pacs numbers: 25.75.-q, 24.10.nz, 47.75+f - arxiv.org · pacs numbers: 25.75.-q, 24.10.nz, 47.75+f...

16
Exact solutions and attractors of higher-order viscous fluid dynamics for Bjorken flow Sunil Jaiswal, 1, * Chandrodoy Chattopadhyay, 2, Amaresh Jaiswal, 3, Subrata Pal, 1, § and Ulrich Heinz 2, 4, 1 Department of Nuclear and Atomic Physics, Tata Institute of Fundamental Research, Mumbai 400005, India 2 Department of Physics, The Ohio State University, Columbus, Ohio 43210-1117, USA 3 School of Physical Sciences, National Institute of Science Education and Research, HBNI, Jatni 752050, Odisha, India 4 Institut f¨ ur Theoretische Physik, J.W.Goethe-Universit¨at, Max-von-Laue-Straße1, D-60438 Frankfurt am Main, Germany (Dated: September 4, 2019) We consider causal higher order theories of relativistic viscous hydrodynamics in the limit of one- dimensional boost-invariant expansion and study the associated dynamical attractor. We obtain evolution equations for the inverse Reynolds number as a function of Knudsen number. The solutions of these equations exhibit attractor behavior which we analyze in terms of Lyapunov exponents using several different techniques. We compare the attractors of the second-order M¨ uller-Israel-Stewart (MIS), transient Denicol-Niemi-Molnar-Rischke (DNMR), and third-order theories with the exact solution of the Boltzmann equation in the relaxation-time approximation. It is shown that for Bjorken flow the third-order theory provides a better approximation to the exact kinetic theory attractor than both MIS and DNMR theories. For three different choices of the time dependence of the shear relaxation rate we find analytical solutions for the energy density and shear stress and use these to study the attractors analytically. By studying these analytical solutions at both small and large Knudsen numbers we characterize and uniquely determine the attractors and Lyapunov exponents. While for small Knudsen numbers the approach to the attractor is exponential, strong power-law decay of deviations from the attractor and rapid loss of initial state memory is found even for large Knudsen numbers. Implications for the applicability of hydrodynamics in far-off- equilibrium situations are discussed. PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an effective macroscopic theory de- scribing the large scale and slowly varying dynamical modes of multiparticle systems. The conventional for- mulation of hydrodynamic equations proceeds by assum- ing a separation of macroscopic and microscopic length and time scales such that gradients of local-equilibrium quantities, normalized to appropriate powers of the sys- tem’s energy density, are small and one can perform an order-by-order gradient expansion around local ther- modynamic equilibrium [1]. Therefore the remarkable success of relativistic hydrodynamics in describing the quark-gluon plasma (QGP) formed in ultra-relativistic heavy-ion collisions initially led to the belief that these collisions create a nearly thermalized medium close to local thermal equilibrium [2] (see also the reviews [1, 35]). On the other hand, with the advent of numerical dissipative relativistic fluid dynamics [612] it became clear that the dynamical evolution of heavy-ion collisions is affected by persistent large dissipative corrections. In spite of this numerical evidence for large deviations from local thermal equilibrium, second-order dissipative rela- tivistic fluid dynamics (especially when coupled with a * [email protected] [email protected] [email protected] § [email protected] [email protected] hadronic cascade to describe the final dilute decoupling stage [13, 14]) met with impressive phenomenological suc- cess and predictive power in the description of heavy-ion collision experiments [7, 11, 12, 1522], and was even successfully applied to explain flow data in small colli- sion systems formed in proton-proton (p-p) and proton- lead (p-Pb) collisions [23, 24]. This “unreasonable effec- tiveness” of hydrodynamics as a dynamical description of high-energy hadronic collisions in situations that are even very far away from local thermal equilibrium has generated much recent interest in the very foundations of fluid dynamics [2552], culminating in the formula- tion of a new “far-from-local-equilibrium fluid dynam- ics” paradigm [5, 37]. The present work is a contribution to this ongoing discussion, adding new analytic results for the heavily studied simple case of (0+1)-dimensional Bjorken expansion of a transversally homogeneous sys- tem with longitudinal boost-invariance [53]. The simplest relativistic dissipative theory, relativis- tic Navier-Stokes theory [54, 55], imposes instantaneous constitutive relations between the dissipative flows and their generating forces, expressed through first-order gra- dients of equilibrium quantities. This approach was found to be plagued by acausality and intrinsic instabil- ity [56, 57]. The phenomenological second-order theory developed by M¨ uller, Israel and Stewart (MIS) [5860] cures these problems by introducing a relaxation type equation for the dissipative flows and thus turning them into independent dynamical degrees of freedom of the system whose evolution is controlled by the competition between macroscopic expansion (driving the system away arXiv:1907.07965v3 [nucl-th] 3 Sep 2019

Upload: others

Post on 01-Sep-2019

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

Exact solutions and attractors of higher-order viscous fluid dynamics for Bjorken flow

Sunil Jaiswal,1, ∗ Chandrodoy Chattopadhyay,2, † Amaresh Jaiswal,3, ‡ Subrata Pal,1, § and Ulrich Heinz2, 4, ¶

1Department of Nuclear and Atomic Physics, Tata Institute of Fundamental Research, Mumbai 400005, India2Department of Physics, The Ohio State University, Columbus, Ohio 43210-1117, USA

3School of Physical Sciences, National Institute of Science Education and Research, HBNI, Jatni 752050, Odisha, India4Institut fur Theoretische Physik, J. W. Goethe-Universitat,

Max-von-Laue-Straße 1, D-60438 Frankfurt am Main, Germany(Dated: September 4, 2019)

We consider causal higher order theories of relativistic viscous hydrodynamics in the limit of one-dimensional boost-invariant expansion and study the associated dynamical attractor. We obtainevolution equations for the inverse Reynolds number as a function of Knudsen number. The solutionsof these equations exhibit attractor behavior which we analyze in terms of Lyapunov exponents usingseveral different techniques. We compare the attractors of the second-order Muller-Israel-Stewart(MIS), transient Denicol-Niemi-Molnar-Rischke (DNMR), and third-order theories with the exactsolution of the Boltzmann equation in the relaxation-time approximation. It is shown that forBjorken flow the third-order theory provides a better approximation to the exact kinetic theoryattractor than both MIS and DNMR theories. For three different choices of the time dependenceof the shear relaxation rate we find analytical solutions for the energy density and shear stress anduse these to study the attractors analytically. By studying these analytical solutions at both smalland large Knudsen numbers we characterize and uniquely determine the attractors and Lyapunovexponents. While for small Knudsen numbers the approach to the attractor is exponential, strongpower-law decay of deviations from the attractor and rapid loss of initial state memory is foundeven for large Knudsen numbers. Implications for the applicability of hydrodynamics in far-off-equilibrium situations are discussed.

PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f

I. INTRODUCTION

Hydrodynamics is an effective macroscopic theory de-scribing the large scale and slowly varying dynamicalmodes of multiparticle systems. The conventional for-mulation of hydrodynamic equations proceeds by assum-ing a separation of macroscopic and microscopic lengthand time scales such that gradients of local-equilibriumquantities, normalized to appropriate powers of the sys-tem’s energy density, are small and one can performan order-by-order gradient expansion around local ther-modynamic equilibrium [1]. Therefore the remarkablesuccess of relativistic hydrodynamics in describing thequark-gluon plasma (QGP) formed in ultra-relativisticheavy-ion collisions initially led to the belief that thesecollisions create a nearly thermalized medium close tolocal thermal equilibrium [2] (see also the reviews [1, 3–5]). On the other hand, with the advent of numericaldissipative relativistic fluid dynamics [6–12] it becameclear that the dynamical evolution of heavy-ion collisionsis affected by persistent large dissipative corrections. Inspite of this numerical evidence for large deviations fromlocal thermal equilibrium, second-order dissipative rela-tivistic fluid dynamics (especially when coupled with a

[email protected][email protected][email protected]§ [email protected][email protected]

hadronic cascade to describe the final dilute decouplingstage [13, 14]) met with impressive phenomenological suc-cess and predictive power in the description of heavy-ioncollision experiments [7, 11, 12, 15–22], and was evensuccessfully applied to explain flow data in small colli-sion systems formed in proton-proton (p-p) and proton-lead (p-Pb) collisions [23, 24]. This “unreasonable effec-tiveness” of hydrodynamics as a dynamical descriptionof high-energy hadronic collisions in situations that areeven very far away from local thermal equilibrium hasgenerated much recent interest in the very foundationsof fluid dynamics [25–52], culminating in the formula-tion of a new “far-from-local-equilibrium fluid dynam-ics” paradigm [5, 37]. The present work is a contributionto this ongoing discussion, adding new analytic resultsfor the heavily studied simple case of (0+1)-dimensionalBjorken expansion of a transversally homogeneous sys-tem with longitudinal boost-invariance [53].

The simplest relativistic dissipative theory, relativis-tic Navier-Stokes theory [54, 55], imposes instantaneousconstitutive relations between the dissipative flows andtheir generating forces, expressed through first-order gra-dients of equilibrium quantities. This approach wasfound to be plagued by acausality and intrinsic instabil-ity [56, 57]. The phenomenological second-order theorydeveloped by Muller, Israel and Stewart (MIS) [58–60]cures these problems by introducing a relaxation typeequation for the dissipative flows and thus turning theminto independent dynamical degrees of freedom of thesystem whose evolution is controlled by the competitionbetween macroscopic expansion (driving the system away

arX

iv:1

907.

0796

5v3

[nu

cl-t

h] 3

Sep

201

9

Page 2: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

2

from local equilibrium) and microscopic scattering (driv-ing it back towards local equilibrium). As discussed in [5],even the minimal causal conformal theory given by MISintroduces new modes called non-hydrodynamic modesthat were absent in Navier-Stokes theory. These non-hydrodynamic modes are now known to play an impor-tant role in the approach to the regime of applicability ofhydrodynamics, also known as the “hydrodynamization”process [26, 30, 61–63]. For conformal systems undergo-ing longitudinal Bjorken expansion, it was shown explic-itly that hydrodynamization occurs at microscopic ther-malization time scales which, in strongly and anisotrop-ically expanding systems, are much shorter than boththe time scales for local isotropization and thermalization[37]. Similar conclusions were obtained in recent studies[32, 42] of systems undergoing Gubser flow [64] wherethe fireball expands not only boost-invariantly in longi-tudinal, but also azimuthally symmetrically in transversedirection. These findings lead to the conclusion that thecriterium of proximity to local thermodynamic equilib-rium for hydrodynamics to be valid is (at least in thesecarefully analytically studied simplified situations) toostrict and should be replaced by a different conditionstating that contributions from the non-hydrodynamicmodes can be neglected [31]. In the present study, wewill focus on yet another interesting feature that appearsin a causal theory of relativistic dissipative hydrodynam-ics, “the hydrodynamic attractor” [26, 32, 36, 38, 42, 43],and its role in the hydrodynamization process.

