paradox - arxiv.org e-print archive · paradox: venkat balasubramanian, 1, alex buchel, 1,2, y...

8
Holographic Thermalization, stability of AdS, and the Fermi-Pasta-Ulam-Tsingou paradox Venkat Balasubramanian, 1, * Alex Buchel, 1, 2, Stephen R. Green, 3, Luis Lehner, 2, § and Steven L. Liebling 4, 1 Department of Applied Mathematics, University of Western Ontario, London, Ontario N6A 5B7, Canada 2 Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada 3 Department of Physics, University of Guelph, Guelph, Ontario N1G 2W1, Canada 4 Department of Physics, Long Island University, Brookville, NY 11548, U.S.A For a real massless scalar field in general relativity with a negative cosmological constant, we uncover a large class of spherically symmetric initial conditions that are close to AdS, but whose numerical evolution does not result in black hole formation. According to the AdS/CFT dictionary, these bulk solutions are dual to states of a strongly interacting boundary CFT that fail to ther- malize at late times. Furthermore, as these states are not stationary, they define dynamical CFT configurations that do not equilibrate. We develop a two-timescale perturbative formalism that captures both direct and inverse cascades of energy and agrees with our fully nonlinear evolutions in the appropriate regime. We also show that this formalism admits a large class of quasi-periodic solutions. Finally, we demonstrate a striking parallel between the dynamics of AdS and the classic Fermi-Pasta-Ulam-Tsingou problem. Introduction.—The gauge theory-string theory corre- spondence [1] has become a valuable tool to study nonequilibrium phenomena in strongly interacting QFTs [2–4]. In a particular limit, this correspondence links gen- eral relativity in d +1-dimensional asymptotically anti-de Sitter (AdS d+1 ) spacetimes with d-dimensional confor- mal field theories. A question of particular importance in field theory is to understand the process of equilibra- tion and thermalization. This corresponds, in the bulk, to collapse of an initial perturbation to a black hole. In the first detailed analysis [5] of dynamics of pertur- bations of global AdS 4 , Bizo´ n and Rostworowski argued that (except for special nonresonant initial data) the evo- lution of a real, massless, spherically symmetric scalar field always results in gravitational collapse, even for arbi- trarily small initial field amplitude . At the linear level, this system is characterized by a normal mode spectrum with natural frequencies ω j =2j + 3. Using weakly non- linear perturbation theory, these authors described the onset of instability as a result of resonant interactions between the normal modes. Because of the presence of a vast number of resonances, they argued that this mecha- nism leads to a direct turbulent cascade of energy to high mode numbers, making gravitational collapse inevitable. Higher mode numbers are more sharply peaked, so this corresponds to an effect of gravitational focusing. The analysis of [5] also showed that, for initial data consisting of a single mode, the dominant effect of res- onant self-interaction could be absorbed into a constant shift in the frequency of the mode. (This time-periodic solution was confirmed to persist at higher nonlinear or- der [6].) However, for two-mode initial data, additional resonances are present that cannot be absorbed into fre- quency shifts. The result is secular growth of higher modes. The turbulent cascade described in [5] is a beautiful mechanism for thermalization of strongly coupled QFTs with holographic gravitational duals. However, it was recently pointed out that this cascade argument breaks down if all modes are initially populated, and the mode amplitudes fall off sufficiently rapidly for high mode num- bers [7]. In this case, all resonant effects may once again be absorbed into frequency shifts and black hole collapse is avoided. Low-lying modes have broadly distributed bulk profiles. Thus, one might expect that if the ini- tial scalar profile is broadly distributed, its evolution might not result in gravitational collapse (see also [8– 10]). This prediction was verified numerically [11]. The physical mechanism responsible for collapse/non-collapse of small amplitude initial data is a competition between two effects: gravitational focusing and nonlinear dis- persion of the propagating scalar field. If the former dominates, gravitational collapse ensues [5]. If the lat- ter does, the system evolves without approaching any identifiable static or stationary solution—the perturbed boundary CFT neither thermalizes nor equilibrates at late times [11]. The perturbation theory of [5] cannot make predic- tions at late times. (The growth of secular terms in the expansion causes a breakdown at time t 1/ 2 .) It also does not properly take into account energy transfer be- tween modes. In this Letter, we undertake a thorough analysis of the dynamics of AdS by making use of a new perturbative formalism for analyzing the effect of reso- nances on the evolution of this system that is valid for long times. We also perform fully nonlinear GR sim- ulations (see [7, 11] for details of our numerical imple- mentation and validation). In the process we uncover a close relationship between the dynamics of AdS and the famous Fermi-Pasta-Ulam-Tsingou (FPUT) prob- lem [12, 13]. Our formalism is based on a two-timescale approach [14], where we introduce a new “slow time” τ = 2 t. The timescale τ characterizes energy transfers between modes, whereas the “fast time” t characterizes arXiv:1403.6471v2 [hep-th] 16 Jun 2014

Upload: others

Post on 20-Sep-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

Holographic Thermalization, stability of AdS, and the Fermi-Pasta-Ulam-Tsingouparadox

