mesons at rest and mesons whose velocity is exactly aligned with the transverse plane
constitute an exception to the argument above, since in this case vz = 0 and their physics
is mostly insensitive to the function H(u) which characterizes the anisotropic direction.
Therefore in this case we expect that umax and Ls will remain finite as we send a → ∞ at
fixed T , and hence that dimensional analysis will imply Ls ∼ T −1 .
In summary, the heuristic argument above suggests that in the limit a/T ≫ 1 we
should have
Ls (T, a) ∼
−1
const. × T
const. × a−1
if the meson is static or in motion within the transverse plane,
(6.3)
otherwise.
The constants may depend on all the dimensionless parameters such as the velocity and the
dipole’s orientation. We will refer to the behavior in the second line as ‘generic’ and to that
in the first line as ‘non-generic’, since the latter only applies if the velocity is exactly zero or
if the motion is exactly aligned with the transverse plane. The generic behavior is of course
consistent with the analysis of Sec. 5.2.1. Indeed, we saw in that section that for motion
outside the transverse plane the ultra-relativistic behavior of Ls is entirely controlled by
the O(u2 ) terms in the metric, which depend on a but not on T .
Fig. 21 shows our numerical results for umax , in units of T −1 and a−1 , as a function
of a/T , for the five physically distinct cases discussed in Sec. 5.4. From the continuous,
magenta curves in the first two rows we see that umax goes to zero at large a/T in the cases
of motion along z, irrespectively of the dipole’s orientation. In contrast, we see that umax
does not go zero for a static meson (dashed, blue curves) or for a meson moving along the
x-direction (continuous, magenta curves in the last three rows).
Recalling that the isotropic screening length is of the form Liso ∝ 1/T , we see that
the quantity plotted on the vertical axes in Figs. 6, 11, 13, 15, 17 and 19 is precisely
proportional to T Ls (T, a). However, the asymptotic behavior (6.3) is not apparent in
these plots because in most cases the horizontal axes do not extend to high enough values
of a/T . For this reason we have illustrated the two possible asymptotic behaviors of Ls in
Fig. 22, where we have extended the horizontal axes to larger values of a/T . We see from
the continuous, magenta curves in the first two rows that Ls ∼ 1/a for motion along the
z-direction. For motion within the transverse plane we see from the same curves in the last
three rows that Ls ∼ 1/T . This approximate scaling relation seems to hold quite precisely
for a dipole oriented within the transverse plane (3rd and 4th rows), whereas for a dipole
oriented in the z-direction the product T Ls seems to retain a slight (perhaps logarithmic)
dependence on a/T at large a/T . We can draw similar conclusions from the dashed, blue
curves in the figure, which correspond to static mesons. We see that for mesons oriented
within the transverse plane (2nd, 3rd and 4th rows) the relation T Ls ∼ constant holds
quite precisely, whereas for mesons oriented in the z-direction (1st and 5th rows) there
seems to be some slight residual dependence on a/T at large a/T .
– 29 –
0.5
0.3
0.2
30
20
10
0.1
0.0
0
100
200
a/T
0.5
0
0
400
14
12
10
8
6
4
2
0
0
0.3
0.2
0.1
100
200
a/T
0.5
400
0.3
0.2
100
200
a/T
100
200
a/T
300
400
x−x
60
40
0
0
400
0.3
0.2
100
200
a/T
100
300
400
x−y
80
a umax
T umax
300
x−y
0.4
60
40
20
0.1
100
200
a/T
0.5
300
0
0
400
0.2
200
a/T
0.1
300
400
x−z
80
a umax
0.3
100
100
x−z
0.4
T umax
400
20
0.5
0.0
0
300
z−x
80
0.1
0.0
0
200
a/T
100
a umax
T umax
300
x−x
0.4
0.0
0
100
a umax
T umax
300
z−x
0.4
0.0
0
z−z
40
a umax
0.4
T umax
50
z−z
60
40
20
100
200
a/T
300
0
0
400
100
200
a/T
300
400
Figure 21. Value of umax in units of 1/T (left) or 1/a (right), as a function of the ratio a/T , for
a dipole at rest (dashed, blue curve) and for a dipole moving with v = 0.45 (continuous, magenta
curve). The first letter on the top right corner of each plot indicates the direction of motion, and
the second one indicates the orientation of the dipole.