Attractor behavior was first identified by consideringthe hydrodynamic formulation as a gradient series expan-sion. Recently, this gradient series was shown, for severalhighly symmetric flow configurations that are amenableto analytic treatment, to have zero radius of conver-gence. This suggests that hydrodynamic theory cannotbe systematically improved by taking into account higherorder-terms in the gradient series [26]. The gradient ex-pansion generates an asymptotic series which exhibits ini-tial signs of convergence for a few terms before eventuallydiverging [26, 28]. The initial appearance of convergencemay explain the observed remarkable phenomenologicalsuccess of hydrodynamic formulations based on trunca-tions of the gradient expansion at second or third orderbut, due to the ultimate divergence of the series, the the-ory cannot be improved beyond a certain order by keep-ing additional terms. Fortunately, the (diverging) gradi-ent expansion series can be Borel resummed, giving riseto a unique hydrodynamic attractor solution (hydrody-namic mode), which is well defined even for large gradi-ents, and a series of rapidly decaying non-hydrodynamicmodes that describe the approach towards this attractorfrom arbitrary initial conditions [25, 26]. This suggeststhat hydrodynamics displays a novel type of universalityeven far from local equilibrium which is independent ofthe initial state of the system, indicating the existenceof a new, far-from-local-equilibrium hydrodynamic the-ory [37]. Recently, it was discovered [65] that this phe-nomenon extends even beyond hydrodynamics: in kinetic

theory also the evolution of non-hydrodynamic higher-order momentum moments of the distribution functionis controlled by attractors.

The theory of dynamical attractors can be complex,and hence it is instructive to have examples were theattractor is quantitatively understood using analyticalmethods. Perhaps the most useful quantities associatedwith an attractor are Lyapunov exponents which charac-terize the rate of separation of infinitesimally close tra-jectories in the phase-space evolution of dynamical sys-tems [66]. While negative Lyapunov exponents are as-sociated with dissipative systems and indicate the exis-tence of attractors, positive values are usually associatedwith chaotic systems. For a conservative system one ob-tains vanishing Lyapunov exponents. In this article, weemploy the theory of Lyapunov exponents to study theattractors for MIS theory and two other, improved ver-sions of causal relativistic dissipative hydrodynamics (seeRefs. [32, 47] for earlier related work).

Motivated by its application to QGP evolution, severalimprovements over MIS theory were proposed during thelast decade [67–78]. For Bjorken and Gubser flows, someof these theories lead to very good agreement with theexact solution of the Boltzmann equation even in sit-uations where the deviations from thermal equilibriumare large [42, 72, 79, 80]. In this article, we considerthe Denicol-Niemi-Molnar-Rischke (DNMR) theory [69]which is an improved version of second-order MIS the-ory, as well as a third-order hydrodynamic theory derivedform relativistic kinetic theory by going to third orderin the Chapman-Enskog expansion [72]. To understandthe emergent attractor behavior we go beyond previousnumerical studies of these theories by finding analyticalsolutions of the corresponding hydrodynamic evolutionequations [81]. Analytical solutions of higher-order dis-sipative hydrodynamics were found previously for a fewvery special cases [35, 82]. We here expand this portfolioand use the new analytical solutions to study their de-pendence on initial conditions, their attractors, and theirlate-time behavior.

The rest of this article is structured as follows. InSec. II we briefly review causal theories of second-order(MIS and DNMR) and third-order relativistic viscoushydrodynamics. In Sec. III we simplify these theoriesfor Bjorken flow and obtain a generic evolution equa-tion for the inverse Reynolds number as a function ofthe inverse Knudsen number, which can be adapted toall three hydrodynamic theories by adjusting a set oftwo parameters. This ordinary differential equation isdecoupled from the hydrodynamic evolution of the en-ergy density (but feeds back into it), and thus it can besolved independently. In Sec. IV we demonstrate numer-ically that the solutions of these differential equations ex-hibit attractor-like behavior which we analyze in terms ofLyapunov exponents, using several different approaches.We also compare the exact numerical attractors of thesecond-order MIS and DNMR theories, as well as ourthird-order theory, with the exact solution of the rela-

Page 3: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

3

tivistic Boltzmann equation in the relaxation-time ap-proximation (RTA). We show that the third-order theoryprovides a better approximation to the exact kinetic the-ory attractor than both the MIS and DNMR attractors.In Sec. V we provide further clarification of the numericalbehavior found in the preceding sections by working outexplicit analytical solutions for the energy density andshear stress as a function of the inverse Knudsen num-ber, for three different forms of the shear relaxation time.These analytical solutions allow us to also study the at-tractors analytically. In Sec. VI, finally, we explore theuniversal behavior of these solutions at both small andlarge Knudsen numbers and use the results to charac-terize and uniquely determine the hydrodynamic attrac-tor and its associated Lyapunov exponents in a new way.Our results are summarized and some conclusions offeredin Sec. VII.

II. RELATIVISTIC DISSIPATIVEHYDRODYNAMICS

In this section, we will briefly review the equationsfor relativistic dissipative hydrodynamics. We considera conformal system, which corresponds in kinetic theoryto a system of massless particles. The energy-momentumtensor for such a system, in the Landau frame, has theform

Tµν = εuµuν − P∆µν + πµν , (1)

where ε and P are the local energy density and pressure.Conformal symmetry implies an equation of state (EoS)ε = 3P and zero bulk viscous pressure, Π = 0. We alsodefine ∆µν ≡ gµν−uµuν which serves as a projectionoperator to the space orthogonal to uµ (i.e. onto thespatial directions in the local rest frame (LRF)). Theshear stress tensor, πµν , is traceless and orthogonal to uµ.The metric convention used here is gµν = diag(+−−−).

Evolution equations for ε and uµ are obtained fromenergy-momentum conservation, DµT

µν = 0:

ε+ (ε+ P )θ − πµνσµν = 0, (2)

(ε+P ) uα −∇αP + ∆αν ∂µπ

µν = 0. (3)

Here we use the notations Dµ for the covariant deriva-

tive, A ≡ uµDµA for the co-moving time derivative,∇α ≡ ∆µαDµ for space-like derivative, θ ≡ Dµu

µ for theexpansion scalar, and σµν ≡ 1

2 (∇µuν+∇νuµ) − 13θ∆µν

for the velocity shear tensor.To close Eqs. (2),(3) we need additional equations

for the shear stress πµν . The simplest form of πµν isthe Navier-Stokes form, which is first order in velocitygradients, πµνNS = 2ησµν , where η is the shear viscos-ity coefficient. As already mentioned in the Introduc-tion, relativistic Navier-Stokes theory violates causalityand is unstable. The simplest way to restore causal-ity is by introducing a relaxation-type equation for πµν .This prescription, also known as the “Maxwell-Cattaneolaw”, requires that the dissipative forces relax to their

Navier-Stokes values in some finite relaxation time, i.e.,τππ〈µν〉+πµν = 2ησµν ,1 where τπ is the shear relaxation

time. For conformally symmetric systems one more termmust be added in the evolution of shear stress [83, 84]:

τππ〈µν〉 + πµν = 2ησµν − 4

3τππ

µνθ. (4)

This equation is a close variant [83] of the one first de-rived by Muller, Israel and Stewart [58–60], and we willtherefore refer to it as the “MIS” theory. Its derivationwas based on an analysis of the entropy current and thesecond law of thermodynamics, without recourse to a spe-cific theory for the underlying microscopic dynamics.2

A systematic derivation of second-order (“transient”)relativistic fluid dynamics from relativistic kinetic theory,using an expansion of the dissipative flows in momentum-moments of the distribution function, was performed in[69]. For conformal systems and an RTA collision term,the result obtained in the 14-moment approximation dif-fers from Eq. (4) by two additional terms that are ofsecond order in gradients:

π〈µν〉+πµν

τπ= 2βπσ

µν+ 2π〈µγ ων〉γ− 10

7π〈µγ σ

ν〉γ− 4

3πµνθ.

(5)Here βπ ≡ η/τπ = 4P/5, while ωµν ≡ 1

2 (∇µuν−∇νuµ) isthe vorticity tensor. This “DNMR” theory [69] can alsobe derived from a Chapman-Enskog like iterative solutionof the RTA Boltzmann equation [85].

Carrying the Chapman-Enskog expansion to one ad-ditional order, a third-order evolution equation for theshear stress was derived for the same system in [72]:

π〈µν〉 =− πµν

τπ+ 2βπσ

µν + 2π〈µγ ων〉γ − 10

7π〈µγ σ

ν〉γ

− 4

3πµνθ +

25

7βππρ〈µων〉γπργ −

1

3βππ〈µγ π

ν〉γθ

− 38

245βππµνπργσργ −

22

49βππρ〈µπν〉γσργ

− 24

35∇〈µ

(πν〉γ uγτπ

)+

4

35∇〈µ

(τπ∇γπν〉γ

)− 2

7∇γ(τπ∇〈µπν〉γ

)+

12

7∇γ(τπu〈µπν〉γ

)− 1

7∇γ(τπ∇γπ〈µν〉

)+

6

7∇γ(τπu

γπ〈µν〉)

− 2

7τπω

ρ〈µων〉γπργ −2

7τππ

ρ〈µων〉γωργ

− 10

63τππ

µνθ2 +26

21τππ〈µγ ω

ν〉γθ. (6)

1 Angular brackets around pairs of Lorentz indices indicate projec-tion of the tensor onto its traceless and locally spatial part, e.g.,π〈µν〉 = ∆µν

αβ παβ , where ∆µν

αβ = 12

(∆µα∆ν

β+∆µβ∆ν

α)− 13

∆µν∆αβ .2 It should be noted that Eq. (4) does not include all possible

second-order terms [84], and that (unlike the theories discussedbelow) its transport coefficients are not matched to an underlyingkinetic theory.

Page 4: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

4

For conformal systems the complete set of possible thirdorder terms was given in [86]. The specific form (6) im-plements transport coefficients obtained from the RTABoltzmann equation [72]. We will here refer to Eq. (6) asthe “third-order” theory. Let us now simplify all of theseequations for Bjorken flow.