Venkat Balasubramanian,1, ∗ Alex Buchel,1, 2, † Stephen R. Green,3, ‡ Luis Lehner,2, § and Steven L. Liebling4, ¶

1Department of Applied Mathematics, University of Western Ontario, London, Ontario N6A 5B7, Canada2Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada3Department of Physics, University of Guelph, Guelph, Ontario N1G 2W1, Canada

4Department of Physics, Long Island University, Brookville, NY 11548, U.S.A

For a real massless scalar field in general relativity with a negative cosmological constant, weuncover a large class of spherically symmetric initial conditions that are close to AdS, but whosenumerical evolution does not result in black hole formation. According to the AdS/CFT dictionary,these bulk solutions are dual to states of a strongly interacting boundary CFT that fail to ther-malize at late times. Furthermore, as these states are not stationary, they define dynamical CFTconfigurations that do not equilibrate. We develop a two-timescale perturbative formalism thatcaptures both direct and inverse cascades of energy and agrees with our fully nonlinear evolutionsin the appropriate regime. We also show that this formalism admits a large class of quasi-periodicsolutions. Finally, we demonstrate a striking parallel between the dynamics of AdS and the classicFermi-Pasta-Ulam-Tsingou problem.

Introduction.—The gauge theory-string theory corre-spondence [1] has become a valuable tool to studynonequilibrium phenomena in strongly interacting QFTs[2–4]. In a particular limit, this correspondence links gen-eral relativity in d+1-dimensional asymptotically anti-deSitter (AdSd+1) spacetimes with d-dimensional confor-mal field theories. A question of particular importancein field theory is to understand the process of equilibra-tion and thermalization. This corresponds, in the bulk,to collapse of an initial perturbation to a black hole.

In the first detailed analysis [5] of dynamics of pertur-bations of global AdS4, Bizon and Rostworowski arguedthat (except for special nonresonant initial data) the evo-lution of a real, massless, spherically symmetric scalarfield always results in gravitational collapse, even for arbi-trarily small initial field amplitude ε. At the linear level,this system is characterized by a normal mode spectrumwith natural frequencies ωj = 2j + 3. Using weakly non-linear perturbation theory, these authors described theonset of instability as a result of resonant interactionsbetween the normal modes. Because of the presence of avast number of resonances, they argued that this mecha-nism leads to a direct turbulent cascade of energy to highmode numbers, making gravitational collapse inevitable.Higher mode numbers are more sharply peaked, so thiscorresponds to an effect of gravitational focusing.

The analysis of [5] also showed that, for initial dataconsisting of a single mode, the dominant effect of res-onant self-interaction could be absorbed into a constantshift in the frequency of the mode. (This time-periodicsolution was confirmed to persist at higher nonlinear or-der [6].) However, for two-mode initial data, additionalresonances are present that cannot be absorbed into fre-quency shifts. The result is secular growth of highermodes.

The turbulent cascade described in [5] is a beautifulmechanism for thermalization of strongly coupled QFTs

with holographic gravitational duals. However, it wasrecently pointed out that this cascade argument breaksdown if all modes are initially populated, and the modeamplitudes fall off sufficiently rapidly for high mode num-bers [7]. In this case, all resonant effects may once againbe absorbed into frequency shifts and black hole collapseis avoided. Low-lying modes have broadly distributedbulk profiles. Thus, one might expect that if the ini-tial scalar profile is broadly distributed, its evolutionmight not result in gravitational collapse (see also [8–10]). This prediction was verified numerically [11]. Thephysical mechanism responsible for collapse/non-collapseof small amplitude initial data is a competition betweentwo effects: gravitational focusing and nonlinear dis-persion of the propagating scalar field. If the formerdominates, gravitational collapse ensues [5]. If the lat-ter does, the system evolves without approaching anyidentifiable static or stationary solution—the perturbedboundary CFT neither thermalizes nor equilibrates atlate times [11].

The perturbation theory of [5] cannot make predic-tions at late times. (The growth of secular terms in theexpansion causes a breakdown at time t ∝ 1/ε2.) It alsodoes not properly take into account energy transfer be-tween modes. In this Letter, we undertake a thoroughanalysis of the dynamics of AdS by making use of a newperturbative formalism for analyzing the effect of reso-nances on the evolution of this system that is valid forlong times. We also perform fully nonlinear GR sim-ulations (see [7, 11] for details of our numerical imple-mentation and validation). In the process we uncovera close relationship between the dynamics of AdS andthe famous Fermi-Pasta-Ulam-Tsingou (FPUT) prob-lem [12, 13]. Our formalism is based on a two-timescaleapproach [14], where we introduce a new “slow time”τ = ε2t. The timescale τ characterizes energy transfersbetween modes, whereas the “fast time” t characterizes

arX

iv:1

403.