– 30 –
0.25
0.20
0.15
0.10
0.05
0.00
0
0.25
0.20
0.15
0.10
0.05
0.00
0
a Ls
30
20
10
100
200
a/T
300
0
0
400
100
200
a/T
50
z−x
a Ls
300
400
z−x
40
30
20
10
100
200
a/T
300
0
0
400
100
200
a/T
50
x−x
300
400
x−x
40
a Ls
0.25
0.20
0.15
0.10
0.05
0.00
0
z−z
40
30
20
10
100
200
a/T
300
0
0
400
100
200
a/T
50
x−y
300
400
x−y
40
a Ls
0.25
0.20
0.15
0.10
0.05
0.00
0
50
z−z
30
20
10
100
200
a/T
300
0
0
400
100
200
a/T
50
x−z
300
400
x−z
40
a Ls
T Ls
T Ls
T Ls
T Ls
T Ls
0.25
0.20
0.15
0.10
0.05
0.00
0
30
20
10
100
200
a/T
300
400
0
0
100
200
a/T
300
400
Figure 22. Screening length in units of 1/T (left) or 1/a (right), as a function of the ratio a/T , for
a dipole at rest (dashed, blue curve) and for a dipole moving with v = 0.45 (continuous, magenta
curve). The first letter on the top right corner of each plot indicates the direction of motion, and
the second one indicates the orientation of the dipole.
– 31 –
x
0.15
0.10
0.05
0.00
0
ℓ adiss
ℓ Tdiss
0.25
0.20
z
z
2
4
x
6
8
10
12
12
z
x
10 x
8
6 z
4
2
0
0.00 0.05 0.10 0.15 0.20 0.25
aℓ
Tℓ
Figure 23. Dissociation temperature (left) Tdiss (a, ℓ) = ℓ−1 f (aℓ) and dissociation anisotropy
(right) adiss (T, ℓ) = ℓ−1 g(T ℓ) for a dipole at rest (dashed curves) and for a dipole moving along the
z-direction with v = 0.45 (continuous curves). The orientation of the dipole is indicated by a letter
next to each curve.
1.0
vlim
0.8
0.6
0.4
0.2
0.0
0
5
10
15
20
aℓ
Figure 24. Limiting velocity, for fixed anisotropy and T = 0, beyond which a meson oriented along
the x-direction and moving along the z-direction will dissociate.
Combining the two plots on the left and the right columns of Fig. 22 we can eliminate
a/T and obtain T Ls as a function of aLs and vice versa. Recalling (6.1) we see that we can
interpret the result in the first case as Tdiss (a, ℓ) = ℓ−1 f (aℓ), whereas in the second case we
get adiss (T, ℓ) = ℓ−1 g(T ℓ). The functions f and g are the curves shown in Fig. 23(left) and
Fig. 23(right), respectively. The right plot is of course the mirror image along a 45 degree
line of the left plot. We see in Fig. 23(left) that the dissociation temperature decreases
monotonically with increasing anisotropy and vanishes at aℓ ≃ 9.75 (for the chosen velocity
and orientation). On the right plot this corresponds to the dissociation anisotropy at zero
temperature. As anticipated above, even at zero temperature, a generic meson of size ℓ will
dissociate if the anisotropy is increased above adiss (T = 0, ℓ) ∝ 1/ℓ. The proportionality
constant in this relation is a decreasing function of the meson velocity in the plasma. This
implies that for a fixed anisotropy there is a limiting velocity vlim above which a meson will
dissociate, even at zero temperature. The form of vlim (aℓ) for T = 0 is plotted in Fig. 24.