III. BJORKEN FLOW

Milne coordinates xµ = (τ, x, y, ηs) (with τ =√t2−z2

and ηs = tanh−1(z/t)) are the natural choice for de-scribing ultra-relativistic heavy-ion collisions where thecolliding nuclei approach each other approximately fol-lowing light-cone trajectories. For Bjorken flow [53]of transversally homogeneous and longitudinally boost-invariant systems, macroscopic fields such as the energydensity, pressure and shear stress, can depend neither onthe transverse coordinates (x, y) nor on the space-timerapidity ηs, but only on the longitudinal proper time τ .The hydrodynamic evolution equations thus reduce to aset of coupled ordinary differential equations (ODEs) inτ . The flow is irrotational (ωµν = 0) and unaccelerated(uµ = (1, 0, 0, 0), uµ = 0), but (owing to the curvilinearnature of Milne coordinates) it has a non-zero local ex-pansion rate, θ = 1/τ , and velocity shear, e.g. σηη =−2/(3τ3).3 Symmetries further constrain the shear ten-sor to be diagonal and space-like in Milne coordinates,leaving only one independent component which we taketo be the ηη component: πxx = πyy = −τ2πηη/2 ≡ π/2.4

Using the following relations that hold for Bjorken flow,

π〈ηη〉=− 1

τ2

dτ, π〈ηγ σ

η〉γ =− π

3τ3, π〈ηγ π

η〉γ =− π2

2τ2,

πρ〈ηπη〉γσργ = − π2

2τ3, ∇〈η∇γπη〉γ =

3τ4, (7)

∇γ∇〈ηπη〉γ =4π

3τ4, ∇2π〈ηη〉 =

3τ4, πργσργ =

π

τ,

the shear evolution equations (4)-(6) can be brought intothe following generic form:

dτ= −1

τ

(4

3ε− π

), (8)

dτ= − π

τπ+

1

τ

[4

3βπ −

(λ+

4

3

)π − χπ

2

βπ

]. (9)

The coefficients βπ, a, λ, χ, and γ appearing in Eq. (9)above and in Eq. (11) below are tabulated in Table I forthe three theories studied in this work.

3 To avoid clutter we drop the subscript on ηs whenever we use itas a sub- or superscript.

4 From here on we will use π (not to be confused with the mathe-matical constant denoted by the same symbol), or its normalizedversion π ≡ π/(ε+P ) = 3π/(4ε), as the independent shear stresscomponent. Note that for Bjorken flow the Navier-Stokes valuefor π is positive, πNS ≥ 0.

Since βπ = 4P/5 = 4ε/15, Eqs. (8) and (9) are mu-tually coupled. Eq. (9) for the shear stress can be com-pletely decoupled from the evolution of the energy den-sity by rewriting it in terms of the normalized shear stress(inverse Reynolds number) π = π/(ε+P ) = π/(4P ). In-troducing at the same time the rescaled time variable[30] τ ≡ τ/τπ (which is the inverse Knudsen number forBjorken flow), Eq. (8) can be used to obtain the relation

π = 3(ττ

) dτdτ− 2. (10)

Here we also used that for a conformal system ε∝T 4

and Tτπ = 5η = const. where η ≡ η/s is the specificshear viscosity. Equations (9) and (10) can now be com-bined to obtain a first-order nonlinear ordinary differ-ential equation (ODE) for the inverse Reynolds numberthat is completely decoupled5 from the evolution of theenergy density:6(

π + 2

3

)dπ

dτ= −π +

1

τ

(a− λ π − γ π2

). (11)

The three hydrodynamic theories studied here can beselected by choosing for λ and γ the appropriate com-binations of constants given in Table I. All three the-ories share the same constant a = 4/15, but we cansolve Eq. (11) numerically7 (and the very closely relatedEq. (28) to be discussed in Sec. V even analytically) forgeneral a and will therefore keep it as a free parameteruntil the end.

An important feature that we observe in Eq. (11) isthat the derivative dπ/dτ diverges for π → −2, indicat-ing a discontinuity in π at −2. This feature is not presentin Eq. (9) and is merely an artifact of changing the evolu-tion parameter from τ to τ . Moreover, we notice that the

βπ a λ χ γ

MIS 4P/5 4/15 0 0 4/3

DNMR 4P/5 4/15 10/21 0 4/3

Third-order 4P/5 4/15 10/21 72/245 412/147

TABLE I. Coefficients for the causal viscous hydrodynamicevolution of the shear stress in Bjorken flow for the threetheories studied in this work.

5 This decoupling works as long as the relaxation time has apower law temperature dependence, τπ ∝ T−∆, in which casein Eq. (11) the coefficient multiplying dπ/dτ must be general-

ized to∆(π−1)+3

3.

6 Even if the relaxation time τπ depends on temperature (as itdoes, for example, in systems with conformal symmetry), thetemperature evolution has been completely absorbed into therescaled time variable (inverse Knudsen number) τ .

7 Equation (11) has the form of an Abel differential equation of thesecond kind for which, to the best of our knowledge, an analyticalsolution does not exist.

Page 5: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

5

transverse pressure, PT ≡ P +π/2 = P (1+2π), becomesnegative for π < −1/2 indicating cavitation in the trans-verse direction which may lead to mechanical instability.Therefore one may conclude that π = −2 already lies inthe physically excluded region. In the next section, westudy several other interesting properties inherent in thesolutions of Eq. (11).

IV. GRADIENT EXPANSION, LYAPUNOVEXPONENTS AND ATTRACTORS

Consistent formulations of relativistic dissipativehydrodynamics involve short-lived non-hydrodynamicmodes. These cannot be captured by a standard gra-dient expansion in terms of the Bjorken expansion rateθ = 1/τ , causing such an expansion to be asymptotic,with zero radius of convergence [25]. Borel resummationof this divergent series leads to a hydrodynamic attractorwhich is well defined even for large gradients. The exis-tence of an attractor in a theory is indicated by negativeLyapunov exponents8 which govern the rate at which thesystem loses information about its initial conditions andevolves towards the attractor. In this section, we studythe gradient expansion, Lyapunov exponents and theirimplications for attractor behavior in solutions of hydro-dynamic equations for Bjorken flow.

A. Gradient expansion

To start with we consider the late proper-time expan-sion of π. Substituting a power series ansatz for π interms of powers of 1/τ ,

π(τ) =

∞∑n=0

cnτn, (12)

into the non-linear ODE (11), one finds a recursion rela-tion for the coefficients cn (n ≥ 1)

cn = a δn,1 +

[2

3(n− 1)− λ

]cn−1

+

n∑m=1

(m−1

3− γ)cn−m cm−1, (13)

with initial value c0 = 0. The first non-vanishing coeffi-cient c1 = a corresponds to the Navier-Stokes term in thegradient expansion.

8 Lyapunov exponents characterize the rate of separation of ini-tially infinitesimally close trajectories in a dynamical system. ALyapunov exponent is defined as

|s(t)| ≈ eΛt |s0|,

where s0 and s(t) are the separation between two trajectoriesat initial time t0 and at a later time t, respectively. NegativeΛ indicates the rate at which the system approaches a regularattractor.

MIS

DNMR

Third-order

0 5 10 15 200

2

4

6

8

10

12

14

n

|cn+

1/c

n|

0 5 10 15 200

5

10

15

FIG. 1. Factorial behavior of coefficients (13) for the threehydrodynamic theories indicated in the legend. In the insetthe black line represents |cn+1/cn| = 2

3n and the red dots are

for MIS theory.

For large n the behavior of the coefficients cn is dom-inated by the term proportional to cn−1 in Eq. (13) andshows factorial growth. This is shown in Fig. 1 wherethe ratio of consecutive coefficients, |cn+1/cn|, is plottedagainst n; as shown in the inset this ratio is proportionalto n for n & 5. Divergence of the hydrodynamic gradientexpansion was found before in a variety of hydrodynamictheories [26, 29, 35]. A new observation from Fig. 1 isthat, while the coefficients of the DNMR and third-ordertheories are very similar, those for MIS theory feature aconstant upward shift which can be attributed to λ = 0in Eq. (13). The inset of Fig. 1 shows the ratio |cn+1/cn|for MIS theory (λ= 0, red dots) compared with the sameratio obtained from Eq. (13) without the nonlinear lastterm (black line). One concludes that this nonlinear termplays no role in the asymptotic factorial growth of theexpansion coefficients. The series solution (12) is thusdominated by the term proportional to cn−1 in Eq. (13).We verified that the same statement holds for the DNMRand third-order theories.

B. “Effective MIS” and Lyapunov exponents

The nonlinear last term in Eq. (13) arises from thequadratic terms ∝ π2, dπ2/dτ in Eq. (11). At late timesπ� 1, hence π2� π, and the nonlinear terms can be ig-nored in (13). The late-time behavior in MIS theory isthus controlled by

2

3

dτ= −π +

a

τ(14)

which we call “effective MIS theory”. Its analytic solu-tion is

π = α e−32 τ +

3a

2e−

32 τ Ei

[3τ

2

](15)

where Ei[z] = −∫∞−z (e−t/t) dt is the exponential integral

and α is the integration constant. Equation (15) implies

Page 6: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

6

that the separation between two solutions for π that areinitialized with different initial conditions is damped ex-ponentially:

∂π

∂α∼ exp

(−3

). (16)

The negative Lyapunov exponent Λ = −3/2 in this equa-tion confirms the existence of an attractor. This attrac-tor, given by Eq. (15) with α= 0, is shown in Fig. 2 asthe solid black line, together with a swarm of particularsolutions of Eq. (14) with different integration constantsα 6= 0 (dashed lines) and the asymptotic Navier-Stokessolution (red solid line) for comparison. One sees thatthe attractor, as well as the other solutions with differ-ent initial conditions, join the asymptotic Navier-Stokesbehavior after τ & 3, i.e. for Knudsen number Kn. 0.3.

The solution (15) can also be written as

π(τ) = π0 e− 3

2 (τ−τ0) +3a

2e−

32 τ

∫ τ

τ0

e32 τ

τ ′dτ ′ (17)

where π0 ≡ π(τ0) is the initial condition at time τ0. FromEq. (17) one can extract the gradient expansion for π byintegrating the last term by parts,

∫ τ

τ0

e32 τ

τ ′dτ ′ =

[m∑n=1

(n−1)!e

32 τ

(3τ ′/2)n

]ττ0

+3

2m!

∫ τ

τ0

dτ ′e

32 τ

(3τ ′/2)m+1, (18)

resulting in

π =e−32 (τ−τ0)

(π0 − a

m∑n=1

cnτn0

)+ a

m∑n=1

cnτn

(19)

+ a

(3

2

)2

m! e−3τ/2

∫ τ

τ0

dτ ′e

32 τ

(3τ ′/2)m+1,

where cn = (n−1)!/(3/2)n−1. The first term withinparentheses contains all dependence on the initial con-ditions memory of which is here seen to decay exponen-tially with a decay time of 2

3τπ. The second term is thedivergent gradient series (12) up to order m. The thirdterm can be thought of as the error in approximating thelate time solution for π using a truncated gradient seriesof order m. Minimizing the error term with respect to mwould give an estimate of the optimal truncation for thegradient series at a given value of τ .

We point out that Eq. (17) also admits a convergentseries in positive powers of τ of the form

π =π0 e− 3

2 (τ−τ0) +3a

2e−

32 τ log

τ0

)+

3a

2e−

32 τ∞∑n=1

(3/2)n

n!n(τn − τn0 ) . (20)

NS

πα = 0

πα≠ 0

0.05 0.10 0.50 1 5-1.0

-0.5

0.0

0.5

1.0

τ

π

FIG. 2. Comparison of Eq. (15) (solid black (α= 0) anddashed blue (α 6= 0) lines) with the asymptotic Navier-Stokesbehavior (solid red line). See text for discussion.