6471

v2 [

hep-

th]

16

Jun

2014

Page 2: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

2

the original normal modes. Importantly, this formalismallows one to study the system for long times and exam-ine energy transfer between modes. In the following wedescribe the Two Time Framework (TTF) and determinea large class of quasi-periodic solutions that extends thesingle-mode periodic solutions of [5, 6]. These solutionshave finely tuned energy spectra such that the net en-ergy flow into each mode vanishes, and they appear tobe stable to small perturbations within both TTF andfull numerical simulations. We then study the behaviorof two-mode initial data of [5] under both approaches. Fi-nally, we use the TTF equations to draw an interestingparallel between scalar collapse in AdS and the FPUTproblem of thermalization of nonlinearly coupled oscilla-tors [12].Model.—Following [5], we consider a self-gravitating, realscalar field φ in asymptotically AdS4 spacetime. Impos-ing spherical symmetry, the metric takes the form

ds2 =1

cos2 x

(−Ae−2δdt2 +A−1dx2 + sin2 x dΩ2

), (1)

where we set the asymptotic AdS radius to one. Sphericalsymmetry implies that A, δ and φ are functions of timet ∈ (−∞,∞) and the radial coordinate x ∈ [0, π2 ).

In terms of the variables Π ≡ eδφ/A and Φ ≡ φ′, theequation of motion for φ is

φ =(Ae−δ −Aδe−δ

)Π +A2e−2δΦ′ (2)

+

(2

sinx cosxA2e−2δ +AA′e−2δ −A2e−2δδ′

)Φ,

while the Einstein equation reduces to the constraints,

A′ =1 + 2 sin2 x

sinx cosx(1−A) + sinx cosxA

(|Φ|2 + |Π|2

),

(3)

δ′ = − sinx cosx(|Φ|2 + |Π|2

). (4)

Two Time Framework.—TTF consists of defining theslow time τ = ε2t and expanding the fields as

φ = εφ(1)(t, τ, x) + ε3φ(3)(t, τ, x) +O(ε5), (5)

A = 1 + ε2A(2)(t, τ, x) +O(ε4), (6)

δ = ε2δ(2)(t, τ, x) +O(ε4). (7)

It is possible to go beyond O(ε3) by introducing addi-tional slow time variables. However, the order of approx-imation used here is sufficient to capture the key aspectsof weakly nonlinear AdS collapse in the ε→ 0 limit.

Perturbative equations are derived by substituting theexpansions (5)–(7) into the equations of motion (2)–(4),and equating powers of ε. It is important to note that,when taking time derivatives of a function of both timevariables we have ∂t → ∂t+ ε2∂τ . At O(ε), we obtain thewave equation for φ(1) linearized off exact AdS,

∂2t φ(1) = φ′′(1) +2

sinx cosxφ′(1) ≡ −Lφ(1). (8)

The operator L has eigenvalues ω2j = (2j + 3)2 (j =

0, 1, 2, . . .) and eigenvectors ej(x) (“oscillons”) [5]. Ex-plicitly,

ej(x) = dj cos3 x 2F1

(−j, 3 + j;

3

2; sin2 x

), (9)

with dj = 4√

(j + 1)(j + 2)/√π. The oscillons form an

orthonormal basis under the inner product

(f, g) =

∫ π/2

0

f(x)g(x) tan2 xdx. (10)

The general real solution to (8) is

φ(1)(t, τ, x) =

∞∑j=0

(Aj(τ)e−iωjt + Aj(τ)eiωjt

)ej(x),

(11)where Aj(τ) are arbitrary functions of τ , to be deter-mined later.

At O(ε2) the constraints (3)–(4) have solutions

A(2)(x) = −cos3 x

sinx

∫ x

0

(|Φ(1)(y)|2 + |Π(1)(y)|2

)tan2 y dy,

(12)

δ(2)(x) = −∫ x

0

(|Φ(1)(y)|2 + |Π(1)(y)|2

)sin y cos y dy.

(13)

Finally, at O(ε3) we obtain the equation for φ(3),

∂2t φ(3) + Lφ(3) + 2∂t∂τφ(1) = S(3)(t, τ, x), (14)

where the source term is

S(3) = ∂t(A(2) − δ(2))∂tφ(1) − 2(A(2) − δ(2))Lφ(1)+ (A′(2) − δ′2)φ′(1). (15)

The solutions (12)–(13) for A(2) and δ(2) are substituteddirectly into S(3). In general, the source term S(3) con-tains resonant terms (i.e., terms proportional to e±iωjt).As noted in [5], for all triads (j1, j2, j3), resonances occurat ωj = ωj1 +ωj2 −ωj3 . In ordinary perturbation theorythese resonances lead to secular growths in φ(3). How-ever, [5] showed that in some cases the growths may beabsorbed into frequency shifts. TTF provides a naturalway to handle these resonances by taking advantage ofthe new term 2∂t∂τφ(1) in (14) and the freedom in Aj(τ).