The existence of a limiting velocity for quarkonium mesons is well known in a strongly
coupled isotropic plasma [35, 36], in which case the dissociation at v = vlim is caused by
– 32 –
Tdiss (v)/Tdiss (0)
Tdiss (v)/Tdiss (0)
1.0
1.0
0.8
0.6
0.4
0.2
0.0
0.0
0.2
0.4
0.6
0.8
0.8
0.6
0.4
0.2
0.0
0.0
1.0
v
0.2
0.4
0.6
0.8
1.0
v
Figure 25. Dissociation temperature for a meson moving along the x-direction and oriented along
the z-direction (left) or along the x-direction (right). Each curve corresponds to a fixed value of
the product aℓ = 0 (blue curve), 1.4 (green curve), 25 (red curve).
the temperature. What we see here is that in our anisotropic plasma this behavior persists
as T → 0 for generic motion. In this limit it is the anisotropy that is responsible for the
dissociation. In the case of ultra-relativistic motion the relation between adiss or Tdiss and
vlim can be obtained by combining the scalings (5.58) and (6.3). For generic motion these
relations yield
1
2 1/2
) ,
adiss (T, ℓ) ∼ (1 − vlim
ℓ
[a ≫ T , vlim . 1]
(6.4)
whereas for motion within the transverse plane we obtain
1
2 1/4
Tdiss (a, ℓ) ∼ (1 − vlim
) .
ℓ
[a ≫ T , vlim . 1]
(6.5)
The scaling (6.5) agrees with the isotropic result [31, 32] and illustrates the fact that,
for motion within the transverse plane, the limiting velocity in our anisotropic plasma
approaches unity as T → 0. This behavior is the same for a meson at rest, as illustrated in
Fig. 23, where we see that a sufficiently small meson will remain bound in the plasma for
any value of the anisotropy provided the plasma is cold enough. In fact, the form of the
dissociation temperature for all anisotropies and all velocities within the transverse plane
is qualitatively analogous to that of the isotropic case, as shown in Fig. 25. The fact that
the curves in this figure approximately overlap one another signals that the dependence of
the dissociation temperature on v and aℓ can be approximately factorized over the entire
range 0 ≤ v ≤ 1.
In contrast, for generic motion we saw above that the limiting velocity is subluminal
even at T = 0, vlim (T = 0, aℓ) < 1. Increasing the temperature simply decreases the value
of the limiting velocity, vlim (T ℓ, aℓ) < vlim (T = 0, aℓ). Turning these statements around we
see that, at a fixed anisotropy, the dissociation temperature is a decreasing function of
the velocity that vanishes at v = vlim (T = 0, aℓ). This is illustrated in Fig. 26, where we
see that vlim (T = 0, aℓ) decreases as the anisotropy increases, in agreement with Fig. 24.
In order to facilitate comparison with the isotropic results of [31–33], in Fig. 26 we have
– 33 –
Tdiss (v)/Tdiss (0)
Tdiss (v)/Tdiss (0)
1.0
1.0
0.8
0.6
0.4
0.2
0.0
0.0
0.2
0.4
0.6
0.8
0.8
0.6
0.4
0.2
0.0
0.0
1.0
v
0.2
0.4
0.6
0.8
1.0
v
Figure 26. Dissociation temperature for a meson moving along the z-direction and oriented along
the x-direction (left) or along the z-direction (right). Each curve corresponds to a fixed value of
the product aℓ. From right to left, aℓ = 0, 1, 5.4, 25.
chosen to normalize the dissociation temperature by its value at v = 0 instead of by the
dipole’s size ℓ. Our numerical results suggest that as v approaches vlim the dissociation
temperature may vanish as
ε
Tdiss (v, aℓ)
2
∼ vlim
− v2 .
(6.6)
Tdiss (0, aℓ)
In this equation vlim = vlim (T = 0, aℓ) and ε = ε(aℓ) > 0 is an anisotropy-dependent
exponent. Unfortunately, the limit v → vlim is difficult to analyze numerically, so our
results are not precise enough to allow us to establish (6.6) unambiguously. To emphasize
this point, in Fig. 26 we have plotted as discontinuous the part of the curves between the
last two data points. The last point lies on the horizontal axis at (v, T ) = (vlim , 0), and the
penultimate point lies at a certain height at (v . vlim , T > 0). Since this last bit of the
curves is an interpolation between these data points, it is difficult to establish whether the
slopes of the curves diverge as they meet the horizontal axis, as would be implied by the
scaling (6.6). Presumably, this scaling could be verified or falsified analytically by including
the first correction in T /a to the scaling in the second line of (6.3).