It is obtained by expanding the exponential term e32 τ

′in

the integral solution for π into a power series. This seriesis, however, of limited use at late times as it would requireincluding a large number of terms for convergence.

We conclude this subsection by offering another simple,yet interesting, form of the “effective MIS” solution forthe inverse Reynolds number:

π = π0 e− 3

2 (τ−τ0) +3a

2e−

32 [τ−β(τ ;τ0)] τ − τ0

β(τ ; τ0). (21)

Here the function β(τ ; τ0) has units of time, and for anygiven pair (τ0, τ) its value can be shown to lie in the inter-val [τ0, τ ]. To obtain this form we used the mean valuetheorem for the integral in Eq. (17). It would be illu-minating to derive an approximate analytical functionalform for β(τ ; τ0); this is left for future work.

C. Lyapunov exponents from linear perturbation

The exercise in the preceding subsection demonstratedexponential loss of memory of initial conditions in the“modified MIS” theory, on a time scale ∼ 2

3τπ. Onearrives at similar conclusions without the assumption ofignoring in Eq. (11) the terms quadratic in π, by studyingthe fate of a linear perturbation around late-time solu-tions of the full equation (11). This is important because,as we will see further below, the effects on the evolutionof π of the non-linear terms that were ignored in “mod-ified” MIS theory last much longer than the memory-decay time over which initial-state information is erased.

We employ linear perturbation theory around the so-lution π, i.e. we set π → π + δπ and insert this intothe generic evolution equation (11). To simplify the non-linear terms ∼ πδπ we expand the solution π for latetimes according to Eq. (12), keeping only the first non-vanishing term which yields the Navier-Stokes approxi-mation: π ≈ a/τ . Keeping terms up to first order in 1/τ

Page 7: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

7

we find

d(δπ)

dτ= −3

2δπ

[1 +

2λ− a2τ

], (22)

with the solution

δπ ∼ τ 34 (a−2λ) exp

(−3

)+O

(1

τ2

). (23)

The prefactor τ34a in front of the exponential decay fac-

tor is easily traced back to the quadratic term π (dπ/dτ)on the left hand side of Eq. (11). The λ-dependenceof this prefactor was not seen in the preceding subsec-tion because λ= 0 for MIS theory. The solution (23) ex-hibits the same negative Lyapunov exponent Λ = −3/2as Eq. (16). This shows that keeping the non-linear termsin Eq. (11) does not change the initial state memory lossrate with which the system approaches the hydrodynamicattractor at late times. This rate is entirely controlledby microscopic physics and independent of the precisemacroscopic dynamical state of the expanding system.

D. Lyapunov exponents from Borel resummation

The results in the preceding subsection may also belooked at from the perspective of Borel resummation,which replaces a divergent series

∑n an by [87]

B

(∑n

an

)≡∫ ∞

0

du e−u∑n

ann!un. (24)

Borel resummation interchanges the order in which thesum and integral in

∑n an =

∑n(an/n!)

∫∞0du e−u un

are performed. While for divergent series this yields in-equivalent results, Borel resummation of asymptotic se-ries can help with their interpretation, as shown below.

Starting from the recursion relation (13) and ignoringthe non-linearities therein one finds the coefficients of thegradient series to be

cn = C Γ(n−3λ/2)

(3/2)n−3λ/2, n ≥ 1, (25)

where the normalisation C ≡ a(3/2)1−3λ/2/Γ(1−3λ/2)ensures c1 = a. The Borel resummed version of Eq. (12)reads

B

(π(τ)

)= C τ−3λ/2

∫ ∞0

du

ue−(3/2)τu u−3λ/2

∞∑n=1

un

= C τ−3λ/2

∫ ∞0

du e−(3/2)τu u−3λ/2

1− u. (26)

The Borel integrand has a pole at u= 1 which results inan ambiguous imaginary part of the sum, Im [B(π)] =±π C e−3τ/2τ−3λ/2, arising from non-unique choices ofdeforming the contour around the pole. In order to re-move the ambiguity in the Borel resummed gradient se-ries one must include in Eq. (12) a “non-hydrodynamic”

NS

MIS

DNMR

Third-order

exact RTA

0.05 0.10 0.50 1 5 100.0

0.1

0.2

0.3

0.4

0.5

τ

π

FIG. 3. Numerical attractors for the inverse Reynolds num-ber π(τ) for the MIS (dashed line), DNMR (dash-dotted line),and third-order (solid line) theories, compared with the exactnumerical attractor of the RTA Boltzmann equation (filledcircles). Also shown for comparison is the Navier-Stokes so-lution (dotted).

term δπ ∼ e−3τ/2τ−3λ/2.9 Notice the similarity betweenthis non-hydrodynamic term and that obtained via themethod of linear perturbation in Eq. (23).10

E. Numerical attractors for various theories

We close this section with a numerical study of the at-tractor solution of Eq. (11) towards which specific solu-tions with generic initial conditions decay exponentially.To identify the attractor we follow the prescription out-lined in Ref. [26]: The initial condition for the attractorsolution is obtained by imposing the boundary condi-tions that both π and dπ/dτ remain finite as τ → 0.As shown in [26] this results in the quadratic equationγ π2 + λ π − a = 0 for the initial value of π. One findstwo solutions, one of which (the positive root) is stableand corresponds to the attractor solution whereas thenegative root corresponds to a repulsor. With this ini-tial condition we then solve Eq. (11) numerically, for thedifferent parameter combinations listed in Table I. Theresulting attractors for the MIS, DNMR, and third-ordertheories are shown in Fig. 3 where we also compare themwith the exact numerical attractor of the RTA Boltzmann

9 For a detailed study of this issue see the discussion of BRSSStheory [84] in Ref. [28].

10 The extra factor of τ3a/4 in the latter stems from non-linearterms in Eq. (13) which were here ignored. It should be em-phasized that when considering the full non-linear version ofEq. (11), not just one, but an entire series of exponentiallydamped terms must be added to the gradient series, turningit into a trans-series [26, 28]. In this case the above-mentionednon-hydrodynamic term plays the role of the leading order cor-rection.

Page 8: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

8

equation [37]. Please note that, unlike any particular so-lution of Eq. (11) which depends on both the initial con-dition π(τ0) and the initial time τ0 at which it is imposed,the attractors are universal, i.e. they attract any partic-ular solution with initial conditions within their basin ofattraction, irrespective of its starting time.

In Fig. 3, we show the attractor solutions for π for theMIS, DNMR, and third-order theories, as well as for theexact solution of RTA Boltzmann equation [45] and theNavier-Stokes solution. We see that of these the MISattractor approaches the exact attractor most slowly,11

while the attractor of the third-order theory exhibits thebest agreement with the exact RTA BE attractor, al-most as good as the anisotropic hydrodynamic (aHydro)attractor studied in [38, 42]. This adds to the evidence ofthe superior performance of the third-order theory overdifferent variants of second-order theories that are basedon expansions around a locally isotropic momentum dis-tribution.12

V. APPROXIMATE ANALYTICAL SOLUTIONS

Up to this point we focused our discussion of the evolu-tion of the inverse Reynolds number π on Eq. (11) whichholds for conformal systems where Tτπ = const. It hasthe advantage of completely absorbing any dependence ofthe shear relaxation time τπ on the energy density or tem-perature into the rescaled time variable (inverse Knudsennumber) τ , but at the expense of not being able to solvethis ODE analytically. In this section we derive analyti-cal solutions for the evolution of π for Bjorken flow, at theexpense of not being able to ensure the conformal rela-tion Tτπ = const. consistently with the evolution of theenergy density. Instead, we find three separate classesof analytical solutions, corresponding to three differentapproximations of τπ as a function of time.

Starting from Eqs. (8),(9), we decouple them as be-fore by rewriting them in terms of the inverse Reynoldsnumber π but without rescaling the time [81]:

1

ετ4/3

d(ετ4/3)

dτ=

4

3

π

τ, (27)

dτ= − π

τπ+

1

τ

(a− λπ − γπ2

). (28)

In the following, we find analytical solutions of Eq. (28),using different approximations for the form of shear re-laxation time τπ.

11 In light of footnote 2 this should perhaps not be too surprising.12 aHydro, a second-order approach that is based on an expansion

around a self-consistently adjusted ellipsoidally deformed localmomentum distribution [38, 74, 76, 77, 79, 80], performs evenbetter than the third-order theory [42].

A. Constant relaxation time

In this subsection we ignore the scaling of τπ with tem-perature, by simply setting it constant. This constitutesa rather drastic violation of conformal symmetry by in-troducing, in addition to the inverse temperature 1/T , asecond, independent length scale τπ. In the following twosubsections we will successively improve on this approx-imation.

The first analytical solution for the evolution in second-order hydrodynamics with Bjorken flow of the energydensity and inverse Reynolds number for a constant shearrelaxation time τπ was found by Denicol and Noronha[35]. For completeness we briefly review this solution,generalizing it to the generic form (27),(28) of the evolu-tion equations which also includes third-order hydrody-namics. Introducing again the rescaled time τ = τ/τπ,for constant τπ Eq. (28) turns directly into

dτ= −π +

1

τ

(a− λπ − γπ2

)(29)

which is similar to Eq (11) but without the nonlinearityon the left hand side.13 As will be discussed below, thisdifference has important consequences for the attractorsolutions and Lyapunov exponents.

Equation (29) is a first-order nonlinear ODE of Riccatitype which can be written as a second-order linear ODEwith the help of the following transformation of variables,

1

y

dy

dτ= γ

π

τ⇐⇒ π =

τ

γy

dy

dτ, (30)

which turns Eq. (29) into

d2y

dτ2+

(1 +

1 + λ

τ

)dy

dτ− aγ

τ2y = 0. (31)

The general solution of this linear ODE can be expressedin terms of Whittaker functions Mk,m(τ) and Wk,m(τ)[35]14:

y(τ) = Aτke−τ/2 [Mk,m(τ) + α Wk,m(τ)] . (32)

Here k = − 12 (λ+1) and m = 1

2

√4aγ + λ2 while A and

α are arbitrary constants. Substituting this solution inEqs. (30) and (27) one finds

π(τ) =(2k+2m+1)Mk+1,m(τ)− 2αWk+1,m(τ)

2γ [Mk,m(τ) + αWk,m(τ)], (33)

ε(τ) =ε0

( τ0τ

)43 (1− kγ )

e−2

3γ (τ−τ0)

×(Mk,m(τ) + αWk,m(τ)

Mk,m(τ0) + αWk,m(τ0)

) 43γ

. (34)

13 Equation (29) can be readily derived by setting ∆ = 0 in foot-note 5.

14 Note that the authors of [35] used γ = 4/3 and a different defi-nition of a.

Page 9: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

9

Here ε0 is the initial energy density at time τ0, and theconstant α encodes the initial normalized shear stress π0.Note that α can only take values for which the energydensity is positive-definite for τ > 0.

