11
Quantum simulations of the early universe B. Opanchuk 1 , R. Polkinghorne 1 , O. Fialko 2 , J. Brand 2 , P. D. Drummond 1* 1 Centre for Atom Optics and Ultrafast Spectroscopy, Swinburne University of Technology, Melbourne 3122, Australia and 2 Institute of Advanced Studies, Massey University, Albany, New Zealand A procedure is described whereby a linearly coupled spinor Bose condensate can be used as a physically accessible quantum simulator of the early universe. In particular, an experiment to gen- erate an analog of an unstable vacuum in a relativistic scalar field theory is proposed. This is related to quantum theories of the inflationary phase of the early universe. There is an unstable vacuum sector whose dynamics correspond to the quantum sine-Gordon equations in one, two or three space dimensions. Numerical simulations of the expected behavior are reported using a truncated Wigner phase-space method, giving evidence for the dynamical formation of complex spatial clusters. Pre- liminary results showing the dependence on coupling strength, condensate size and dimensionality are obtained. I. INTRODUCTION The use of quantum simulation as a route to better understanding of complex quantum dynamics has much to recommend it. Conceptually, this creates an analog quantum computer. In quantum simulations, a table- top experiment is carried out to mimic a more com- plex quantum system which we would like to understand. This method has been used to treat, for example, models of condensed matter phase-transitions [1]. Such an ap- proach can be complemented by the use of approximate computer simulations. A numerical simulation can then be verified and tested in the table-top experiment, while allowing a wider variety of parameters to be treated. But why stop at condensed matter: why not model the entire universe? This seems presumptuous, and pos- sibly is. The entire universe will never be shoehorned into a table-top experiment with every complexity in- tact. Nevertheless, cosmologists today often use quan- tum field theory models to describe the early universe. This can be traced back to the pioneering works of Higgs and colleagues [2–5], studying the origins of mass, and to Coleman’s groundbreaking studies on unstable quantum vacuum decay [6, 7]. A combination of these approaches, together with the inclusion of general relativity, leads to the current in- flationary universe scenario [8–13]. Inflation provides a possible explanation for the origin of structure in the universe. Can we model at least part of this picture of the early universe in a laboratory? The simplest model for the scalar inflaton field φ(x) is described by the La- grangian L = 1 2 μ φ∂ μ φ - V (φ), (1.1) where V (φ) is the potential down which the scalar field rolls or decays. In this paper, we show how to real- ize the above relativistic scalar field model in a coupled * [email protected] Bose condensate, and model the fate of the scalar field of the early universe. We note that there have been earlier proposals using somewhat different, albeit related tech- niques [14, 15]. While the potential V (φ) is specific, depending on the model of inflation under consideration, coupled Bose con- densates provide a situation with V (φ) ∝- cos(φ). In the spirit of inflation, we consider a cold, unstable vac- uum with φ = π as the initial condition. The vacuum gradually decays [9–11], to produce a hot universe, re- plete with dynamical clumping into random structures. This quantum dynamical “universe on a table-top” exper- iment is probed by means of an internal Rabi rotation, and imaged. Relativistic quantum field theory is usually tested ex- perimentally at large accelerators. However, energies at high energy particle accelerators like CERN are not high enough for these field theories, although observational evidence for the Higgs boson [16, 17] is thought to pro- vide evidence for the low energy sector of scalar quantum fields in the universe. Instead of accelerating particles to light speed, we propose, essentially, to slow the velocity of light down to atomic speeds. This allows us to study novel phenomena outside of CERN’s limits. A quantum simulation of interacting relativistic fields in two, three or four space-time dimensions has the useful feature that it allows us to access quantum physics that we cannot do experiments on by any other means. We note some recent experiments have explored rele- vant physics with ultra-cold atoms, demonstrating that our proposal is indeed feasible. These show that the physics of interest here is very close to realization. The investigations that are similar to our proposal include long time-scale interferometry with a two-level BEC [18], and the study of the thermalization of a BEC [19]. More recent experiments have realized a flat trap or “can” for a BEC in three dimensions [20]. In principle, simula- tions of this type might eventually produce results that can predict the early universe temperature maps [21] pro- duced recently by the Planck telescope surveys, or very large-scale density inhomogeneities. Here we have a more cautious goal, of simply being able to demonstrate quan- arXiv:1305.5314v2 [cond-mat.quant-gas] 21 Aug 2013

arXiv:1305.5314v2 [cond-mat.quant-gas] 21 Aug 2013 · + g 1 y 1 1 + g c 2 2 ˙ 1 2; i~@ t 2 = ˆ ~2 2m r2 x + g 2 y 2 2 + g c 1 1 ˙ 2 1: (2.1) Here, r2 x representstheD-dimensionalLaplacianop-erator,

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Quantum simulations of the early universe

    B. Opanchuk1, R. Polkinghorne1, O. Fialko2, J. Brand2, P. D. Drummond1∗1Centre for Atom Optics and Ultrafast Spectroscopy,

    Swinburne University of Technology, Melbourne 3122, Australia and2Institute of Advanced Studies, Massey University, Albany, New Zealand

    A procedure is described whereby a linearly coupled spinor Bose condensate can be used as aphysically accessible quantum simulator of the early universe. In particular, an experiment to gen-erate an analog of an unstable vacuum in a relativistic scalar field theory is proposed. This is relatedto quantum theories of the inflationary phase of the early universe. There is an unstable vacuumsector whose dynamics correspond to the quantum sine-Gordon equations in one, two or three spacedimensions. Numerical simulations of the expected behavior are reported using a truncated Wignerphase-space method, giving evidence for the dynamical formation of complex spatial clusters. Pre-liminary results showing the dependence on coupling strength, condensate size and dimensionalityare obtained.

    I. INTRODUCTION

    The use of quantum simulation as a route to betterunderstanding of complex quantum dynamics has muchto recommend it. Conceptually, this creates an analogquantum computer. In quantum simulations, a table-top experiment is carried out to mimic a more com-plex quantum system which we would like to understand.This method has been used to treat, for example, modelsof condensed matter phase-transitions [1]. Such an ap-proach can be complemented by the use of approximatecomputer simulations. A numerical simulation can thenbe verified and tested in the table-top experiment, whileallowing a wider variety of parameters to be treated.