7
Discussion
We have considered an anisotropic N = 4 SYM plasma in which the x, y directions are
rotationally symmetric, but the z-direction is not. In the context of heavy ion collisions the
latter would correspond to the beam direction, and the former to the transverse plane. The
screening length of a quarkonium meson in motion in the plasma depends on the relative
orientation between these directions, on the one hand, and the direction of motion of the
meson and its orientation, on the other. This dependence can be parametrized by three
angles (θv , θ, ϕ), as shown in Fig. 8. We have determined the screening length for the most
general geometric parameters and for any anisotropy. Our results are valid in the strongcoupling, large-Nc limit, since we have obtained them by means of the gravity dual [14, 15]
of the anisotropic N = 4 plasma. The anisotropy is induced by a position-dependent theta
– 34 –
term in the gauge theory, or equivalently by a position-dependent axion on the gravity side.
One may therefore wonder how sensitive the conclusions may be to the specific source of
the anisotropy. In this respect it is useful to note that the gravity calculation involves
only the coupling of the string to the background metric. This means that any anisotropy
that gives rise to a qualitatively similar metric (and no Neveu-Schwarz B-field) will yield
qualitatively similar results for the screening length, irrespectively of the form of the rest
of supergravity fields.
An example of a rather robust conclusion is the ultra-relativistic behavior2 of the
screening length (5.58), which for motion not exactly aligned with the transverse plane is
Ls ∼ (1 − v 2 )1/2 . The 1/2 exponent contrasts with the 1/4 isotropic result [31, 32], and
follows from the fact that the near-boundary fall-off of the metric (2.2) takes the schematic
form
L2
(4)
(2)
gµν = 2 ηµν + u2 gµν
+ ··· .
(7.1)
+ u4 gµν
u
As v grows closer and closer to 1 the point of maximum penetration of the string into the
bulk, umax , moves closer and closer to the AdS boundary at u = 0. As a consequence, the
physics in this limit is solely controlled by the near-boundary behavior of the metric. For
generic motion the behavior is in fact governed by the O(u2 ) terms alone, and a simple
scaling argument then leads to the 1/2 exponent above. In the isotropic case the O(u2 )
terms are absent and the same scaling argument leads to the 1/4 exponent.
In fact, a similar reasoning allowed us to determine the large-anisotropy limit. Since
the metric component gzz ∝ H(u) grows as one moves from the boundary to the horizon,
a subluminal velocity of the meson at the boundary would eventually translate into a
superluminal proper velocity (6.2) at a sufficiently large value of u.3 This sets an upper
limit on the maximum penetration length umax of the string into the bulk and hence on Ls .
Moreover, gzz becomes steeper as a/T increases, so in the limit a/T ≫ 1 the point umax
approaches the AdS boundary (unless the motion is aligned with the transverse plane), just
as in the ultra-relativistic limit. In this limit the physics is again controlled by the O(u2 )
terms in the metric, which depend on a but not on T . Therefore dimensional analysis
implies that Ls = const. × a−1 , were the proportionality ‘constant’ is a decreasing function
of the velocity. This led us to one of our main conclusions: even in the limit T → 0, a
generic meson of size ℓ will dissociate at some high enough anisotropy adiss ∼ ℓ−1 . Similarly,
for fixed a and T , even if T = 0, a generic meson will dissociate if its velocity exceeds a
limiting velocity vlim (a, T ) < 1, as shown in Fig. 24 for T = 0. As explained in Sec. 6, the
conclusions in this paragraph would remain unchanged if we worked at constant entropy
density instead of at constant temperature, since in the limit a ≫ s1/3 the physics would
again be controlled only by the O(u2 ) terms in the metric.
The above discussion makes it clear that, at the qualitative level, much of the physics
(2)
depends only on a few features of the solution: The presence of the gµν term in the near2
We recall that we first send the quark mass to infinity and then v → 1 (see Sec. 1).