It is easy to see from Eq. (33) that the solution for π(τ)loses all memory about initial conditions at late timeswhen Mk,m(τ) dominates over Wk,m(τ): for large argu-ments τ →∞ the ratio

Wk,m(τ)

Mk,m(τ)−→

Γ(m−k+ 1

2

)Γ(2m+1)

τ2k e−τ (35)

decays exponentially. Thus at late times the terms pro-portional to α in (33), (34) decay like e−τ , correspondingto a Lyapunov exponent of Λ = − 1. This obviously dif-fers from Λ = − 3

2 for the conformally invariant theoriesdescribed by Eq. (11); the difference is a direct conse-quence of the breaking of conformal symmetry by settingτπ constant instead of ∝ 1/T .

B. Relaxation time from ideal hydrodynamics

A better approximation to Eq. (11) can be obtainedby setting Tτπ = const. but, instead of using the exacttime dependence of the temperature T , approximatingthe latter such that the evolution equations can still beintegrated analytically. Seeing that for Bjorken flow thesystem asymptotically approaches local thermal equilib-rium, we may approximate the time-dependence of T atlate times by the ideal fluid law [53]

Tid(τ) = T0

(τ0τ

)1/3

, (36)

where T0 is the temperature at initial time τ0. For Tτπ =5η this yields

τπ(τ) = b τ1/3, with b =5η

T0τ1/30

. (37)

Using this to define the scaled time variable τ ≡ τ/τπone finds

dτ=

2

3τπ=

2

3bτ−1/3,

and Eq. (28) turns into

2

3

dτ= −π +

1

τ

(a− λπ − γπ2

), (38)

independent of b. This equation again misses the nonlin-ear term on the left hand side (l.h.s.) of Eq. (11) and,except for the factor 2/3 on the l.h.s., has the same struc-ture as Eq. (29). Its analytical solution is therefore verysimilar to Eq. (33), except for a change of the argument

of the Whittaker functions by a factor 2/3:

π(τ) =(2k+2m+1)Mk+1,m(3τ /2)− 2αWk+1,m(3τ /2)

3γ [Mk,m(3τ /2) + αWk,m(3τ /2)],

(39)

ε(τ) =ε0

( τ0τ

)43 ( 3

2−kγ )e−

1γ (τ−τ0)

×(Mk,m(3τ /2) + αWk,m(3τ /2)

Mk,m(3τ0/2) + αWk,m(3τ0/2)

) 43γ

. (40)

Here k = − 3λ+24 and m = 3

4

√4aγ + λ2, different from

Eqs. (33)-(34). The asymptotic behavior (35) of theWhittaker functions now tells us that, for the choice (37),memory of the initial conditions is lost exponentially ac-cording to e−

32 τ , corresponding to the same Lyapunov

exponent Λ = − 32 as for the conformally invariant theo-

ries described by Eq. (11).

C. Relaxation time from Navier-Stokes evolution

We can further improve our approximation by account-ing for first-order gradient effects in the evolution of thetemperature, by replacing the ideal fluid law (36) by theNavier-Stokes result [83, 88, 89]

TNS

= T0

(τ0τ

)1/3[1 +

3τ0T0

{1−

(τ0τ

)2/3}]

. (41)

For η= 0 this reduces to (36). Substituting this intoTτπ = 5η we find

τπ =τ1/3

d− 215τ−2/3

, d ≡(T0τ05η

+2

15

)τ−2/30 . (42)

For the scaled time variable τ ≡ τ/τπ we now have

dτ=

2

3τπ

(1 +

2

15τ

). (43)

Using this in Eq. (28) one obtains(a/τ + 2

3

)dπ

dτ= −π +

1

τ

(a− λπ − γπ2

), (44)

independent of the constant d. This shares with Eq. (38)the factor 2/3 on the l.h.s. which, as we saw in thepreceding subsection, leads to the correct Lyapunov ex-ponent for conformally symmetric systems. Comparingwith Eq. (11) one sees that they are identical up to thesubstitution π 7→ a/τ (which is the first non-zero termin the series expansion (12)). This should not be sur-prising as the term within parenthesis on the l.h.s. ofEq. (11) stems solely from the energy (or, equivalently,temperature) evolution equation (8) which, when mak-ing the replacement π 7→ a/τ , leads to the Navier-Stokessolution T

NS.

Comparing with the preceding subsection, this sug-gests that it might be possible to account for the time

Page 10: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

10

evolution of the temperature in the relation τπ(τ) ∼1/T (τ) with ever increasing precision by substituting thegradient series (12) for π in the prefactor on the l.h.s.of Eq. (11) and truncating it at increasingly higher or-der: zeroth order for ideal hydrodynamics, first order forNavier-Stokes dynamics, and so on. Unfortunately, this isnot justified as Eq. (12) is an asymptotic (i.e. divergent)series. The consequences of this on the hydrodynamicattractor will be discussed in the next section.

Equation (44) can be recast in a form similar toEq. (29) and solved again analytically:

π(τ) =(2k+2m+1)Mk+1,m(w)− 2αWk+1,m(w)

3γ [Mk,m(w) + αWk,m(w)], (45)

ε(τ) =ε0

(w0

w

)43 ( 3

2−kγ )e−

1γ (τ−τ0)

×(Mk,m(w) + αWk,m(w)

Mk,m(w0) + αWk,m(w0)

) 43γ

. (46)

This looks formally identical to Eqs. (39)-(40), except forthe substitution 3

2 τ 7→ w ≡ 32 (τ + a

2 ) in the arguments ofthe Whittaker functions on the right hand side (r.h.s.),together with modified definitions for the indices k andm:

k =3a− 4− 6λ

8, m =

3

8

√a2 + 16aγ − 4aλ+ 4λ2.

Using Eq. (35) for the asymptotic behavior of the Whit-taker functions we see that once again memory of theinitial conditions is lost exponentially according to e−

32 τ ,

with no effect from the constant shift of the time vari-able in the arguments of the Whittaker functions. Thiscorresponds to the same Lyapunov exponent Λ = − 3

2 asin the preceding subsection and, more generally, for allthe conformally invariant theories described by Eq. (11).

We note in passing that including higher order termscn/τ

n for n> 1 in the gradient series while approximatingπ on the l.h.s. of Eq. (44) spoils its reducibility to theanalytically solvable form explored in this work. Whetherfor such approximations Eq. (11) can still be solved interms of known functions remains to be seen.

We summarize this section by observing that, while theassumption of a fixed shear relaxation time τπ (i.e. of afixed temperature T when writing τπ ∝ 1/T ) leads to anincorrect Lyapunov exponent describing the rate of ap-proach towards the hydrodynamic attractor, the correctdecay rate is recovered as soon as one allows the temper-ature T to vary with time hydrodynamically even if theexact time dependence is replaced by an approximationbased on a truncated hydrodynamic gradient expansion.

VI. ANALYTICAL ATTRACTORS

In this section we investigate the hydrodynamic attrac-tors associated with the analytic approximate solutionsof the evolution equation for the inverse Reynolds num-ber π found in the preceding section. In Section IV E

Attractor

Other solutions

0.05 0.10 0.50 1 5-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

τ

π

FIG. 4. Attractor behavior of the analytical solution (39)for the third-order theory (λ= 10/21, γ= 412/147). The two(attracting and repulsing) fixed points at τ → 0 are clearlyvisible. Please note that some of the dashed lines have poles,i.e. as τ → 0 they go to +∞ before reappearing from −∞and approaching the negative fixed point.

we saw that, as τ → 0, π can take one of only two fi-nite values of opposite sign, and we identified the attrac-tor as the unique solution which connects to the positivevalue. All other solutions were found to connect in thelimit τ → 0 to the negative value. An illustration of thisgeneric behavior is shown in Fig. 4 for the case of the at-tractor corresponding to the analytical solution (39) forthe third-order theory.

In this section we introduce the following procedure foridentifying the hydrodynamic attractor [81]: In terms ofthe parameter α encoding the initial condition for π, wesearch for the value α0 at which the quantity

ψ(α0) ≡ limτ→τ0

∂π

∂α

∣∣∣∣α=α0

(47)

diverges at the time τ0 where the two fixed points of theevolution trajectories (see Fig. 4) are located [32]. ForBjorken flow, this time is usually τ0 = 0. For the numer-ical solution of Eq. (11) in Sec. IV E this can be seen inFig. 3, and for the analytical solutions Eqs. (33) and (39)in Secs. V A and V B, respectively, this can be seen bystudying their behavior near τ = 0. The structural sim-ilarity of Eqs. (45) and (39) shows that for the solution(45) the fixed points are instead located at w0 = 0, whichcorresponds to a negative (i.e. unphysical) longitudinalproper time τ0 = −a/2. This arises from the Navier-Stokes substitution π 7→ a/τ on the l.h.s. of Eq. (44)which breaks down (and thus renders the analytic solu-tion (45) unreliable) in the region τ � 1.

We can use the same approach of studying the sensitiv-ity to initial conditions to obtain the Lyapunov exponentΛ from the formula

Λ = limτ→∞

∂τ

[ln

(∂π

∂α

)]. (48)

For the analytical solutions (33), (39), (45) this prescrip-tion reproduces the same results as obtained from the

Page 11: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

11

late-time behavior (35) of the Whittaker functions but itsadvantage is that it can also be used numerically whereexact solutions for π are not available (such as for thenumerical solutions of the generic equation (11)).

For the case shown in Fig. 4 we have verified explicitlythat indeed there are two fixed points at τ0 = 0 and thatonly for the attractor solution, characterized by α0 = 0,ψ(0) (defined in Eq. (47)) diverges. For all other solutionsα 6= 0 we found that ψ(α) = 0 in the limit τ → 0,indicating that they all converge to the negative branch,as shown in Fig. 4.

The approximate analytical solutions (33), (39), (45)discussed in Sec. V can be written in the generic form

π(w) =(k+m+ 1

2 )Mk+1,m(w)− αWk+1,m(w)

γ|Λ| [Mk,m(w) + αWk,m(w)], (49)

τπ= const.

τπ~ 1/Tid

τπ~ 1/TNS

exact attr.

exact RTA

0.1

0.2

0.3

0.4

π

MIS

(a)

0.05

0.10

0.15

0.20

0.25

0.30

π

DNMR

(b)

0.05 0.10 0.50 1 5 10

0.05

0.10

0.15

0.20

0.25

τ

π

Third-order

(c)

FIG. 5. Approximate analytical attractors for the MIS (a),DNMR (b), and third-order (c) theories, compared with theirexact numerical attractors (solid green lines) and the exactanalytical attractor for the RTA Boltzmann equation (blackdots).