We now project (14) onto an individual oscillon modeej and substitute for φ(1),(

ej , ∂2t φ(3) + ω2

jφ(3))− 2iωj

(∂τAje

−iωjt − ∂τ Ajeiωjt)

=(ej , S(3)

). (16)

By exploiting the presence of terms proportional to e±iωjt

on the left hand side of the equation, we may cancel off

Page 3: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

3

the resonant terms on the right hand side. Denoting byf [ωj ] the part of f proportional to eiωjt, we set

−2iωj∂τAj = (ej , S(t, τ, x))[−ωi] =∑klm

S(j)klmAkAlAm,

(17)

where S(j)klm are real constants representing different res-onance channel contributions. The right hand side is acubic polynomial in Aj and Aj . Thus, we have obtaineda set of coupled first order ODEs in τ for Aj , which weshall refer to as the TTF equations. The equations are tobe solved given the initial conditions for φ. This proce-dure fixes the arbitrariness in the solution (11) for φ(1).While we could also solve for φ(3), this would be of lit-tle interest since the lack of resonances remaining in (14)implies that φ(3) remains bounded.

Under evolution via the TTF equations, both the am-plitude and phase of the complex coefficients Aj(τ) canvary. Thus, in contrast to the perturbative analysis in [5],the energy per mode Ej = ω2

j |Aj |2 can change withtime in a very nontrivial manner. However, it can bechecked that the total energy E =

∑j Ej is conserved.

TTF thus describes an energy-conserving dynamical sys-tem. The TTF equations also possess a scaling symme-try Aj(τ)→ εAj(τ/ε

2). This symmetry was observed inFig. 2b of [5], which indicates that the instability mech-anism is captured by TTF.

In practice, it is necessary to truncate the TTF equa-

tions at finite j = jmax. We evaluated S(j)klm up tojmax = 47. In particular, under truncation to jmax = 0,the equations reduce to

iπ∂τA0 = 153A20A0, (18)

with solution A0(τ) = A0(0) exp(−i 153π |A0(0)|2τ

). This

reproduces precisely the single-mode frequency shift re-sult of [5].Quasi-periodic solutions.—To understand the dynamicsof TTF, we first look for quasi-periodic solutions. Forjmax = 0 this is the periodic solution above. For generaljmax > 0 we take as ansatz Aj = αj exp(−iβjτ), whereαj , βj ∈ R are independent of τ . These solutions haveEj = constant, so they represent a balancing of energyfluxes such that each mode has constant energy. Substi-tuting into the TTF equations, the τ -dependence can becanceled by requiring βj = β0 + j(β1 − β0). This leavesjmax + 1 algebraic equations,

− 2ωjαj [β0 + j(β1 − β0)] =∑kmn

S(j)kmnαkαmαn, (19)

for jmax + 3 unknowns (β0, β1, αj). The equations forj = 0, 1 may be used to eliminate (β0, β1), leaving jmax−1equations to be solved for αj—two parameters of un-derdetermination. The scaling symmetry allows for elim-ination of one parameter, so we set αjr = 1 for some fixed

0 ≤ jr < jmax. Taking the remaining free parameter tobe αjr+1 and requiring solutions to be insensitive to thevalue of jmax (i.e., stable to truncation), it is straightfor-ward to construct solutions perturbatively in αjr+1/αjr .We find a single solution for jr = 0 and precisely twootherwise (see Fig. 1).

0 5 10 15 20 25 30

j+1

-60

-40

-20

0

log

10(E

j/Eto

tal)

jr=0

jr=1, A

jr=1, B

jr=2, A

jr=2, B

jr=3, A

jr=3, B

FIG. 1. Energy spectra of quasi-periodic solutions withαjr+1/αjr = 0.1 for jr = 0, 1, 2, 3. Dashed and solid linesdistinguish different branches for jr > 0. The solid branch iswell-approximated by an exponential to each each side of jr.For αjr+1/αjr too large, it becomes difficult to obtain solu-tions to (19), but for jr = 0, we can go up to α1/α0 ≈ 0.42.(Constructed for jmax = 30.)

Stability of quasi-periodic solutions.—Ref. [6] extendedsingle-mode, time-periodic solutions to higher order in εand found these solutions to be stable to perturbations.Similarly, we examine the stability of our extended classof quasi-periodic solutions, both using full numerical rel-ativity simulations and by numerically solving the TTFODEs.

0 200 400 600 800 1000t

10-4

10-3

10-2

Ej

FIG. 2. Energy per mode for 0 ≤ j ≤ 9 for TTF solutionwith initial data Aj(0) ∝ exp(−0.3j)/(2j + 3).

We consider initial data Aj(0) = ε exp(−µj)/(2j + 3),which well-approximates jr = 0 quasi-periodic solutions.Varying µ and also adding random perturbations, we ob-serve periodic oscillations about the quasi-periodic so-lution, providing evidence for stability (see Fig. 2). Forsmaller values of µ, energy levels are more closely spaced,resulting in more rapid energy transfers between modes,leading to larger-amplitude oscillations. Likewise, largerrandom perturbations increase the amplitude of oscilla-tion, as the initial data deviates more strongly from aquasi-periodic solution. Results from TTF and full nu-merical relativity simulations are in close agreement.