    But why stop at condensed matter: why not modelthe entire universe? This seems presumptuous, and pos-sibly is. The entire universe will never be shoehornedinto a table-top experiment with every complexity in-tact. Nevertheless, cosmologists today often use quan-tum field theory models to describe the early universe.This can be traced back to the pioneering works of Higgsand colleagues [2–5], studying the origins of mass, and toColeman’s groundbreaking studies on unstable quantumvacuum decay [6, 7].

    A combination of these approaches, together with theinclusion of general relativity, leads to the current in-flationary universe scenario [8–13]. Inflation provides apossible explanation for the origin of structure in theuniverse. Can we model at least part of this picture ofthe early universe in a laboratory? The simplest modelfor the scalar inflaton field φ(x) is described by the La-grangian

    L = 12∂µφ∂

    µφ− V (φ), (1.1)

    where V (φ) is the potential down which the scalar fieldrolls or decays. In this paper, we show how to real-ize the above relativistic scalar field model in a coupled

    [email protected]

    Bose condensate, and model the fate of the scalar field ofthe early universe. We note that there have been earlierproposals using somewhat different, albeit related tech-niques [14, 15].

    While the potential V (φ) is specific, depending on themodel of inflation under consideration, coupled Bose con-densates provide a situation with V (φ) ∝ − cos(φ). Inthe spirit of inflation, we consider a cold, unstable vac-uum with φ = π as the initial condition. The vacuumgradually decays [9–11], to produce a hot universe, re-plete with dynamical clumping into random structures.This quantum dynamical “universe on a table-top” exper-iment is probed by means of an internal Rabi rotation,and imaged.

    Relativistic quantum field theory is usually tested ex-perimentally at large accelerators. However, energies athigh energy particle accelerators like CERN are not highenough for these field theories, although observationalevidence for the Higgs boson [16, 17] is thought to pro-vide evidence for the low energy sector of scalar quantumfields in the universe. Instead of accelerating particles tolight speed, we propose, essentially, to slow the velocityof light down to atomic speeds. This allows us to studynovel phenomena outside of CERN’s limits. A quantumsimulation of interacting relativistic fields in two, threeor four space-time dimensions has the useful feature thatit allows us to access quantum physics that we cannot doexperiments on by any other means.

    We note some recent experiments have explored rele-vant physics with ultra-cold atoms, demonstrating thatour proposal is indeed feasible. These show that thephysics of interest here is very close to realization. Theinvestigations that are similar to our proposal includelong time-scale interferometry with a two-level BEC [18],and the study of the thermalization of a BEC [19]. Morerecent experiments have realized a flat trap or “can” fora BEC in three dimensions [20]. In principle, simula-tions of this type might eventually produce results thatcan predict the early universe temperature maps [21] pro-duced recently by the Planck telescope surveys, or verylarge-scale density inhomogeneities. Here we have a morecautious goal, of simply being able to demonstrate quan-

    arX

    iv:1

    305.

    5314

    v2 [

    cond

    -mat

    .qua

    nt-g

    as]

    21

    Aug

    201

    3

    mailto:[email protected]

  • 2

    tum simulations of an unstable quantum vacuum in arelativistic quantum field theory.

    Our starting point is a trapped Bose condensate withtwo internal quantum levels. The levels are coupled by amicrowave field, and the atoms interact through S-wavescattering. We assume zero temperatures, homogeneouscouplings and perfect microwave phase stability. Whiledepartures from such perfection can be treated, and theserequirements can be relaxed, we will study the ideal casehere. We show that by introducing a simple π phaseshift into the coupling field, it is possible to generatean experimentally accessible model of an unstable rela-tivistic vacuum. A quantum simulation of the decay ofthis unstable vacuum then provides the simplest of earlyuniverse models. The relevant field variable is the rela-tive phase of the two condensates. In a regime of smallcoupling, the relative phase obeys the sine-Gordon equa-tion [22, 23], a popular model for a relativistic scalar fieldtheory where the field potential V (φ) is a cosine.

    This paper demonstrates the feasibility of a class ofexperiments which can simulate fundamental aspects ofmodels of inflation. Such experiments, as well as theoret-ical calculations and numerical simulation, should followa step-by-step programme that starts with the simplestpossible theories. In the early stages, naturally, one willlearn a lot about how to extract meaningful data fromexperiments and the adequacy of approximate numeri-cal techniques but not much about the universe. Herewe illustrate our approach by a numerical simulation ofquantum effects, using a truncated Wigner approxima-tion described later. The results show the importanceof a hybrid approach, combining numerical and experi-mental techniques. Each method involves different phys-ical approximations. By comparing them, one can hopeto come to an improved understanding of how quantumfluctuations behave in the nonlinear regime, where exactquantum field predictions are hard to come by. As theprogramme is extended to include additional quantumfields and the effects of gravity one can hope to connectthe results of experiments with astronomical observationsand refine our models of the universe.

    In general, the longer term fate of vacuum fluctuationsin the nonlinear regime is of great fundamental inter-est. It is not known how to calculate this exactly, dueto quantum complexity and the failure of perturbationtheory. We speculate that these nonlinear effects may berelated to very large scale inhomogeneities in the massdistribution of the universe. This in turn could providethe basis for observational tests of inflationary universetheories beyond those available now. Unlike some mod-ern models of inflation [9–11], our system can supportdomain walls of relative phase [24]. For experimentalstudies, these provide an easily measured signature ofvacuum decay, and may have cosmological implications.

    The paper is organized as follows. In Section (II) wedescribe the general theory of coupled Bose fields, asfound in Bose-Einstein condensate (BEC) experimentson ultra-cold atomic gases at nanoKelvin temperatures.

    In Section (III), we treat the different types of classi-cal vacuum that exist, and the density-phase representa-tion. In Section (IV) we transform to the sine-GordonLagrangian, and demonstrate how this can behave as arelativistic field theory with an unstable vacuum. In Sec-tion (V) we carry out detailed, though approximate, nu-merical simulations of the original field theory, to indicatewhat to expect in an experiment. Finally, in Section (VI)we discuss the conclusions.

    II. COUPLED BOSE FIELDS

    Our first task then, is to slow down light: to a fewcentimeters per second if possible. How is this to beachieved? For the purpose, we use quasi-particles withdispersion relations equivalent to relativistic equations,in a coupled ultra-cold Bose condensate (BEC). To un-derstand this, let us consider the dynamical equationsfor a coupled condensate; a D-dimensional Bose gas withtwo spin components that are linearly coupled by an ex-ternal microwave field. This system obeys the followingnon-relativistic Heisenberg equation:

    i~∂tΨ1 ={− ~

    2

    2m∇2x + g1Ψ

    †1Ψ1 + gcΨ

    †2Ψ2

    }Ψ1 − νΨ2,

    i~∂tΨ2 ={− ~

    2

    2m∇2x + g2Ψ

    †2Ψ2 + gcΨ

    †1Ψ1

    }Ψ2 − νΨ1.