T (τ) w Λ k m

const. τ −1 − 12

(λ+1) 12

√4aγ+λ2

ideal 32τ − 3

2− 3λ+2

434

√4aγ+λ2

NS 32

(τ+a

2

)− 3

2− 6λ+4−3a

838

√16aγ+a2−4aλ+4λ2

TABLE II. Arguments and parameters of Eq. (49) for theanalytic approximations studied in Secs. V A, B, and C, re-spectively.

with arguments and parameters for the three cases (i.e.for τπ ∝ 1/T with T either constant or with time depen-dence taken from ideal or Navier-Stokes hydrodynamics)compiled for convenience in Table II. The attractor so-lutions are obtained from Eq. (49) by setting the initialcondition parameter α= 0:

πattr(w) =k+m+ 1

2

γ|Λ|Mk+1,m(w)

Mk,m(w). (50)

They are shown in Fig. 5 for the three different hydrody-namic theories discussed in this paper (MIS (a), DNMR(b), and third-order (c)) and compared with the corre-sponding exact numerical attractors as well as with theattractor for the exact analytical solution of the RTABoltzmann equation [37] (the latter is, of course, thesame in all three subpanels). Comparison of these attrac-tors provides insights not only about the performance ofthe three different hydrodynamic theories as approxima-tions to the underlying kinetic theory, but also about therelative accuracy of the additional approximations madein Sec. V in order to obtain analytical results.

For all three hydrodynamic theories, we note (espe-cially at early times when the system is farthest awayfrom local equilibrium) that the differences between theexact numerical attractors and their analytical approxi-mations from Sec. V are significantly smaller than theirdiscrepancy from the attractor of the underlying kinetictheory. At late times the breaking of conformal symme-try by choosing a constant relaxation time leads gener-ally to the largest difference between the exact numericaland analytically approximated hydrodynamic attractors;this may be attributed to the fact that the τπ = const.approximation underestimates the rate of approach to-wards the attractor by a factor 2/3. More surprisingly,the analytic approximation that performs best at latetimes (which uses τπ(τ) ∼ 1/TNS(τ)) performs worst atearly times when compared with the exact numerical re-sult. This reflects a different value for the fixed pointof the inverse Reynolds number compared to the otheranalytic approximations and may be related to the factthat in this case the fixed point is shifted outside thephysical region to τ = −a/2. As already seen in Fig. 3,in comparison with the exact solution of the RTA Boltz-mann equation, the third-order theory performs muchbetter than both the MIS and DNMR theories, bothwhen evaluated exactly numerically or approximately an-

Page 12: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

12

alytically. The only known theory that performs evenbetter than the third-order theory studied here is second-order anisotropic hydrodynamics [38, 42, 71] which effec-tively resums terms of all orders in the inverse Reynoldsnumber.

Numerical studies, such as those shown in Figs. 2 and4, show that at late times τ > 1 any initial deviationfrom the attractor approaches the attractor exponen-tially, with the Lyapunov exponents discussed before.The approximate analytical result (49) allows to under-stand this approach analytically over the entire range ofτ , i.e. also for large Knudsen numbers τ � 1. Followingthe analysis [90] we write15

π(w) = πattr(w) + δ(w)

= π(w)|α=0 + α∂π

∂α(w)

∣∣∣∣α=0

+O(α2) (51)

where in the last expression we expanded the deviation δto first order in the deviation of the initial value param-eter α from the value α= 0 characterizing the attractorsolution. The decay of the inverse Reynolds number π to-wards its attractor value is for small deviations16 δ givenby

δ(w) =− α

γ|Λ|Mk+1,m(w)

Mk,m(w)(52)

×[(k+m+ 1

2

) Wk,m(w)

Mk,m(w)+Wk+1,m(w)

Mk+1,m(w)

].

This is a function of the scaling variable w ∝ |Λ|τ whichis proportional to the inverse Knudsen number τ . Atfirst sight this suggests that the competition between theglobal expansion rate 1/τ and the microscopic relaxationrate 1/τπ not only rules the evolution of the hydrody-namic attractor πattr itself, but also the way initial ex-cursions of the inverse Reynolds number from this attrac-tor decay as the full dynamical solution approaches theattractor. However, as first pointed out in [90], a deeperanalysis exhibits that the dependence of the lifetime ofsuch excursions on the interaction rate 1/τπ changes dra-matically between early times τ� τπ/|Λ| and late timesτ� τπ/|Λ|.

Using the following Whittaker function properties forsmall arguments,

Wk,m(x)

Mk,m(x)

∣∣∣∣x→0

≈ Γ[2m]

Γ[m−k+ 12 ]x−2m,

Mk+1,m(x)

Mk,m(x)

∣∣∣∣x→0

≈ 1,

(53)where the approximation holds at leading order, we seefrom Eq. (52) that, in contrast to the exponential decay

15 Note that δ as defined in [90] differs from ours by a factor 4/3.16 It is worth pointing out that, for fixed initial deviation δ0 at

initial time τ0, the corresponding initial state parameter α ap-proaches 0 as τ0 → 0. Small δ0 thus implies small α (especiallyfor small values of τ0), but not vice versa.

(35) at late times τ � 1, the decay of excursions awayfrom the attractor decay with a power law at early timesτ � 1 [90]. The transition from power law to exponentialdecay around w ∝ |Λ|τ = 1 is illustrated in Fig. 6.17,18

Wk,m (x)

Mk,m (x)

~ x-2m

~ x2k ⅇ-x

0.001 0.010 0.100 1 1010-6

10-4

0.01

1

100

104

x

f(x)

1 2 3 4 50.001

0.0050.010

0.0500.100

0.5001

FIG. 6. The ratioWk,m(x)

Mk,m(x)as a function of the scaling variable

x (which is proportional to time in units of the microscopicrelaxation time, i.e. to the inverse Knudsen number). Theexact result (solid black line) is compared with the first or-der approximation (53) for small arguments x < 1 (dashedred line) and the analogous approximation (35) for large ar-guments x > 1 (dashed green line). The power-law decayat early times manifests itself as an approximately straightline in the double-logarithmic representation of the main plotwhile the transition to exponential decay at late times leadsto the approximately straight line behavior seen in the semi-logarithmic inset plot.

An immediate consequence of this discussion is that forearly starting times τ0� τπ/|Λ| (where, for sufficientlysmall δ, δ(τ) ∝ (|Λ|τ)−2m) the decay time τ1/e (defined

17 Note that the “effective MIS” approximation (15) discussed inSec. IV B misses the early-time power-law decay of excursionsaway from the hydrodynamic attractor. We found that early-time power-law decay of initial deviations from the attractor re-quires that in the generic evolution equation (11) for the inverseReynolds number at least one of the two coefficients λ and γmust be nonzero. This is not the case for the “effective MIS”theory.

18 A few comments regarding the domain of validity of the linearisedapproach are in order. Using Eq. (49) to determine the devia-tion δ(w) from the attractor solution, one readily checks that,for w � 1, terms beyond the linear order in Eq. (51) are pro-portional to αn(w−2m)n for n ≥ 2. Even for fixed small valuesof α, this series diverges for sufficiently early “times” w2m � α,where linearization in α must break down. This is closely relatedto footnote 16: in order to keep δ0 ≡ δ(w0) fixed for small valuesof w0, α must be tuned down to ensure αw−2m

0 � 1 such thathigher-order contributions may be safely neglected. The break-down of the linearised method at very early times owes itself tothe presence of the repulsive fixed point at w = 0, i.e., as long asα is not strictly equal to 0, all solutions, π(α,w) will ultimatelyhit the repulsor at w = 0, where the deviation from the attrac-tor no longer stays infinitesimal. This mathematical feature ofπ(α,w = 0) being discontinuous at α = 0 was, in fact, used touniquely determine the attractor solution via Eq. (47) [81].

Page 13: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

13

as the time over which the magnitude of δ decreases bya factor 1/e) is given by τ1/e = ρ τ0 (where, to leadingorder in δ and τ , ρ is a constant given in terms of the at-

tractive fixed point πattr,0 by ρ = (e1/2m−1)3

πattr,0+2 ) andthus independent of τπ, whereas for late starting timesτ0� τπ/|Λ| it is given by τ1/e = τπ/|Λ|, independent ofτ0. τ1/e thus smoothly increases from 0 to τπ/|Λ| (whereit saturates) as τ0 increases from 0 to values much largerthan τ /|Λ|. Sufficiently small excursions of the inverseReynolds number from its attractor thus decay faster atearly times τ� τπ/|Λ| than at late times τ� τπ/|Λ|.

Based on the observation that for τ � τπ/|Λ| the decayrate is controlled by the initial expansion rate 1/τ0 andindependent of the scattering rate 1/τπ whereas the op-posite holds for τ � τπ/|Λ|, the authors of [90] attributethe transition from power-law decay at early times to ex-ponential decay at later times to a transition from the“pre-hydrodynamic” to the “hydrodynamic” stage, i.e.they see it as associated with the process of hydrody-namization. We have a different view of this matter: It isknown from earlier numerical studies [32, 38, 42] of bothBjorken and Gubser flows that some hydrodynamic ap-proximations (in particular anisotropic and third-orderhydrodynamics, but to a somewhat lesser degree alsoDNMR theory) evolve the inverse Reynolds number veryaccurately (when compared with the evolution predictedby the underlying RTA Boltzmann equation) already atvery early times τ � 1, even when initialized far awayfrom the attractor. That is, not only do the attractorsfor these hydrodynamic approximations agree well withthe exact attractor even for large Knudsen numbers, asshown for DNMR and third-order theories here in Fig. 3and for anisotropic hydrodynamics in Fig. 3 of [38], butthe theories also describe accurately the evolution of thesystem towards the attractor even when initialized faraway from it. To us this implies that the system hydro-dynamizes well before τ = 1/|Λ|, and power-law ratherthan exponential decay of deviations from the attractorare not a tell-tale signature for “pre-hydrodynamic” be-havior.

VII. SUMMARY AND CONCLUSIONS

We studied analytically and numerically the evolu-tion of the inverse Reynolds number in causal theoriesof second- and third-order relativistic viscous fluid dy-namics for Bjorken flow. In this situation there is only asingle non-vanishing component of the shear stress, de-scribing an anisotropy between longitudinal and trans-verse pressure, which is generically very large at earlytimes. For Bjorken flow the evolution of the associatedinverse Reynolds number (i.e. the ratio of the shear stressto the enthalpy density of the system) decouples fromthat of the energy density and temperature and thus canbe solved independently. When expressed as a functionof time in units of the microscopic shear relaxation time(which measures the inverse Knudsen number of Bjorkenflow), the solution is universal, i.e. independent of the

specific shear viscosity of the medium. For three dif-ferent macroscopic hydrodynamic theories, we studiedthese solutions, numerically and with various analyticalapproximations, their hydrodynamic attractors, the rateof initial state memory loss and approach to the attractor(expressed through Lyapunov exponents), and comparedall these with the corresponding solution of the relativis-tic Boltzmann equation in RTA approximation which de-scribes the underlying microscopic dynamics and can, forBjorken flow, be solved exactly.