Page 4: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

4

Two-mode initial data.—Our main interest is to under-stand which initial conditions can be expected to col-lapse. Thus it is necessary to study initial data thatare not expected to closely approximate a quasi-periodicsolution. A particularly interesting case consists of twomodes initially excited (all others zero) as this case waskey to the argument of [5] showing the onset of the tur-bulent cascade. In contrast to results of the previoussection, two-mode initial data,

Aj(0) =ε

3

(δ0j + κδ1j

), (20)

involves considerable energy transfer among modes pro-vided κ is sufficiently large. [For κ 1, (20) may beconsidered as a perturbation about single-mode data.]We examined several choices of κ using both TTF andfull numerical relativity, with similar results. Here werestrict to κ = 3/5—the equal-energy case.

0 500 1000 1500t

0

1

2

3

4

5

log

10Π

2(x

=0)

Full Numerical GRTTF, j

max=15

TTF, jmax

=23TTF, j

max=31

TTF, jmax

=47

FIG. 3. Full numerical and TTF results for 2-mode equal-energy initial data with ε = 0.09. As jmax is increased, theTTF solutions achieve better agreement with the full numer-ics. Recurrence behavior observed in the full numerical solu-tion is reasonably well captured by TTF.

The upper envelope of Π2(x = 0) is often used as anindicator of the onset of instability [5, 7, 11]. We plot thisquantity in Fig. 3, both for full GR simulations and TTFsolutions with varying jmax. In the full GR simulation,Π2(x = 0) grows initially, but, in contrast to blowupobserved in [5] for Gaussian scalar field profile, it thendecreases close to its initial value. This recurrence phe-nomenon repeats and—for sufficiently small ε—collapsenever occurs for as long as we have run the simulation.Recurrence was also observed in previous work [11] forbroadly distributed Gaussian profiles.

Also in Fig. 3, TTF solutions appear to converge to thefull numerical GR solution as jmax is increased. (Strictlyspeaking, the TTF and numerical approaches converge asboth jmax → ∞ and ε → 0; see the accompanying sup-plemental material for more discussion.) This illustratesnicely the cascade/collapse mechanism: Higher-j modesare more sharply peaked at x = 0, so as the (conserved)energy is transferred to these modes, Π2(x = 0) attainshigher values. Truncating the system at finite jmax arti-ficially places a bound on values of Π2(x = 0) that canbe reached. In particular, Π2(x = 0) can never blow upfor jmax <∞.

0

0.005

0.01

0.015

Ej

j=0j=1j=2j=3j=4j=5

0 500 1000 1500t

10-5

10-4

10-3

10-2

Ej

FIG. 4. Full numerical (solid) and TTF (dotted) energy (toppanel) and running time-average energy (bottom panel) permode, for 2-mode equal-energy initial data. Notice the re-peated approximate return of the initial energies to the firsttwo modes in the top panel and the running time-averageenergies asymptoting to distinct values. For this run withjmax = 47,

∑11j=0 |ETTF

j − Enumericalj |/Etotal does not exceed

0.19. For jmax = (31, 23, 15) the bounds are (0.28, 0.42, 0.57).(The horizontal offset is partially attributed to a slight differ-ence in time-normalization for our numerical and TTF codes.)

It is useful to examine the solution mode by mode,and in Fig. 4 we show the energy per mode as a functionof time. Initially, energy is distributed evenly betweenmodes j = 0, 1. It then flows out of j = 1 to mode j = 2,then j = 3, etc. At some point in time, energy beginsto flow back to mode j = 1—an inverse energy cascade.By t ≈ 450 the state has nearly returned to the originalconfiguration. This recurrence behavior then repeats.

The bottom plot of Fig. 4 illustrates the runningtime-average energy per mode Ej(t) ≡ t−1

∫ t0Ej(t

′) dt′.Rather than cascading to ever-higher modes, the energysloshes primarily between low-j modes, in a “metastable”state. We never observe thermalization, i.e., no equipar-tition of energy occurs.

Fig. 4 is remarkably similar in appearance to plots ofFPUT [12] (cf. Figs. 4.1 and 4.2 of [15].) FPUT nu-merically simulated a collection of nonlinearly coupledharmonic oscillators and expected to see thermalization.Instead, they observed the same recurrence we see here.Indeed, as the TTF formulation (17) of our system makesclear, small-amplitude scalar collapse in AdS reduces pre-cisely to a (infinite) set of nonlinearly coupled oscillators,so the similar behavior should not be surprising. Moreprecisely, our system is related to the FPUT β-model[15]. (Of course, the particular resonances and nonlinearinteractions differ between our system and FPUT.) Pre-dicting when the FPUT system of oscillators thermalizesis a longstanding problem in nonlinear dynamics, and is

Page 5: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

indeed known as the FPUT paradox [15–17].