Note that the overall conformal factor 1/u2 in (2.2) plays no role in this argument, since it cancels out
in the ratio (6.2).
3
– 35 –
log(s2D /s2D
iso )
0.08
0.06
0.04
0.02
0.00
-4
-2
0
2
4
6
8
log(a/T )
Figure 27. Log-log plot of the entropy density per unit 2-area in the xy-directions on a constant-z
π
2 2
slice as a function of a/T , normalized to the isotropic result s2D
iso = 2 Nc T .
boundary expansion of the metric, the fact that the metric (7.1) be non-boost-invariant
(2)
at order u2 (i.e. that gµν not be proportional to ηµν ), and the fact that gzz increases as
a function of both u and a/T .4 The second condition is necessary because otherwise the
physics of a meson in motion would be equivalent to that of a meson at rest, and we
have seen that the latter is very similar to that of a meson in an isotropic plasma. The
third condition ensures that umax moves close to the boundary as a/T increases. Note that
(4)
adding temperature to an otherwise boost-invariant metric will only affect gµν , and thus
(2)
this is not enough to make gµν non-boost-invariant. This conclusion is consistent with the
(2)
fact that gµν is only a function of the external sources which the theory is coupled to.
From the gauge theory viewpoint, some heuristic intuition can be gained by recalling
that the anisotropy is induced by dissolving along the z-direction objects that extend
along the xy-directions [14, 15, 26]. The number density of such objects along the zdirection, dn/dz, is proportional to a. On the gravity side these are D7-branes that wrap
the five-sphere in the metric (2.2), extend along the xy-directions, and are homogeneously
distributed in the z-direction. Increasing a has a large effect on the entropy density per
unit 3-volume in the xyz-directions, in the sense that s/T 3 → ∞ as a/T → ∞, as shown in
Fig. 2. In contrast, the entropy density per unit 2-area in the xy-directions on a constant-z
slice, s2D /T 2 , approaches a constant in the limit a/T → ∞. This is illustrated in Fig. 27,
which is based on our numerical calculations, but it can also be proven analytically following
the argument in Sec. 2.5 of Ref. [26]. In view of these differences, it is perhaps not surprising
that the anisotropy has the largest effect on the physics of mesons moving along the zdirection, and the smallest effect on the physics of mesons moving within the transverse
plane. Mesons at rest are also more sensitive to the anisotropy if they extend along the
z-direction than if they are contained within the transverse plane. Presumably, the correct
intuition behind this physics is that moving against the D7-branes is harder than moving
along them.
We close with a few comments on existing weak-coupling results on the physics of
4
Again, up to possible overall conformal factors.
– 36 –
quarkonium dissociation in the real-world QGP. In the isotropic case the velocity dependence of the heavy quark potential has been studied using perturbative and effective field
theory methods, see e.g. [37–40]. These analyses include modifications of both the real
and imaginary parts of the potential, which are related to screening and to the thermal
width of the states, respectively. They find that meson dissociation at non-zero velocity
results form a complex interplay between the real and the imaginary parts of the potential.
However, the general trend that seems to emerge is that screening effects increase with the
velocity, while the width of the states decreases. The behavior of the real part is thus in
qualitative agreement with the isotropic limit of our results. However, the extraction of a
screening length from these analyses is not immediate due to the fact that the real part of
the potential is not approximately Yukawa-like [39, 40], in contrast with the holographic
result. In any case, an interesting consequence of the dominance of the real part of the
potential is that, at sufficiently high velocities, dissociation is caused by screening rather
than by Landau damping [39, 40]. In the holographic framework, the thermal widths of
our mesons could presumably be computed along the lines of [41].
To the best of our knowledge no results at non-zero velocity exist in the presence of
anisotropies, so in this case we will limit ourselves to the static situation. We emphasize
though that any comparison between these results and ours should be interpreted with
caution, because the sources of anisotropy in the QGP created in a heavy ion collision
and in our system are different. In the QGP the anisotropy is dynamical in the sense
that it is due to the initial distribution of particles in momentum space, which will evolve
in time and eventually become isotropic. In contrast, in our case the anisotropy is due
to an external source that keeps the system in an equilibrium anisotropic state that will
not evolve in time. We hope that, nevertheless, our system might provide a good toy
model for processes whose characteristic time scale is sufficiently shorter than the time
scale controlling the time evolution of the QGP.