    (2.1)

    Here, ∇2x represents the D-dimensional Laplacian op-erator, and g1, gc are the D−dimensional coupling con-stants, which are assumed positive. These have knownrelations with the measured scattering length. Thecoupling-constant is renormalizable at large momentumcutoff, although our simulations are in the regime of lowmomentum cutoff, which makes renormalization unnec-essary. The field commutators are:[

    Ψ1 (x) ,Ψ†1 (x

    ′)]

    = δD (x− x′) (2.2)

    We use spinor notation, introducing Ψ = (Ψ1,Ψ2)Tand write the energy functional as W =

    ´dDxw, where

    the energy density w(x) is

    w =~2

    2m∇xΨ† · ∇xΨ− νΨ†σxΨ +

    gs2

    : (Ψ†Ψ)2 : +

    +gsa2

    : (Ψ†σzΨ)2 : +ga : (Ψ

    †Ψ)(Ψ†σzΨ) : .

    (2.3)

    Here we have defined 2gs = 12 (g1+g2)+gc, 4ga = g1−g2,and 2gsa = 12 (g1 + g2)− gc. From now on, for simplicitywe will assume the case of symmetric self coupling g1 =g2 ≡ g , so that ga = 0. In the case where gc = g,we further find gsa = 0. It is important for our purposes

  • 3

    that the cross-term gc have a different strength to the self-term gi. This is generally achievable in ultra-cold atomicphysics, depending on the atomic species and particularFeshbach resonance used for magnetically tunable cases.The fully symmetric case with gc = g is less interesting,and we will assume that gc = 0 for definiteness in laternumerical examples.

    We recover the Heisenberg equation from i∂tψi =δW/δψ∗i . Alternatively, we will also recover these equa-tions from a stationary action principle δS = 0 with

    S =

    ˆdt dDx [R{Ψ†i~∂tΨ} − w] =

    ˆdt dDx [LB − w] .

    (2.4)

    The coupling field ν is supplied by an external microwavesource, and couples the hyperfine levels in the atomic con-densate together. In experimental realizations [18], thishas an adjustable amplitude and phase, good homogene-ity, and a long coherence time.

    Before proceeding further, we rescale the equationsinto natural units. The time and distance scale is cho-sen so that mean-field frequency shifts are of order unity,as are the corresponding Laplacian terms. For typicalatomic densities per Bose field of n = N/V , where N isthe particle number of a given species, and V the volume,this leads to the choice τ = t/t0, and ζ = x/x0, where:

    t0 = ~/gnx0 = ~/

    √gnm . (2.5)

    Scaling the fields so that they correspond to densitiesin the new fields, i.e. ψ = ΨxD/20 , the resulting dimen-sionless field equations in the symmetric case are

    i∂τψ1 =

    {−1

    2∇2 + γψ†1ψ1 + γcψ

    †2ψ2

    }ψ1 − ν̃ψ2,

    i∂τψ2 =

    {−1

    2∇2 + γψ†1ψ2 + γcψ

    †1ψ1

    }ψ2 − ν̃ψ1. (2.6)

    Here, ∇ represents derivatives with respect to ζ, γ =1/(nxD0 ), γc = gc/(gnxD0 ), and ν̃ = ν/(gn). The dimen-sionless field commutators are:[

    ψ1 (ζ) , ψ†1 (ζ′)]

    = δD (ζ − ζ′) . (2.7)

    In one dimension, γ2 = γLL ≡ mg/(~2n), whichis the famous Lieb-Liniger parameter [25] of the one-dimensional quantum Bose gas.

    III. STABLE AND UNSTABLE VACUA

    We now consider a semi-classical approach. We defineψ̃1 as a classical mean field, and we linearize around theclassical equilibrium solutions. Vacuum solutions withconstant fields, apart from an oscillating phase, are eas-ily found from Eq. (2.6) as ψ̃01 =

    √ñ = ±ψ̃02 . An overall

    phase is chosen to make ψ̃01 positive and ñ = nxD0 is thedimensionless density. The different signs of ψ̃02 corre-spond to two inequivalent vacua. In order to study theirstability properties and elementary excitations, we in-vestigate small oscillations around the vacuum solutions.This generalizes the analysis presented in Ref. [26] to in-clude the cross-coupling with γc.

    A. Linearized solutions

    Writing ψ̃j = [ψ̃0j+(ujkei(k̃ζ−ω̃τ)+vj∗k e

    −i(k̃ζ−ω̃τ))]e−iµ̃τ ,where µ̃ is the vacuum chemical potential, substitutinginto Eq. (2.6) and keeping linear terms we obtain thesecular equation det[B − ω̃] = 0 with

    B =

    H0 γñ ∓ν̃ + γcñ γcñ−γñ −H0 −γcñ ±ν̃ − γcñ∓ν̃ + γcñ γcñ H0 γñ−γcñ ±ν̃ − γcñ −γñ −H0

    ,(3.1)

    where H0 = − 12 k̃2 + (2γ + γc)ñ− µ̃ and consistency de-

    mands that 2γsñ = µ̃± ν̃, where we define 2γs = γ + γc.The sign "+" corresponds to the in-phase and "-" corre-sponds to the out-of-phase vacua. We find two indepen-dent solution branches of the secular equation. The firstone is gapless

    ω̃1 =

    √1

    2k̃2(

    1

    2k̃2 + 4γsñ

    ), (3.2)

    with eigenvector ∼ (1,−1, 1,−1)T for small k, which cor-responds to the phase fluctuations of the fields δ arg ψ̃1 =δ arg ψ̃2 ∝ sin(k̃ζ − ω̃1τ). This leads to a sound wavewith sound speed ṽBog =

    √2γsñ. Solutions with real

    frequencies indicate stable, propagating waves of smallamplitude, whereas imaginary roots indicate an instabil-ity. Excitations along this branch are always stable when2γs ≡ γ + γc > 0.