When comparing the exact numerical solutions for theattractor of the inverse Reynolds number for the threedifferent hydrodynamic theories studied here with theexact solution from the Boltzmann equation we find sig-nificant differences at early times, i.e. at large Knud-sen numbers where the dynamics happens far away fromequilibrium. While for the third-order theory the discrep-ancy remains always below 10%,19 it increases to ∼ 30%for DNMR theory and to almost a factor of 2 for MIStheory (both being second-order hydrodynamic theories).Compared to these, the additional discrepancies causedby the various approximations we made to arrive at ana-lytical solutions for the dynamics of π are small. At latetimes (small Knudsen numbers) all attractors approachthe Navier-Stokes solution. Again third-order hydrody-namics is closest to the exact solution of the underly-ing Boltzmann kinetics, but the excellent agreement isspoiled somewhat if conformal symmetry is broken by anapproximation that sets the relaxation time τπ to a con-stant rather than allowing it to vary inversely with thetemperature.

As τ → 0, the Navier-Stokes value of the inverseReynolds number diverges while its attractor value ap-proaches the finite value π = 0.25, corresponding to ashear stress over thermal pressure ratio π/P = 1. TheLyapunov exponent associated with the evolution of π,Λ = − 3

2 , indicates that even far-from-equilibrium initialconditions, π0/P0 � 1, relax exponentially to the attrac-tor value with a decay time of order 2

3τπ = 103ηT . For

minimal specific shear viscosity η = 14π and a medium

temperature of, say, T = 0.5 GeV, this corresponds to adecay time of ≈ 0.1 fm/c. Subsequently, the hydrody-namic evolution follows essentially the hydrodynamic at-tractor of the theory which agrees, within the precisionstated above, with the exact attractor associated with ofthe underlying microscopic Boltzmann kinetics.

While the ODE describing the evolution of the inverseReynolds number for Bjorken flow is easily solved on acomputer, with arbitrary precision, the analytic approxi-mations studied here are surprisingly accurate, and theyyield valuable insights into the details of initial statememory loss and the approach to attractor dynamicsin Bjorken flow. Similar methods may be applicable to

19 It has been shown elsewhere [38, 42] that the attractor foranisotropic hydrodynamics (aHydro) is even closer to the exactBoltzmann attractor than the one for the third-order theory.

Page 14: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

14

different situations (for example, Gubser flow, which isphysically quite different from Bjorken flow but shareswith it many mathematical similarities) where they canlead to similarly valuable qualitative insights. General-ization to Gubser flow should be particularly interestingbecause it does not thermalize at late times but ratherapproaches an asymptotic free-streaming state. In sucha situation the questions of initial state memory loss andthe approach to a hydrodynamic attractor [32, 42, 43]have not yet been fully understood.

ACKNOWLEDGMENTS

S.J. and A.J. acknowledge kind hospitality of IISERPune where parts of this work were completed. We grate-fully acknowledge interesting discussions with A. Be-htash, S. Kamata, M. Martinez and H. Shi as well asthe stimulating effect of Ref. [90] which appeared si-

multaneously when the present work was first circu-lated and which led to several additional insights re-ported in Sec. VI of this final version of our work.U.H. thanks Urs Wiedemann for an important clarify-ing discussion of Ref. [90]. A.J. is supported in part bythe DST-INSPIRE faculty award from the Indian De-partment of Science and Technology under Grant No.DST/INSPIRE/04/2017/000038. The research of U.H.and C.C. was supported in part by the U.S. Departmentof Energy (DOE), Office of Science, Office for NuclearPhysics under Award No. DE-SC0004286 and in part bythe National Science Foundation (NSF) within the frame-work of the JETSCAPE Collaboration under Award No.ACI-1550223. S.J. thanks Pranav S. and C.C. thanksYuri Kovchegov for illuminating discussions. U.H. ac-knowledges the kind hospitality of the Institut fur The-oretische Physik at the J. W. Goethe-Universitat duringa research visit supported by a Research Prize from theAlexander von Humboldt Foundation. He thanks Micha lHeller and Viktor Svensson for fruitful discussions thatmotivated some of the research reported here.

[1] W. Florkowski, M. P. Heller and M. Spaliski, Newtheories of relativistic hydrodynamics in the LHC era,Rept. Prog. Phys. 81 (2018) 046001 [1707.02282].

[2] U. Heinz and P. F. Kolb, Early thermalization at RHIC,Nucl. Phys. A702 (2002) 269 [hep-ph/0111075].

[3] U. Heinz and R. Snellings, Collective flow and viscosityin relativistic heavy-ion collisions, Ann. Rev. Nucl.Part. Sci. 63 (2013) 123 [1301.2826].

[4] A. Jaiswal and V. Roy, Relativistic hydrodynamics inheavy-ion collisions: general aspects and recentdevelopments, Adv. High Energy Phys. 2016 (2016)9623034 [1605.08694].

[5] P. Romatschke and U. Romatschke, Relativistic FluidDynamics In and Out of Equilibrium, CambridgeMonographs on Mathematical Physics. CambridgeUniversity Press, 2019, 10.1017/9781108651998,[1712.05815].

[6] U. Heinz, H. Song and A. K. Chaudhuri, Dissipativehydrodynamics for viscous relativistic fluids, Phys. Rev.C73 (2006) 034904 [nucl-th/0510014].

[7] P. Romatschke and U. Romatschke, Viscosityinformation from relativistic nuclear collisions: howperfect is the fluid observed at RHIC?, Phys. Rev. Lett.99 (2007) 172301 [0706.1522].

[8] H. Song and U. Heinz, Suppression of elliptic flow in aminimally viscous quark-gluon plasma, Phys. Lett.B658 (2008) 279 [0709.0742].

[9] K. Dusling and D. Teaney, Simulating elliptic flow withviscous hydrodynamics, Phys. Rev. C77 (2008) 034905[0710.5932].

[10] H. Song and U. Heinz, Causal viscous hydrodynamics in2+1 dimensions for relativistic heavy-ion collisions,Phys. Rev. C77 (2008) 064901 [0712.3715].

[11] M. Luzum and P. Romatschke, Conformal relativisticviscous hydrodynamics: applications to RHIC results at√sNN = 200 GeV, Phys. Rev. C78 (2008) 034915

[0804.4015].

[12] H. Song and U. Heinz, Multiplicity scaling in ideal andviscous hydrodynamics, Phys. Rev. C78 (2008) 024902[0805.1756].

[13] H. Song, S. A. Bass and U. Heinz, Viscous QCD matterin a hybrid hydrodynamic+Boltzmann approach, Phys.Rev. C83 (2011) 024912 [1012.0555].

[14] U. Heinz, C. Shen and H. Song, The viscosity ofquark-gluon plasma at RHIC and the LHC, AIP Conf.Proc. 1441 (2012) 766 [1108.5323].

[15] H. Song and U. W. Heinz, Extracting the QGP viscosityfrom RHIC data - a status report from viscoushydrodynamics, J. Phys. G36 (2009) 064033[0812.4274].

[16] B. Schenke, S. Jeon and C. Gale, (3+1)D hydrodynamicsimulation of relativistic heavy-ion collisions, Phys. Rev.C82 (2010) 014903 [1004.1408].

[17] H. Song, S. A. Bass, U. Heinz, T. Hirano and C. Shen,200 A GeV Au+Au collisions serve a nearly perfectquark-gluon liquid, Phys. Rev. Lett. 106 (2011) 192301[1011.2783].

[18] H. Song, S. A. Bass, U. Heinz, T. Hirano and C. Shen,Hadron spectra and elliptic flow for 200 A GeV Au+Aucollisions from viscous hydrodynamics coupled to aBoltzmann cascade, Phys. Rev. C83 (2011) 054910[1101.4638].

[19] H. Song, S. A. Bass and U. Heinz, Elliptic flow in 200 AGeV Au+Au collisions and 2.76 A TeV Pb+Pbcollisions: insights from viscous hydrodynamics +hadron cascade hybrid model, Phys. Rev. C83 (2011)054912 [1103.2380].

[20] C. Shen, U. Heinz, P. Huovinen and H. Song, Radialand elliptic flow in Pb+Pb collisions at the LargeHadron Collider from viscous hydrodynamic, Phys. Rev.C84 (2011) 044903 [1105.3226].

[21] B. Schenke, S. Jeon and C. Gale, Higher flow harmonicsfrom (3+1)D event-by-event viscous hydrodynamics,Phys. Rev. C85 (2012) 024901 [1109.6289].

Page 15: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

15

[22] C. Gale, S. Jeon, B. Schenke, P. Tribedy andR. Venugopalan, Event-by-event anisotropic flow inheavy-ion collisions from combined Yang-Mills andviscous fluid dynamics, Phys. Rev. Lett. 110 (2013)012302 [1209.6330].

[23] P. Bozek, A. Bzdak and G.-L. Ma, Rapidity dependenceof elliptic and triangular flow in protonnucleuscollisions from collective dynamics, Phys. Lett. B748(2015) 301 [1503.03655].

[24] P. Bozek, Small systems hydrodynamics, Nucl. Phys.A956 (2016) 208.

[25] M. P. Heller, R. A. Janik and P. Witaszczyk,Hydrodynamic gradient expansion in gauge theoryplasmas, Phys. Rev. Lett. 110 (2013) 211602[1302.0697].

[26] M. P. Heller and M. Spaliski, Hydrodynamics beyond thegradient expansion: resurgence and resummation, Phys.Rev. Lett. 115 (2015) 072501 [1503.07514].

[27] I. Aniceto and M. Spaliski, Resurgence in extendedhydrodynamics, Phys. Rev. D93 (2016) 085008[1511.06358].

[28] G. Basar and G. V. Dunne, Hydrodynamics, resurgence,and transasymptotics, Phys. Rev. D92 (2015) 125011[1509.05046].

[29] W. Florkowski, R. Ryblewski and M. Spaliski, Gradientexpansion for anisotropic hydrodynamics, Phys. Rev.D94 (2016) 114025 [1608.07558].

[30] M. P. Heller, A. Kurkela, M. Spaliski and V. Svensson,Hydrodynamization in kinetic theory: Transient modesand the gradient expansion, Phys. Rev. D97 (2018)091503 [1609.04803].

[31] P. Romatschke, Do nuclear collisions create a locallyequilibrated quark-gluon plasma?, Eur. Phys. J. C77(2017) 21 [1609.02820].

[32] A. Behtash, C. N. Cruz-Camacho and M. Martinez,Far-from-equilibrium attractors and nonlineardynamical systems approach to the Gubser flow, Phys.Rev. D97 (2018) 044041 [1711.01745].