Discussion.—Common intuition suggests that a finite-sized strongly interacting system driven off-equilibrium,even by a small amount, eventually thermalizes. Thisthermalization would imply, via AdS/CFT, that arbi-trarily small perturbations about global AdS must re-sult in gravitational collapse. However, we have uncov-ered in this Letter a large class of initial conditions fora massless, self-gravitating real scalar field in AdS4, thatfail to collapse. We constructed and evolved these ini-tial conditions within a newly proposed TTF, as well asthrough full numerical GR simulations. TTF shows thatscalar perturbations of AdS are in the same universalityclass as the famous FPUT problem [12]. Thus, perturbedAdS spacetimes act as a holographic bridge between non-equilibrium dynamics of CFTs and the dynamics of non-linearly coupled oscillators and the FPUT paradox. Inthis Letter we focused on the dynamics of low-energy 2+1dimensional CFT excitations “prepared” with nonzeroexpectation values of dimension three (marginal) opera-tors. Extensions to higher-dimensional CFTs, as well asto states generated by (ir)relevant operators are straight-forward.

Acknowledgments: We would like to thank A.Polkovnikov and J. Santos for interesting discussions andcorrespondence. We also thank D. Minic and A. Zim-merman for pointing us to the FPUT problem. Finally,we thank P. Bizon and A. Rostworowski for discussionsand direct comparison of numerical solutions. This workwas supported by the NSF under grants PHY-0969827 &PHY-1308621 (LIU), NASA under grant NNX13AH01G,NSERC through a Discovery Grant (to A.B. and L.L.)and CIFAR (to L.L.). S.R.G. acknowledges support bya CITA National Fellowship. L.L. and S.L.L. also ac-knowledge the Centre for Theoretical Cosmology at theUniversity of Cambridge for their hospitality and its par-ticipants for helpful discussions at the “New Frontiersin Dynamical Gravity” meeting during which this workwas completed. Research at Perimeter Institute is sup-ported through Industry Canada and by the Province ofOntario through the Ministry of Research & Innovation.Computations were performed at Sharcnet.

[email protected][email protected][email protected]; CITA National Fellow§ [email protected][email protected]

[1] O. Aharony, S. S. Gubser, J. M. Maldacena, H. Ooguri,and Y. Oz, Phys. Rept. 323, 183 (2000), arXiv:hep-th/9905111 [hep-th].

[2] P. M. Chesler and L. G. Yaffe, Phys. Rev. Lett. 102,211601 (2009), arXiv:0812.2053 [hep-th].

[3] V. Balasubramanian, A. Bernamonti, J. de Boer, N. Cop-

land, B. Craps, et al., Phys. Rev. Lett. 106, 191601(2011), arXiv:1012.4753 [hep-th].

[4] A. Buchel, R. C. Myers, and A. van Niekerk,Phys. Rev. Lett. 111, 201602 (2013), arXiv:1307.4740[hep-th].

[5] P. Bizon and A. Rostworowski, Phys. Rev. Lett. 107,031102 (2011), arXiv:1104.3702 [gr-qc].

[6] M. Maliborski and A. Rostworowski, Phys. Rev. Lett.111, 051102 (2013), arXiv:1303.3186 [gr-qc].

[7] A. Buchel, L. Lehner, and S. L. Liebling, Phys. Rev.D86, 123011 (2012), arXiv:1210.0890 [gr-qc].

[8] M. Maliborski and A. Rostworowski, (2014),arXiv:1403.5434 [gr-qc].

[9] O. J. Dias, G. T. Horowitz, D. Marolf, and J. E. Santos,Class. Quant. Grav. 29, 235019 (2012), arXiv:1208.5772[gr-qc].

[10] J. Abajo-Arrastia, E. da Silva, E. Lopez, J. Mas, andA. Serantes, (2014), arXiv:1403.2632 [hep-th].

[11] A. Buchel, S. L. Liebling, and L. Lehner, Phys. Rev.D87, 123006 (2013), arXiv:1304.4166 [gr-qc].

[12] E. Fermi, J. Pasta, and S. M. Ulam, Studies of nonlin-ear problems. I , Technical Report LA-1940 (1955) alsoin Enrico Fermi: Collected Papers, volume 2, edited byEdoardo Amaldi, Herbert L. Anderson, Enrico Persico,Emilio Segre, and Albedo Wattenberg. Chicago: Univer-sity of Chicago Press, 1965.

[13] T. Dauxois, Physics Today 61, 010000 (2008),arXiv:0801.1590 [physics.hist-ph].

[14] J. Kevorkian and J. D. Cole, Multiple scale and singu-lar perturbation methods, Applied mathematical sciences,Vol. 114 (Springer, New York, 1996).

[15] G. Benettin, A. Carati, L. Galgani, and A. Giorgilli, TheFermi-Pasta-Ulam Problem and the Metastability Per-spective, Lecture Notes in Physics, Vol. 728 (Springer,Berlin, 2008).

[16] P. Poggi, S. Ruffo, and H. Kantz, Phys. Rev. E 52, 307(1995).

[17] G. P. Berman and F. M. Izrailev, Chaos 15, 015104(2005), nlin/0411062.