    The other branch is gapped, since

    ω̃2 =

    √(1

    2k̃2 ± 2ν̃

    )(1

    2k̃2 + 4γsañ+ 2ν̃

    ), (3.3)

    where 2γsa = γ − γc. The nonlinearity parameter in thissecond branch differs from the first branch and can betuned independently due to the presence of the cross-coupling γc. The eigenvector corresponding to ω2 is∼ (1,−1,−1, 1)T for small k and ν, which corresponds toan excitation of relative phase fluctuations of the fieldsδ arg ψ̃1 = −δ arg ψ̃2 ∝ sin(k̃ζ − ω̃2τ). It is interesting tonote that the two branches are decoupled in the linear ap-proximation. We will be particularly concerned with therelative phase dynamics, which has characteristic prop-erties of the elementary excitations of a relativistic fieldtheory. Indeed, expanding to second order in k̃ we maywrite ω̃22 = k̃2c̃22 + m̃22c̃42, which is a relativistic dispersion

  • 4

    relation with a light speed of c̃22 = 2γsañ±2ν̃ and rest en-ergy m̃22c̃42 = 4ν̃(2γsañ± ν̃). The relative phase dynamicswill analyzed be more rigorously in the next section.

    The second branch also supports unstable modes andgoverns the dynamics of the unstable vacuum. We areparticularly interested in the situation where both γs andγsa are positive. In this case the second branch becomesunstable if the linear coupling is such that ±ν̃ < 0, withthe unstable modes satisfying −4γsañ ∓ 2ν̃ < k̃2/2 <∓2ν̃. This occurs either with the fields having the samesign and ν̃ < 0, or with the fields having the oppositesign and ν̃ > 0.

    Thus, in a sufficiently large system where k is continu-ous, one of the two possible vacua is stable and the otherone unstable. The classical field dynamics resulting fromthe unstable vacuum has been discussed for some specialcases in Refs. [27, 28] (see also Ref. [26]). In this pa-per we are going to simulate the decay from the unstablevacuum due to quantum fluctuations. The characteristiclength scale of the decay modes is larger than π/

    √|ν̃|,

    which will become the largest length scale in the systemin the interesting regime where |ν̃| is small. The timescales of the unstable mode become large in the sine-Gordon regime. This is compatible with the followinginflation requirement: When the scalar field rolls downthe potential hill very slowly compared to the expansionof the universe, inflation occurs.

    B. Density-phase representation

    We will continue to consider these equations as clas-sical, or mean-field equations. We use this procedure togive us a better understanding of how the dynamics canbe reduced to those of a relativistic scalar field. Once wehave made this transformation, we will use Lagrangianquantization methods to construct a low-energy effectivequantum field theory for the elementary excitations.

    We parametrize the spinor as follows:

    ψ̃1 = uei(φs+φa)/2 cos(θ)

    ψ̃2 = uei(φs−φa)/2 sin(θ), (3.4)

    by introducing the density mixing angle θ and total den-sity u2, the relative phase φa and total phase φs. Thefirst part of the dimensionless Lagrange density becomes

    L̃B =t0x0~LB = R{ψ̃†i∂τ ψ̃}

    =u2

    2∂τφs +

    u2

    2cos(2θ)∂τφa. (3.5)

    Writing the energy density as w̃ = K̃ + Ṽ, the kinetic

    energy part becomes

    K̃ =12∇ψ̃† · ∇ψ̃

    =1

    2

    {(∇u)2 + u2(∇θ)2 +

    [(∇φs)2 + (∇φa)2

    ] u24

    +∇φs · ∇φau2

    4cos(2θ)

    }. (3.6)

    For the potential part we find

    V =− ν̃u2 cosφa sin(2θ) +γs2u4

    +γsa4u4[1 + cos(4θ)], (3.7)

    where 2γs = γ+γc and 2γsa = γ−γc. The potential is afunction of the three real field variables u, θ, and φa. Acut for constant u is shown in Fig. 1. Note:

    • As long as γsa ≥ 0, there is a valley at θ = π/4corresponding to a vacuum with equal density inthe two fields. This expresses a well-known con-dition for the stability of two coupled Bose fields,g1 + g2 ≥ 2gc, where only the total particle num-ber is conserved. This condition is weaker thanthe miscibility condition for binary mixtures with-out interconversion [29], g1g2 > g2c , and turnsinto it when g1 = g2. The vacuum is locatedat (θ = π/4, φa = 0). An equivalent point is(θ = 3π/4, φa = π). Working at fixed particle num-ber N , we find for the dimensionless total densityof the vacuum u2 = 2ñ ≡ 2nxD0 = γ−1.

    • As θ terms primarily couple to the relative phaseφa, there are dynamical solutions following the bot-tom of the θ valley with φa = 0 and θ = π/4corresponding to solutions of the type ψ1(x, t) =ψ2(x, t). This is just the dynamics of a singleD-dimensional Bose field involving compressionalwaves (Bogoliubov phonons) and dark solitons.

    • The point (θ = π/4, φa = π) or, equivalently, (θ =3π/4, φa = 0) is an unstable point.

    • Excitations of the relative phase φa have low energycost when |ν̃| � γsau2, i.e. in the regime of smallcoupling and strong nonlinearity. This is the mainfocus of this paper.

    We can see from Fig. 1 that within each large canyonin θ-space, there is a secondary feature where a hill andvalley form in the φ−direction. We wish to focus onthis relative phase feature. It is able to form a quantumfield with a relativistic dispersion relation, albeit with aneffective “light” velocity of very much smaller size than c.

    C. Domain walls

    Domain walls of relative phase appear as a connec-tion between two degenerate vacua along a θ valley.

  • 5

    Figure 1. Plotted is V for ν̃ = 0.1, u = 1, γsa = 0.5 as a func-tion of θ and φa. At θ = 3π/4 a saddle point occurs, whichcorresponds to π phase jump between the two components. Itis a known and dynamically unstable stationary state of theGP equations [26, 27].

    The corresponding solution were discussed in the con-text of three-dimensional two-component BECs with in-ternal coupling in Ref. [24]. In the context of one di-mensional coupled BECs they are also known as atomicJosephson vortices or rotational fluxons, since they arerelated to circular atomic currents, and were discussedin Refs. [22, 30–32]. As features of the relative phase,they only exist at sufficiently small linear coupling ν̃ <1/3 [30]. They are unstable and can decay if a mecha-nism for dissipation of energy is provided unless ν̃ . 0.14,where they become local energy minima with topologi-cal stability [33]. In this regime, they acquire the typicalproperties of topological solitons, or kinks, in the sine-Gordon equation. By tuning the coupling parameter ν̃,one can thus move between regimes of different stabilityproperties of domain walls. The regime dominated bysine-Gordon physics is found for small ν̃.