A general conclusion of Refs. [42–44] is that, if the comparison between the anisotropic
plasma and its isotropic counterpart is made at equal temperatures, then the screening
length increases with the anisotropy. This effect occurs for dipoles oriented both along and
orthogonally to the anisotropic direction, but it is more pronounced for dipoles along the
anisotropic direction. The dependence on the anisotropy in these weak-coupling results is
the opposite of what we find in our strongly coupled plasma. In our case the screening
length in the anisotropic plasma is smaller than in its isotropic counterpart if both plasmas
are taken to have the same temperature, as shown in Fig. 6(left). We also find that the effect
is more pronounced for dipoles extending along the anisotropic direction, as illustrated in
Fig. 7(left).
Refs. [44, 45] argued that if the comparison between the anisotropic and the isotropic
plasmas is made at equal entropy densities, then the physics of quarkonium dissociation
exhibits little or no sensitivity to the value of the anisotropy. This is again in contrast to
our results since, as shown in Fig. 6(right) and Fig. 7(right), the screening length in this
case is just as sensitive to the anisotropy as in the equal-temperature comparison. The
– 37 –
difference in the equal-entropy case is simply that the screening length may increase or
decrease with the anisotropy depending on the dipole’s orientation.
Acknowledgments
It is a pleasure to thank M. Strickland for helpful discussions. MC is supported by a postdoctoral fellowship from Mexico’s National Council of Science and Technology (CONACyT). We acknowledge financial support from 2009-SGR-168, MEC FPA2010-20807-C0201, MEC FPA2010-20807-C02-02 and CPAN CSD2007-00042 Consolider-Ingenio 2010 (MC,
DF and DM), and from DE-FG02-95ER40896 and CNPq (DT).
References
[1] J. Adams et al. [STAR Collaboration], “Experimental and theoretical challenges in the
search for the quark gluon plasma: The STAR collaboration’s critical assessment of the
evidence from RHIC collisions,” Nucl. Phys. A 757, 102 (2005) [arXiv:nucl-ex/0501009].
[2] K. Adcox et al. [PHENIX Collaboration], “Formation of dense partonic matter in relativistic
nucleus nucleus collisions at RHIC: Experimental evaluation by the PHENIX collaboration,”
Nucl. Phys. A 757, 184 (2005) [arXiv:nucl-ex/0410003].
[3] See the contributions on elliptic flow at the LHC in the proceedings of Quark Matter 2011:
J. Phys. GG 38, number 12 (December 2011).
[4] E. Shuryak, “Why does the quark gluon plasma at RHIC behave as a nearly ideal fluid?,”
Prog. Part. Nucl. Phys. 53, 273 (2004) [arXiv:hep-ph/0312227].
[5] E. V. Shuryak, “What RHIC experiments and theory tell us about properties of quark-gluon
plasma?,” Nucl. Phys. A 750, 64 (2005) [arXiv:hep-ph/0405066].
[6] J. M. Maldacena, “The Large N limit of superconformal field theories and supergravity,”
Adv. Theor. Math. Phys. 2, 231-252 (1998) [hep-th/9711200].
[7] S. S. Gubser, I. R. Klebanov, A. M. Polyakov, “Gauge theory correlators from noncritical
string theory,” Phys. Lett. B428, 105-114 (1998) [hep-th/9802109].
[8] E. Witten, “Anti-de Sitter space and holography,” Adv. Theor. Math. Phys. 2, 253-291
(1998) [hep-th/9802150].
[9] J. Casalderrey-Solana, H. Liu, D. Mateos, K. Rajagopal, U. A. Wiedemann, “Gauge/String
Duality, Hot QCD and Heavy Ion Collisions,” [arXiv:1101.0618 [hep-th]].