Page 6: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

Holographic Thermalization, stability of AdS, and the Fermi-Pasta-Ulam-Tsingouparadox:

Supplementary Material

Venkat Balasubramanian,1, ∗ Alex Buchel,1, 2, † Stephen R. Green,3, ‡ Luis Lehner,2, § and Steven L. Liebling4, ¶

1Department of Applied Mathematics, University of Western Ontario, London, Ontario N6A 5B7, Canada2Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada3Department of Physics, University of Guelph, Guelph, Ontario N1G 2W1, Canada

4Department of Physics, Long Island University, Brookville, NY 11548, U.S.A

Summary.—We provide additional description of the ap-proximations involved in TTF, as well as expectations forconvergence in both ǫ and jmax to full general relativity.Last, we present results of tests validating our full GRcode.Approximations.—As described in our Letter, TTF in-volves two approximations: small ǫ and truncation inmode number j. By working to O(ǫ3) we neglect interac-tions of higher than cubic order in metric perturbations,but we keep the lowest order nontrivial interactions. Thisgives rise to an infinite set of coupled oscillons, whichin practice must be truncated at a finite mode numberj = jmax. We were able to explicitly determine the equa-tions up to jmax = 47.Small ǫ.—The TTF equations possess a scaling symme-try, Aj(τ) → ǫAj(τ/ǫ

2). The full Einstein equation doesnot possess this symmetry, however, and it is evidentthat had we included higher order in ǫ interactions inTTF, they would break the scaling symmetry. The pres-ence of the scaling symmetry in a set of solutions to thefull Einstein equation therefore indicates that higher or-der in ǫ interactions are negligible for these solutions. InFig. 1 we plot the upper envelope over rapid oscillationsof Π2(x = 0) for full numerical solutions obtained withseveral values of ǫ for 2-mode equal-energy initial data,as well as a plot of rescaled values. It is clear that thesesolutions do possess an approximate scaling symmetryfor sufficiently small ǫ (see also Fig. 2 of [1]).Initial data for Fig. 1 is the same (up to overall normal-

ization) as that of Figs. 3 and 4 of the Letter. Because ofthe approximate scaling symmetry in Fig. 1, the discrep-ancy between TTF and full GR in Fig. 3 of the Letteris therefore primarily attributed to truncation in modenumber j. In addition, as ǫ is decreased, comparison ofFig. 1 and Fig. 3 of the Letter reveals that the small de-gree of non-scaling of the full GR solution brings it intocloser agreement with TTF.Mode number truncation.—For 2-mode equal-energy ini-tial data, under evolution the majority of the energy re-mains in the lowest-j modes (see Fig. 4 of the Letter).Nevertheless, Π2(x = 0) can become very peaked. This isbecause higher-j modes are more sharply peaked aboutthe origin. Even small amounts of energy in high-j modescan lead to growth of Π2(x = 0), despite the dynamics be-ing dominated by low-j modes. This property can lead to

0 500 1000 1500t

0123456

log 10

Π2 (x

=0)

0 5 10

ε2t

2

4

6

8

log 10

ε-2Π

2 (x=

0) ε=0.125ε=0.1ε=0.09

FIG. 1. (Color online) Full numerical GR solutions for 2-modeequal-energy initial data for different values of ǫ. The bottomplot shows the same data, appropriately rescaled in time andamplitude by ǫ. The ǫ = 0.125 case collapses to black holeat t ≈ 550. However, for smaller ǫ black hole formation isavoided with an inverse cascade. As ǫ increases and collapseto black hole ensues, the scaling symmetry in ǫ is broken ashigher order terms (above third order) become important.

collapse in the case where sufficient energy is transferredto high-j modes (e.g., Gaussian initial data in [1]).

We have attempted to show in Fig. 3 of the Letter thatincreasing jmax gives rise to a better approximation of thefull solution. This is best illustrated at early times, beforethe higher modes are populated, in which case the lackof available higher modes in a truncated TTF system isirrelevant. However, once the truncation begins to affectthe dynamics, the overall solution is altered and it ismore difficult to compare with full GR. Nevertheless, itis clear throughout the simulation that for higher jmax,larger values of Π2(x = 0) can be attained.

This discussion and the accompanying figures supportthe assertions that: (i) the TTF and the fully nonlinearsolutions converge as jmax → ∞ and ǫ → 0, and (ii) thelevel to which the nonlinear solutions scale improves withdecreasing ǫ.

The truncated TTF equations are able to accuratelymodel the dynamics of low-j modes for the 2-mode data,

Page 7: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

2

as shown in Fig. 4 of the main paper. However, the exactsolution involves excitations of very high mode numbersat the peaks of Π2(x = 0), and the TTF equations withjmax = 47 are not sufficient to match the fully nonlinearresult. However, we can instead consider a case wherewe expect better agreement. To that end, we include inFig. 2 a plot of the full numerical GR and TTF solu-tions for an initially exponential energy spectrum. Thissolution is (a) close to a quasi-periodic solution, and (b)has very little energy in high-j modes. In particular, theexponential fall-off in the energy spectrum with j over-whelms the polynomial growth in peakiness about x = 0.Thus in this case, the truncated TTF and full numericalGR are in very close agreement1.

0 200 400 600 800 1000t

-0.5

-0.4

-0.3

-0.2

log 10

Π2 (x

=0)

Full Numerical SimulationTTF, j

max=47

FIG. 2. (Color online) Full numerical GR and TTF solutionsfor initial data Aj(0) = 0.05 exp(−j)/(2j +3) which approxi-mates a quasi-periodic solution.