    IV. SINE-GORDON REGIME

    Aiming at the dynamics of the relative phase φa, wesimplify the action and corresponding equations of mo-tion by assuming that all dynamics takes place in a θvalley and that both θ and u only have small devia-tions from their respective equilibrium values. We conse-quently write θ = π/4+y and expand the action to lead-ing orders in y. Using cos(2θ) = −2y, sin(2θ) = 1− 2y2,

    cos(4θ) = 8y2 − 1 we obtain

    Ṽ = ν̃u2 cosφa(1− 2y2)−γs2u4 − 2γsau4y2

    2K̃ = (∇u)2 + u2(∇y)2 +{

    (∇φs)2 + (∇φa)2} u2

    4

    −∇φs∇φau2

    2y

    L̃B =u2

    2∂τφs − u2y∂τφa. (4.1a)

    Low energy dynamics is possible in two sectors: com-pressional waves involving u and φs on the one hand, anddynamics of the relative phase involving φa and y on theother. Keeping only the terms that are relevant for thelatter, the Lagrangian density reads

    L̃ = −u2(∂τφa)y −u2

    8(∇φa)2 + ν̃u2 cos(φa)− 2γsau4y2.

    (4.2)

    Variation with respect to the fields gives the Euler-Lagrange equation

    ∂L∂f−∇ ∂L

    ∂∇f− ∂τ

    ∂L∂fτ

    = 0, (4.3)

    which implies that∂τφa + 4γsau2y = 0. This allows us toreplace y in the Lagrangian to yield

    L̃ = 18γsa

    (∂τφa)2 − ñ

    4(∇φa)2 + 2ν̃ñ cos(φa). (4.4)

    This is the D-dimensional sine-Gordon Lagrangian.Now we need to extract parameters. The critical speed

    c̃ is found by dividing the ∇ pre-factor by the ∂τ pre-factor and taking the square root to yield c̃ =

    √2γsañ.

    The critical velocity thus matches the Bogoliubov speedof sound ṽbog =

    √2γsñ only in the case of vanishing

    cross coupling γc, where γsa = γs. We note that so farwe have used classical arguments. However, our originalequations are also valid as quantum field equations. If wequantize the Lagrangian, we accordingly have a quantumsine-Gordon equation, which describes the physics at suf-ficiently low real temperatures.

    On quantizing the Lagrangian, Eq. (4.4), the canonicalmomentum field is:

    π =∂τφa4γsa

    ,

    with corresponding canonical commutators of:[φa (ζ) , π

    (ζ′)]

    = iδD(ζ − ζ′

    ).

    The sine-Gordon action can be brought into Lorenz-invariant form by rescaling time as ζ0 = c̃τ to read

    S = ~ˆdD+1ζ

    1

    2

    {(∂ζ0φ

    ′)2 − (∇φ′)2}

    +α̃

    β2cos(βφ′),

    (4.5)

  • 6

    with the rescaled field φa = βφ′ and β2 = 2√

    2γsaγ, α̃ =4ν̃. Here, β2 is the universal dimensionless parameterof the quantum sine-Gordon equation. In one dimen-sion and the case of vanishing cross coupling γc = 0 and2γsa = γ, we obtain β2 = 2γ = 2

    √γLL, which again links

    to the Lieb-Liniger coupling parameter γLL = mg/(~2n).The stationary action principle leads to the (classical)

    sine-Gordon equation

    ∇2φa − ∂ζ0ζ0φa + α̃ sinφa = 0. (4.6)

    Note that the original field φa reappears and the parame-ter β completely drops out of the classical equation. Thisis a relativistic field equation with a dimensionless speedof light of unity, as in all sensible relativistic field theo-ries. Alternatively, using atomic units, we consider c̃ asthe effective light-speed.

    The parameter α = α̃x−20 is connected to the char-acteristic length scale of the sine-Gordon equation. Inphysical units

    `SG =1√α

    =~

    2√νm

    . (4.7)

    Remarkably, this length scale depends on the tunnel cou-pling but is independent of the nonlinearity. It is thesecond characteristic length scale besides the GP healinglength x0 of Eq. (2.5). It sets the spatial scales of thedomain walls in the sine-Gordon regime.

    The parameter α̃ is also connected to the restmass of the elementary excitations of the sine-Gordonequation. Indeed, small amplitude oscillations ∼exp−i(k̃ζ − ω̃sGτ) of the vacuum have a relativistic dis-persion relation of ω̃2sG = k̃

    2c̃2 + α̃c̃2, with a rest massof m̃sG =

    √α̃/c̃ and a rest energy of m̃2sGc̃

    4 = 8ν̃γsañ.Note that the sine-Gordon light speed c̃ and rest massm̃sG agree to leading order in ν̃ with the correspondingparameters of the gapped second branch of the Bogoli-ubov spectrum of the stable vacuum (3.3). The relativis-tic sine-Gordon equation thus describes asymptotically asector of excitations of the vacuum that, for small ampli-tude, completely decouple from the compressional wavesof the Bogoliubov sound.

    In the course of inflation tiny quantum fluctuationsare important: they form the primordial seeds for allstructure created in the later Universe. We now turn tocomputational quantum simulations of these equations.

    V. COMPUTATIONAL SIMULATIONS

    To investigate the feasibility and likely behavior ofthese experiments, we have carried out a number of nu-merical simulations of the expected dynamics. We simu-lated the full coupled Bose fields, using atomic densities,confinement parameters and S-wave scattering lengthsgenerally similar to those employed in recent experimen-tal studies with ultra-cold atoms. We employed an accu-rate fourth-order interaction-picture algorithm. Careful

    monitoring of step-size errors was needed to ensure accu-rate results.

    The theoretical method used is a truncated, proba-bilistic version of the Wigner-Moyal [34, 35] phase-spacerepresentation [36–40]. This is a probabilistic phase-space method for quantum fields, that correctly simulatessymmetrically-ordered quantum dynamics in the limit oflarge numbers of atoms per mode. It has been shownto give results in agreement with quantum-limited pho-tonic and BEC experiments [41, 42], provided the largeatom-number restriction is satisfied in the experiment.