[10] P. M. Chesler and L. G. Yaffe, “Boost invariant flow, black hole formation, and
far-from-equilibrium dynamics in N = 4 supersymmetric Yang-Mills theory,” Phys. Rev. D
82 (2010) 026006 [arXiv:0906.4426 [hep-th]].
[11] P. M. Chesler and L. G. Yaffe, “Holography and colliding gravitational shock waves in
asymptotically AdS5 spacetime,” Phys. Rev. Lett. 106 (2011) 021601 [arXiv:1011.3562
[hep-th]].
[12] M. P. Heller, R. A. Janik and P. Witaszczyk, “The characteristics of thermalization of
boost-invariant plasma from holography,” Phys. Rev. Lett. 108, 201602 (2012)
[arXiv:1103.3452 [hep-th]].
– 38 –
[13] M. P. Heller, R. A. Janik and P. Witaszczyk, “A numerical relativity approach to the initial
value problem in asymptotically Anti-de Sitter spacetime for plasma thermalization - an
ADM formulation,” Phys. Rev. D 85, 126002 (2012) [arXiv:1203.0755 [hep-th]].
[14] D. Mateos and D. Trancanelli, “The anisotropic N=4 super Yang-Mills plasma and its
instabilities,” Phys. Rev. Lett. 107, 101601 (2011) [arXiv:1105.3472 [hep-th]].
[15] D. Mateos and D. Trancanelli, “Thermodynamics and Instabilities of a Strongly Coupled
Anisotropic Plasma,” JHEP 1107, 054 (2011) [arXiv:1106.1637 [hep-th]].
[16] A. Rebhan and D. Steineder, “Probing Two Holographic Models of Strongly Coupled
Anisotropic Plasma,” JHEP 1208, 020 (2012) [arXiv:1205.4684 [hep-th]].
[17] R. A. Janik and P. Witaszczyk, “Towards the description of anisotropic plasma at strong
coupling,” JHEP 0809, 026 (2008). [arXiv:0806.2141 [hep-th]].
[18] D. Giataganas, “Probing strongly coupled anisotropic plasma,” JHEP 1207, 031 (2012)
[arXiv:1202.4436 [hep-th]].
[19] A. Rebhan and D. Steineder, “Violation of the Holographic Viscosity Bound in a Strongly
Coupled Anisotropic Plasma,” Phys. Rev. Lett. 108, 021601 (2012) [arXiv:1110.6825
[hep-th]].
[20] K. A. Mamo, “Holographic Wilsonian RG Flow of the Shear Viscosity to Entropy Ratio in
Strongly Coupled Anisotropic Plasma,” arXiv:1205.1797 [hep-th].
[21] M. Chernicoff, D. Fernandez, D. Mateos and D. Trancanelli, “Drag force in a strongly
coupled anisotropic plasma,” arXiv:1202.3696 [hep-th].
[22] M. Chernicoff, D. Fernandez, D. Mateos and D. Trancanelli, “Jet quenching in a strongly
coupled anisotropic plasma,” JHEP 1208, 041 (2012) [arXiv:1203.0561 [hep-th]].
[23] K. B. Fadafan and H. Soltanpanahi, “Energy loss in a strongly coupled anisotropic plasma,”
arXiv:1206.2271 [hep-th].
[24] A. Gynther, A. Rebhan and D. Steineder, “Thermodynamics and phase diagram of
anisotropic Chern-Simons deformed gauge theories,” arXiv:1207.6283 [hep-th].
[25] C. Krishnan, “Baryon Dissociation in a Strongly Coupled Plasma,” JHEP 0812, 019 (2008)
[arXiv:0809.5143 [hep-th]].
[26] T. Azeyanagi, W. Li and T. Takayanagi, “On String Theory Duals of Lifshitz-like Fixed
Points,” JHEP 0906, 084 (2009) [arXiv:0905.0688 [hep-th]].
[27] S. Kachru, X. Liu, M. Mulligan, “Gravity Duals of Lifshitz-like Fixed Points,” Phys. Rev.
D78 (2008) 106005 [arXiv:0808.1725 [hep-th]].