Full numerical GR code validation.—The code we use hasbeen described and thoroughly tested in [2, 3] to studyscalar collapse in AdS. In particular we have confirmed:convergence of the obtained solutions with increased res-olution, mass conservation, and convergence to zero ofthe constraint residuals (similar to Fig. 3 of [2]).We present another, longer example of such a conver-

gence study in Fig. 3. The results unambiguously indi-cate a convergent code over a large number of dynamicaltimescales (roughly 600/π ≈ 190). The test also directlyconfirms that, while a direct cascade dominates at earlytimes, the system exhibits both direct and inverse cas-cades that compete with each other, a central result of

1 The horizontal offset in Fig. 2 (and Figs. 3 and 4 of the Letter)between TTF and full numerical GR is attributed to a smalldifference in time-normalization between our codes. Specifically,the full numerical GR code sets δ = 0 at the AdS boundary,whereas the TTF equations set δ = 0 at x = 0. This gaugefreedom amounts to a difference in time-normalization. The firstnonzero contribution to δ is at O(ǫ2), so this is a small effect.

this Letter; note in the bottom panel that the three high-est resolutions have essentially converged to such a tran-sition at t ≈ 230.This test also helps understand the failure of the TTF

in Fig. 3 of the Letter to resolve the first peak of Π2(0, t).As mentioned previously, the TTF simply cannot resolveenergy in high frequency modes compared to jmax. Eventhe fully nonlinear code must extend to very high resolu-tions to resolve it spatially. Nevertheless, as indicated inFig. 4 of the Letter and in Fig. 2 of this supplement, theTTF captures the essential dynamics of the system andmatches the behavior of the fully nonlinear evolutions forthe low frequency modes.This study tests evolutions with a single grid-spacing,

but in practice we use adaptive mesh refinement (AMR)to greatly reduce computational overhead. With anAMR code such as ours, one can vary a number of pa-rameters to increase the effective resolution, and we oftenincrease both the resolution of the base grid as well thecriterion for refinement. We verify that such changes donot produce different results (see Figs. 5 and 6 of Ref. [3]for past examples of tests of our AMR code). In particu-lar, AMR results are consistent with unigrid results thatshow a series of direct and indirect cascades (see Fig. 4).

10−910−810−710−610−510−410−310−210−1

|∆M

/M0|

x20x21x22x23x24x25

-9-8-7-6-5-4-3-2

Log

10|H

am.R

esid| 2

0 100 200 300 400 500 600t

-0.50.00.51.01.52.02.53.03.5

Log

10|Π

2(0,t)|

FIG. 3. (Color online) Demonstration of convergence for un-igrid evolutions of equal-energy, two-mode initial data withǫ = 0.1. Shown for subsequent doublings of resolution are:(top) total mass loss, (middle) norm of the residual of theHamiltonian constraint, and (bottom) the value of Π2(0, t).Note that the top panels display a measure of the error whichprogressively diminishes with resolution, indicating conver-gence. The bottom panel displays a feature of the evolutionand quickly approaches a unique solution consistent with con-vergence. The base resolution (red, solid line) uses 316 + 1points while the highest resolution (black, dashed line) uses10, 112 + 1 points.

As a final comment we note that in keeping withRef. [4], we refer to the famous FPUT paradox within the

Page 8: paradox - arXiv.org e-Print archive · paradox: Venkat Balasubramanian, 1, Alex Buchel, 1,2, y Stephen R. Green, 3, z Luis Lehner, 2, x and Steven L. Liebling 4, { 1, Alex Buchel,

3

0 50 100 150 200 250t

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0L

og10|Π

2(0,t)|

Unigrid x22AMR x20 (base)AMR x21 (higher)AMR x22 (highest)

FIG. 4. (Color online) Demonstration of convergence forAMR evolutions of equal-energy, two-mode initial data withǫ = 0.1. Shown is the value of Π2(0, t) for three different AMRresolutions as well as the highest unigrid result from Fig. 3.Each successive AMR evolution uses a base grid with twicethe resolution and a smaller threshold for refinement. Thehighest resolution AMR result has the same base resolutionas the unigrid result.

main text by all four names of the contributors, Fermi,Pasta, Ulam, and Tsingou (later Menzel).

[email protected][email protected][email protected]; CITA National Fellow§ [email protected][email protected]

[1] P. Bizon and A. Rostworowski,Phys. Rev. Lett. 107, 031102 (2011),arXiv:1104.3702 [gr-qc].

[2] A. Buchel, L. Lehner, and S. L. Liebling,Phys. Rev. D86, 123011 (2012), arXiv:1210.0890 [gr-qc].

[3] A. Buchel, S. L. Liebling, and L. Lehner,Phys. Rev. D87, 123006 (2013), arXiv:1304.4166 [gr-qc].

[4] T. Dauxois, Physics Today 61, 010000 (2008),arXiv:0801.1590 [physics.hist-ph].