    Other methods of this type are also possible. One ex-ample is the positive-P representation [43, 44]. In thisnormal-ordered approach, there is no truncation. Un-like the truncated Wigner method, no systematic errorsoccur at small occupation numbers [45]. However, withcurrent algorithms, the statistical sampling error growsin time for long time-scale nonlinear problems withoutdamping. This rules out the positive-P approach for thepresent simulations, until better algorithms are found.

    Although the resulting equations are similar to theGross-Pitaevskii (GP) equations, the initial noise termshave a precise meaning from quantum mechanics. Theycorrespond to the exact vacuum noise needed to generatesymmetrically ordered operator moments that includethe quantum fluctuations of the initial quantum state.The truncated Wigner method therefore includes quan-tum terms correct to order 1/n, which are omitted in theGP approach. Additional stochastic terms are neededif there is decoherence or atomic absorption [40]. How-ever, in these preliminary studies, such effects will beomitted. We also omit the cross-coupling between spinsfor simplicity. Following the dimensional notation intro-duced previously, except that our c-number fields are nowstochastic representations of quantum fields, we obtain:

    i∂τ ψ̃1 =

    {−1

    2∇ζ2 + γ

    ∣∣∣ψ̃1∣∣∣2} ψ̃1 − ν̃ψ̃2,i∂τ ψ̃2 =

    {−1

    2∇ζ2 + γ

    ∣∣∣ψ̃2∣∣∣2} ψ̃2 − ν̃ψ̃1. (5.1)For the truncated Wigner calculations, with an initial

    coherent state of ψ̄(ζ), so that ψ̃(ζ) = ψ̄(ζ) + ∆ψ̃(ζ), onewould have:〈

    ∆ψ̃(ζ)∆ψ̃∗(ζ ′)〉

    =1

    2δ (ζ − ζ ′) .

    We emphasize here that our assumption of an initial co-herent state is certainly not the only one possible. Itcan be replaced by other quantum states. For example,a finite temperature ground-state followed by a suddenjump in phase is probably closer to the way that exper-iments will take place. Nevertheless, in the absence of agood model for the quantum state of the early universe,a coherent state is as reasonable a choice as any other.

    Experimental state preparation is a complex issue, anddepends on such details as the magnitude of the mi-crowave coupling during cooling. Our initial state cor-responds approximately to one in which the fields are

  • 7

    0 5 10 15 20 25τ

    −0.50

    −0.25

    0.00

    0.25

    0.50J

    Figure 2. Decay of 1D quantum field coherence J , as an indi-cator of Sine-Gordon unstable dynamics. γ = 0.1. Couplings,ν̃ = 0.001, 0.01, 0.1, 1 (blue solid, red dashed, green dash-dotted, yellow dotted). Condensate size L = 80, 256 spatialgrid points, 10000 time steps, 100 ensembles, opposite initialphases to give an unstable vacuum.

    cooled to a very weakly interacting ground state, witha coherent coupling of ν < 0, so that the lowest energystate has fields with an opposite sign. Then the inter-actions are increased to a large value, and the couplingphase is reversed to ν > 0, in order to create an unstablevacuum. Naturally, other models are possible, includinga strongly interacting initial state at finite temperature.These will be treated elsewhere.

    A. Dependence on couplings

    We firstly investigate the dependence of the system onthe coupling strength. This is a crucial part of the cal-culation, as it creates the potential minimum such thatthe relative phase is able to be treated as an independentscalar field. The measurable quantities in an experimentare the densities obtained after interfering the two hy-perfine field amplitudes [18]. These are the quantities:

    n± =1

    2

    〈(ψ̂1 ± ψ̂2

    )† (ψ̂1 ± ψ̂2

    )〉(5.2)

    They correspond simply to the atomic densities aftera rapid Rabi rotation of π/2 is performed to allow theatomic fields to interfere. Substituting the mean fields ofEq. (3.4) into Eq. (5.2), these measurable quantities are

    n± = u2

    [1

    2± 1

    2cosφa +O

    (y2)].

    At the vacuum phase, φa = 0, the atoms are all in theeven mode, with n+ = u2 and n− = 0. At the unstableequilibrium, φa = π, n+ = 0, and n− = u2. We plot therelative visibility, which is a quadrature-like measure ofthe phase, namely [46]:

    J =

    ´dDζ (n+ − n−)

    4N≈ 1

    2cosφa .

    0 5 10 15 20 25τ

    −0.50

    −0.25

    0.00

    0.25

    0.50

    J

    Figure 3. Decay of average 1D quantum field coherence J fordifferent condensate sizes, to indicate the effects of bound-aries. Condensate size L = 10, 20, 40, 80 (blue solid, reddashed, green dash-dotted, yellow dotted), spatial grid sizes32, 64, 128, 256, 10000 time steps, 100 ensembles, γ = 0.1.Coupling ν̃ = 0. Opposite initial phases.

    We see in Fig. 2, that for very weak couplings, there isa relative phase decay induced purely by quantum phasediffusion. This is the opposite to sine-Gordon physics,and occurs because the two condensates are uncoupled,and experience an increasingly random relative phase.For stronger couplings, with ν̃ ≈ 0.1, one can see the dis-tinctive rapid decay of an unstable vacuum, which is thesignature of sine-Gordon physics with this initial condi-tion. The slight oscillation near ν̃ ≈ 0.1 may indicatesome form of collective instability of coupled domains,and deserves further investigation.

    B. Effects of condensate size

    Although experimentally one would use a finite trap,we simulate a system with periodic boundary conditionsfor simplicity. To some extent, this allows us to ignorethe detrimental effects of boundaries, and hence to allowa small computer simulation mimic a larger experimentalcondensate. However, this is not entirely the case.

    As shown in Fig. 3, the finite coherence length of thequantum fluctuations has an effect on the decay statis-tics. Initially, the condensate size has no effect, as the co-herence length is very small, and certainly much smallerthan the condensate size as long as it is larger than ahealing length. For increasing times, if ν̃ = 0, the co-herence length increases until it reaches the size of thecondensate itself. At this stage, the decay accelerates.For very large condensates, a boundary-independent be-havior is found. We note that this effect is strongest inthe limit of zero coupling ν̃, which is plotted here. Atlarger couplings the formation of domain walls limits thegrowth of coherence.

  • 8

    0 10 20 30 40 50 60 70τ

    403020100

    10203040

    z

    0.50.40.30.20.1

    0.00.10.20.30.40.5

    j

    Figure 4. Plot of 1D space-time dynamics of the local visibil-ity j for a single trajectory, showing contour plot vs time andspace. Coupling ν̃ = 0.1, condensate size 80, 256 spatial gridpoints, γ = 0.1, 10000 time steps.