[33] J.-P. Blaizot and L. Yan, Onset of hydrodynamics for aquark-gluon plasma from the evolution of moments ofdistribution functions, JHEP 11 (2017) 161[1703.10694].

[34] J.-P. Blaizot and L. Yan, Fluid dynamics of out ofequilibrium boost invariant plasmas, Phys. Lett. B780(2018) 283 [1712.03856].

[35] G. S. Denicol and J. Noronha, Analytical attractor andthe divergence of the slow-roll expansion in relativistichydrodynamics, Phys. Rev. D97 (2018) 056021[1711.01657].

[36] P. Romatschke, Relativistic hydrodynamic attractorswith broken symmetries: non-conformal andnon-homogeneous, JHEP 12 (2017) 079 [1710.03234].

[37] P. Romatschke, Relativistic fluid dynamics far fromlocal equilibrium, Phys. Rev. Lett. 120 (2018) 012301[1704.08699].

[38] M. Strickland, J. Noronha and G. Denicol, Anisotropicnonequilibrium hydrodynamic attractor, Phys. Rev. D97(2018) 036020 [1709.06644].

[39] M. Spaliski, On the hydrodynamic attractor ofYang-Mills plasma, Phys. Lett. B776 (2018) 468[1708.01921].

[40] I. Aniceto, B. Meiring, J. Jankowski and M. Spaliski,The large proper-time expansion of Yang-Mills plasmaas a resurgent transseries, JHEP 02 (2019) 073

[1810.07130].[41] A. Behtash, S. Kamata, M. Martinez and C. N.

Cruz-Camacho, Non-perturbative rheological behavior ofa far-from-equilibrium expanding plasma, 1805.07881.

[42] C. Chattopadhyay, U. Heinz, S. Pal and G. Vujanovic,Higher order and anisotropic hydrodynamics forBjorken and Gubser flows, Phys. Rev. C97 (2018)064909 [1801.07755].

[43] G. S. Denicol and J. Noronha, Hydrodynamic attractorand the fate of perturbative expansions in Gubser flow,Phys. Rev. D99 (2019) 116004 [1804.04771].

[44] M. P. Heller and V. Svensson, How does relativistickinetic theory remember about initial conditions?, Phys.Rev. D98 (2018) 054016 [1802.08225].

[45] M. Strickland, The non-equilibrium attractor for kinetictheory in relaxation time approximation, JHEP 12(2018) 128 [1809.01200].

[46] L. Tinti, G. Vujanovic, J. Noronha and U. Heinz,Resummed hydrodynamic expansion for a plasma ofparticles interacting with fields, Phys. Rev. D99 (2019)016009 [1808.06436].

[47] A. Behtash, S. Kamata, M. Martinez and H. Shi,Dynamical systems and nonlinear transient rheology ofthe far-from-equilibrium Bjorken flow, Phys. Rev. D99(2019) 116012 [1901.08632].

[48] J.-P. Blaizot and L. Yan, Emergence of hydrodynamicalbehavior in expanding quark-gluon plasmas, 1904.08677.

[49] S. Grozdanov, P. K. Kovtun, A. O. Starinets andP. Tadi, The complex life of hydrodynamic modes,1904.12862.

[50] S. Grozdanov, P. K. Kovtun, A. O. Starinets andP. Tadi, Convergence of the Gradient Expansion inHydrodynamics, Phys. Rev. Lett. 122 (2019) 251601[1904.01018].

[51] U. W. Heinz and J. S. Moreland, Hydrodynamic flow insmall systems or: “How the heck is it possible that asystem emitting only a dozen particles can be describedby fluid dynamics?, J. Phys. Conf. Ser. 1271 (2019)012018 [1904.06592].

[52] M. Strickland and U. Tantary, Exact solution for thenon-equilibrium attractor in number-conservingrelaxation time approximation, 1903.03145.

[53] J. D. Bjorken, Highly relativistic nucleus-nucleuscollisions: the central rapidity region, Phys. Rev. D27(1983) 140.

[54] C. Eckart, The thermodynamics of irreversibleprocesses. 1. The simple fluid, Phys. Rev. 58 (1940) 267.

[55] L. D. Landau and E. M. Lifshitz, Fluid mechanics, 2nded., Course of Theoretical Physics Volume 6,Butterworth-Heinemann, Oxford (1987) .

[56] W. A. Hiscock and L. Lindblom, Stability and causalityin dissipative relativistic fluids, Annals Phys. 151(1983) 466.

[57] W. A. Hiscock and L. Lindblom, Generic instabilities infirst-order dissipative relativistic fluid theories, Phys.Rev. D31 (1985) 725.

[58] I. Muller, Zum Paradoxon der Warmeleitungstheorie, Z.Phys. 198 (1967) 329.

[59] W. Israel, Nonstationary irreversible thermodynamics:A Causal relativistic theory, Annals Phys. 100 (1976)310.

[60] W. Israel and J. M. Stewart, Transient relativisticthermodynamics and kinetic theory, Annals Phys. 118(1979) 341.

Page 16: PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f - arxiv.org · PACS numbers: 25.75.-q, 24.10.Nz, 47.75+f I. INTRODUCTION Hydrodynamics is an e ective macroscopic theory de-scribing the

16

[61] M. P. Heller, R. A. Janik, M. Spaliski andP. Witaszczyk, Coupling hydrodynamics tononequilibrium degrees of freedom in strongly interactingquark-gluon plasma, Phys. Rev. Lett. 113 (2014) 261601[1409.5087].

[62] A. Kurkela and Y. Zhu, Isotropization andhydrodynamization in weakly coupled heavy-ioncollisions, Phys. Rev. Lett. 115 (2015) 182301[1506.06647].

[63] M. P. Heller, Holography, hydrodynamization andheavy-ion collisions, Acta Phys. Polon. B47 (2016) 2581[1610.02023].

[64] S. S. Gubser, Symmetry constraints on generalizationsof Bjorken flow, Phys. Rev. D82 (2010) 085027[1006.0006].

[65] M. Strickland, The non-equilibrium attractor: Beyondhydrodynamics, Acta Phys. Polon. B50 (2019) 1243[1904.00413].

[66] C. P. H. Goldstein and J. Safko, Classical mechanics,3rd ed., Addison-Wesley (1980) .

[67] M. Martinez and M. Strickland, Dissipative dynamics ofhighly anisotropic systems, Nucl. Phys. A848 (2010)183 [1007.0889].

[68] W. Florkowski and R. Ryblewski, Highly-anisotropicand strongly-dissipative hydrodynamics for early stagesof relativistic heavy-ion collisions, Phys. Rev. C83(2011) 034907 [1007.0130].

[69] G. S. Denicol, H. Niemi, E. Molnar and D. H. Rischke,Derivation of transient relativistic fluid dynamics fromthe Boltzmann equation, Phys. Rev. D85 (2012) 114047[1202.4551].

[70] M. Strickland, Anisotropic hydrodynamics: Threelectures, Acta Phys. Polon. B45 (2014) 2355[1410.5786].

[71] D. Bazow, U. Heinz and M. Strickland, Second-order(2+1)-dimensional anisotropic hydrodynamics, Phys.Rev. C90 (2014) 054910 [1311.6720].

[72] A. Jaiswal, Relativistic third-order dissipative fluiddynamics from kinetic theory, Phys. Rev. C88 (2013)021903 [1305.3480].

[73] A. Jaiswal, R. Ryblewski and M. Strickland, Transportcoefficients for bulk viscous evolution in the relaxationtime approximation, Phys. Rev. C90 (2014) 044908[1407.7231].

[74] L. Tinti, Anisotropic matching principle for thehydrodynamic expansion, Phys. Rev. C94 (2016) 044902[1506.07164].

[75] E. Molnar, H. Niemi and D. H. Rischke, Derivation ofanisotropic dissipative fluid dynamics from theBoltzmann equation, Phys. Rev. D93 (2016) 114025[1602.00573].

[76] E. Molnar, H. Niemi and D. H. Rischke, Closing theequations of motion of anisotropic fluid dynamics by ajudicious choice of a moment of the Boltzmannequation, Phys. Rev. D94 (2016) 125003 [1606.09019].

[77] M. McNelis, D. Bazow and U. Heinz, (3+1)-dimensionalanisotropic fluid dynamics with a lattice QCD equationof state, Phys. Rev. C97 (2018) 054912 [1803.01810].

[78] M. Nopoush and M. Strickland, Including off-diagonalanisotropies in anisotropic hydrodynamics, Phys. Rev.C100 (2019) 014904 [1902.03303].

[79] W. Florkowski, R. Ryblewski and M. Strickland,Testing viscous and anisotropic hydrodynamics in anexactly solvable case, Phys. Rev. C88 (2013) 024903[1305.7234].

[80] M. Martinez, M. McNelis and U. Heinz, Anisotropicfluid dynamics for Gubser flow, Phys. Rev. C95 (2017)054907 [1703.10955].

[81] C. Chattopadhyay, A. Jaiswal, S. Jaiswal and S. Pal,Analytical solutions of causal relativistic hydrodynamicequations for Bjorken and Gubser flows, Nucl. Phys.A982 (2019) 911 [1807.05544].

[82] Y. Hatta, J. Noronha and B.-W. Xiao, Exact analyticalsolutions of second-order conformal hydrodynamics,Phys. Rev. D89 (2014) 051702 [1401.6248].

[83] R. Baier, P. Romatschke and U. A. Wiedemann,Dissipative hydrodynamics and heavy ion collisions,Phys. Rev. C73 (2006) 064903 [hep-ph/0602249].

[84] R. Baier, P. Romatschke, D. T. Son, A. O. Starinetsand M. A. Stephanov, Relativistic viscoushydrodynamics, conformal invariance, and holography,JHEP 04 (2008) 100 [0712.2451].

[85] A. Jaiswal, Relativistic dissipative hydrodynamics fromkinetic theory with relaxation time approximation, Phys.Rev. C87 (2013) 051901 [1302.6311].

[86] S. Grozdanov and N. Kaplis, Constructing higher-orderhydrodynamics: The third order, Phys. Rev. D93 (2016)066012 [1507.02461].

[87] H. Kleinert and V. Schulte-Frohlinde, Critical propertiesof φ4-theories. World Scientific Publishing Co., RiverEdge, USA, 2001.

[88] H. Kouno, M. Maruyama, F. Takagi and K. Saito,Relativistic hydrodynamics of quark-gluon plasma andstability of scaling solutions, Phys. Rev. D41 (1990)2903.

[89] A. Muronga, Second order dissipative fluid dynamics forultrarelativistic nuclear collisions, Phys. Rev. Lett. 88(2002) 062302 [nucl-th/0104064].

[90] A. Kurkela, U. A. Wiedemann and B. Wu, Whatattracts to attractors?, 1907.08101v1.