[28] K. Copsey, R. Mann, “Pathologies in Asymptotically Lifshitz Spacetimes,” JHEP 1103
(2011) 039 [arXiv:1011.3502 [hep-th]].
[29] S. -J. Rey, S. Theisen and J. -T. Yee, “Wilson-Polyakov loop at finite temperature in large N
gauge theory and anti-de Sitter supergravity,” Nucl. Phys. B 527, 171 (1998)
[hep-th/9803135].
[30] A. Brandhuber, N. Itzhaki, J. Sonnenschein and S. Yankielowicz, “Wilson loops in the large
N limit at finite temperature,” Phys. Lett. B 434, 36 (1998) [hep-th/9803137].
[31] H. Liu, K. Rajagopal and U. A. Wiedemann, “An AdS/CFT Calculation of Screening in a
Hot Wind,” Phys. Rev. Lett. 98, 182301 (2007) [hep-ph/0607062].
– 39 –
[32] H. Liu, K. Rajagopal, U. A. Wiedemann, “Wilson loops in heavy ion collisions and their
calculation in AdS/CFT,” JHEP 0703, 066 (2007). [hep-ph/0612168].
[33] M. Chernicoff, J. A. Garcia, A. Guijosa, “The Energy of a Moving Quark-Antiquark Pair in
an N=4 SYM Plasma,” JHEP 0609, 068 (2006). [hep-th/0607089].
[34] C. P. Herzog, A. Karch, P. Kovtun, C. Kozcaz and L. G. Yaffe, “Energy loss of a heavy
quark moving through N=4 supersymmetric Yang-Mills plasma,” JHEP 0607, 013 (2006)
[hep-th/0605158].
[35] D. Mateos, R. C. Myers and R. M. Thomson, “Thermodynamics of the brane,” JHEP 0705,
067 (2007) [hep-th/0701132].
[36] Q. J. Ejaz, T. Faulkner, H. Liu, K. Rajagopal and U. A. Wiedemann, “A Limiting velocity
for quarkonium propagation in a strongly coupled plasma via AdS/CFT,” JHEP 0804, 089
(2008) [arXiv:0712.0590 [hep-th]].
[37] T. Song, Y. Park, S. H. Lee and C. -Y. Wong, “The Thermal width of heavy quarkonia
moving in quark gluon plasma,” Phys. Lett. B 659, 621 (2008) [arXiv:0709.0794 [hep-ph]].
[38] F. Dominguez and B. Wu, “On dissociation of heavy mesons in a hot quark-gluon plasma,”
Nucl. Phys. A 818, 246 (2009) [arXiv:0811.1058 [hep-ph]].
[39] M. A. Escobedo, J. Soto and M. Mannarelli, “Non-relativistic bound states in a moving
thermal bath,” Phys. Rev. D 84, 016008 (2011) [arXiv:1105.1249 [hep-ph]].
[40] M. A. Escobedo, “Non-relativistic bound states in a moving thermal bath,”
http://quark.phy.bnl.gov/www/rhic2/talks_qprogram/moving.pdf
[41] T. Faulkner and H. Liu, “Meson widths from string worldsheet instantons,” Phys. Lett. B
673, 161 (2009) [arXiv:0807.0063 [hep-th]].
[42] A. Dumitru, Y. Guo and M. Strickland, “The Heavy-quark potential in an anisotropic
(viscous) plasma,” Phys. Lett. B 662, 37 (2008) [arXiv:0711.4722 [hep-ph]].
[43] A. Dumitru, Y. Guo, A. Mocsy and M. Strickland, “Quarkonium states in an anisotropic
QCD plasma,” Phys. Rev. D 79, 054019 (2009) [arXiv:0901.1998 [hep-ph]].
[44] Y. Burnier, M. Laine and M. Vepsalainen, “Quarkonium dissociation in the presence of a
small momentum space anisotropy,” Phys. Lett. B 678, 86 (2009) [arXiv:0903.3467 [hep-ph]].
[45] O. Philipsen and M. Tassler, “On Quarkonium in an anisotropic quark gluon plasma,”
arXiv:0908.1746 [hep-ph].
– 40 –