    C. Single trajectory examples

    Quantum mechanics only predicts ensemble averages.However, it is now common to use quantum theory topredict the behavior of the universe, which is only a sin-gle ensemble member. Of course, this experiment is noteasy to repeat, especially as the observer is part of it.In the laboratory, one can obtain ensemble averages byrepeating the state preparation. Even so, the densitypatterns and evolution in a single ensemble member isinstructive. We expect the Wigner ensemble membersto have a characteristic behavior like individual labora-tory runs, and perhaps to the universe itself, if currenttheories are correct.

    In the one-dimensional case, this is shown in Fig. 4.This corresponds to the largest of the condensates plot-ted in Fig. 3, and it is visually clear that the densitycorrelation length is still substantially smaller than thecondensate size throughout the time evolution. The localfringe visibility is defined in general as:

    j =n+ − n−

    2 (n+ + n−)≈ 1

    2cosφa .

    Since the total density is nearly constant through-out, we used the excellent approximation of j ≈(n+ − n−) /(2ū2) for simplicity of calculation and plot-ting.

    D. Two-dimensional behavior

    The universe is certainly not one-dimensional. Any-one might reasonably question the applicability of a one-dimensional simulation, even though this is the mostamenable to theoretical treatment. In the laboratory, theavailability of engineered traps means that dimensional-ity can be readily adjusted to obtain artificial universesof one, two or three space dimensions.

    0 5 10 15 20 25τ

    −0.50

    −0.25

    0.00

    0.25

    0.50

    J

    Figure 5. Plot of 2D quantum field coherence J , as an indi-cator of sine-Gordon unstable dynamics. γ = 0.1. Couplings,ν̃ = 0.001, 0.01, 0.1, 1 (blue solid, red dashed, green dash-dotted, yellow dotted). Condensate size L = 80, 256 × 256spatial grid points, 4000 time steps, 200 ensembles, oppositeinitial phases.

    The average dynamics of the sine-Gordon field dependon dimensionality as shown in Fig. 5, which shows thatthe two-dimensional averages differ from Fig. 2, althoughthere are qualitative similarities. However, Fig. 6 is muchmore interesting, demonstrating clearly the growth ofcomplex spatial structures. These are two-dimensionalcross-sections, taken at dimensionless times of τ = 5,τ =10 andτ = 20. These single-trajectory plots show howthe early, relatively high momentum quantum noise pat-terns have decayed into coherent spatial domains with atypical spatial coherence length of order ζ̄ = 10 spatialunits at τ = 10, and start to evaporate at later times. Itremains to be investigated whether this later evaporationis an artifact of the coupled BEC system, or a genuinesine-Gordon feature.

    E. Three-dimensional behavior

    Recent experiments have shown the possibility of ex-ploring a condensate in a “can” or cylindrical trap. Thisis a flat trap in three-dimensions, with reflecting bound-ary conditions. While our simulations do not have thistype of boundary, it is feasible to carry out large three-dimensional calculations relatively quickly using GPUhardware [42]. In Fig. 7, we show a slice through a con-densate, in which one coordinate is held fixed, while theother two are varied. This is a single ensemble mem-ber, and the other parameters are similar to the two-dimensional case, Fig. 6.

    The resulting spatial geometry of clusters of hot andcold matter is distinctly different to the two-dimensionalcase, with greater connectedness, as larger coherentstructures are formed over the same time-scale. Here wehave plotted the real spatial image of a two-dimensionalslice, to show domain-like clustering effects. This demon-

  • 9

    40 30 20 10 0 10 20 30 40y

    403020100

    10203040

    x

    0.50.40.30.20.1

    0.00.10.20.30.40.5

    j

    τ=5 τ=10 τ=20

    Figure 6. Plot of 2D spacial clustering of the local visibility j for a single trajectory, showing contour plot vs space at finaltime. Coupling ν̃ = 0.1, condensate size 80× 80, 128× 128 spatial grid points, γ = 0.1, 5000 time steps, opposite initial phases,time τ = 10.

    40 30 20 10 0 10 20 30 40y

    403020100

    10203040

    x

    0.50.40.30.20.1

    0.00.10.20.30.40.5

    j

    τ=5 τ=10 τ=20

    Figure 7. Plot of 3D spacial clustering geometry of the local visibility j for single trajectory, showing contour plot vs space,sliced through a single z-coordinate. Coupling ν̃ = 0.1, condensate size 80×80×80, 128×128×128 spatial grid points, γ = 0.1,5000 time steps, opposite initial phases. Plotted: a slice orthogonal to x-axis at τ = 10.

    strates that very large-scale clustering effects occur evenin the absence of explicit gravitational interactions, forour model.

    Rapid progress in imaging technology means that itis now possible to image such 2D slices in the experi-ment [47], although it is technologically more difficultthan typical BEC experimental techniques which producea column density image of an expanded condensate.

    VI. SUMMARY

    The general behavior and unstable dynamics of an ef-fective quantum sine-Gordon field using a two-mode BECwith linear coupling has been clearly demonstrated. Ex-perimentally, for the scenario outlined here, it is impor-

    tant to use an atomic species whose cross-coupling dif-fers from its self-coupling. This largely rules out the mostcommonly used atomic species, 87Rb, which has very sim-ilar self and cross couplings apart from a small regionnear a Feshbach resonance. A Feshbach-type experimentis still possible, although losses are greater in this regime.Apart from this case, there are many species which areboth able to be evaporatively cooled and have multiplehyperfine levels. For example,39K can be used. It has asizable difference of g1g2 − g2c [48, 49], and it is thereforemore suitable for realization of our proposal.

    Our results will be extended to a more complete anal-ysis of the numerical simulations elsewhere, in which wewill analyze in greater detail the effects of different typesof state preparation. However, it is clear that in prin-ciple the use of ultra-cold quantum gases as quantumsimulators is not restricted to condensed matter and non-

  • 10

    relativistic analogies. Simulating interacting relativisticquantum fields in one, two or three dimensions appearscompletely feasible, both experimentally and computa-tionally. An experimental implementation would throwmuch-needed light on the question of how accurate ournumerical approximations are for these parameter values.

    While we have focused on the practical issues of howto carry out the experiment and corresponding numericalsimulations, these issues pale somewhat in comparisonto another issue. We hope our proposal will stimulateinterest in this fundamental question, which is:

    • how do we carry out measurements on the quantumuniverse, when we are part of it?

    ACKNOWLEDGEMENTS

    The authors acknowledge helpful discussions with An-drei Sidorov, Tim Langen and Uli Zülicke. OF andJB were supported by the Marsden Fund (contractNo. MAU0910). PDD was funded by the Australian Re-search Council.

    [1] D. Porras and J. I. Cirac, Phys. Rev. Lett. 92, 207901(2004).

    [2] P. W. Higgs, Phys. Lett. 12, 132 (1964).[3] F. Englert and R. Brout, Phys. Rev. Lett. 13, 321 (1964).[4] P. W. Higgs, Phys. Rev. Lett. 13, 508 (1964).[5] G. Guralnik, C. Hagen, and T. Kibble, Phys. Rev. Lett.

    13, 585 (1964).[6] S. Coleman, Phys. Rev. D 15, 2929 (1977).[7] C. Callan and S. Coleman, Phys. Rev. D 16, 1762 (1977).[8] A. H. Guth, Phys. Rev. D 23, 347 (1981).[9] A. Linde, Phys. Lett. B 259, 38 (1991).

    [10] A. Linde, Phys. Rev. D 49, 748 (1994).[11] A. Linde, Phys. Rev. D 58, 083514 (1998).[12] E. D. Stewart and D. H. Lyth, Phys. Lett. B 302, 171

    (1993).[13] A. Liddle, P. Parsons, and J. Barrow, Phys. Rev. D 50,

    7222 (1994).[14] U. Fischer and R. Schützhold, Phys. Rev. A 70, 063615

    (2004).[15] C. Neuenhahn, A. Polkovnikov, and F. Marquardt, Phys.

    Rev. Lett. 109, 085304 (2012).[16] ATLAS Collaboration, Phys. Lett. B 716, 1 (2012).[17] CMS Collaboration, Phys. Lett. B 716, 30 (2012).[18] M. Egorov, R. P. Anderson, V. Ivannikov, B. Opanchuk,

    P. D. Drummond, B. V. Hall, and A. I. Sidorov, Phys.Rev. A 84, 021605(R) (2011).

    [19] M. Gring, M. Kuhnert, T. Langen, T. Kitagawa,B. Rauer, M. Schreitl, I. Mazets, D. A. Smith, E. Demler,and J. Schmiedmayer, Science 337, 1318 (2012).

    [20] A. L. Gaunt, T. F. Schmidutz, I. Gotlibovych, R. P.Smith, and Z. Hadzibabic, Phys. Rev. Lett. 110, 200406(2013).

    [21] Planck Collaboration, “Planck 2013 results. I. Overviewof products and scientific results,” arXiv:1303.5062(2013).

    [22] V. Kaurov and A. Kuklov, Phys. Rev. A 73, 013627(2006).

    [23] V. Gritsev, A. Polkovnikov, and E. Demler, Phys. Rev.B 75, 174511 (2007).

    [24] D. Son and M. Stephanov, Phys. Rev. A 65, 063621(2002).

    [25] E. H. Lieb and W. Liniger, Phys. Rev. 130, 1605 (1963).[26] J. Brand, T. J. Haigh, and U. Zülicke, Phys. Rev. A 81,

    025602 (2010).[27] I. Lesanovsky and W. von Klitzing, Phys. Rev. Lett. 98,

    050401 (2007).

    [28] T. W. A. Montgomery, R. G. Scott, I. Lesanovsky, andT. M. Fromhold, Phys. Rev. A 81, 063611 (2010).

    [29] B. D. Esry, C. Greene, J. Burke, Jr., and J. Bohn, Phys.Rev. Lett. 78, 3594 (1997).

    [30] V. Kaurov and A. Kuklov, Phys. Rev. A 71, 011601(2005).

    [31] J. Brand, T. J. Haigh, and U. Zülicke, Phys. Rev. A 80,011602 (2009).

    [32] M. I. Qadir, H. Susanto, and P. C. Matthews, J. Phys.B 45, 035004 (2012).

    [33] S.-W. Su, S.-C. Gou, A. S. Bradley, O. Fialko, andJ. Brand, “Kibble-Zurek scaling and its breakdown forspontaneous generation of Josephson vortices in Bose-Einstein condensates,” arXiv:1302.3304 (2013).

    [34] E. P. Wigner, Phys. Rev. 40, 749 (1932).[35] J. E. Moyal, Math. Proc. Cambridge 45, 99 (1947).[36] P. D. Drummond and A. D. Hardman, Europhys. Lett.

    21, 279 (1993).[37] M. Steel, M. K. Olsen, L. I. Plimak, P. D. Drummond,

    S. Tan, M. J. Collett, D. F. Walls, and R. Graham, Phys.Rev. A 58, 4824 (1998).

    [38] A. Sinatra, C. Lobo, and Y. Castin, J. Phys. B 35, 3599(2002).

    [39] P. B. Blakie, A. S. Bradley, M. J. Davis, R. J. Ballagh,and C. W. Gardiner, Adv. Phys. 57, 363 (2008).

    [40] B. Opanchuk and P. D. Drummond, J. Math. Phys. 54,042107 (2013).

    [41] J. F. Corney, J. Heersink, R. Dong, V. Josse, P. D. Drum-mond, G. Leuchs, and U. L. Andersen, Phys. Rev. A 78,023831 (2008).

    [42] B. Opanchuk, M. Egorov, S. E. Hoffmann, A. I. Sidorov,and P. D. Drummond, Europhys. Lett. 97, 50003 (2012).

    [43] P. D. Drummond and C. W. Gardiner, J. Phys. A: Math.Gen. 13, 2353 (1980).

    [44] S. J. Carter, P. D. Drummond, M. D. Reid, and R. M.Shelby, Phys. Rev. Lett. 58, 1841 (1987).

    [45] P. P. Deuar and P. D. Drummond, Phys. Rev. Lett. 98,120402 (2007).

    [46] Q.-Y. He, T. G. Vaughan, P. D. Drummond, and M. D.Reid, New J. Phys. 14, 093012 (2012).

    [47] R. Bücker, A. Perrin, S. Manz, T. Betz, C. Koller, T. Plis-son, J. Rottmann, T. Schumm, and J. Schmiedmayer,New J. Phys. 11, 103039 (2009).

    [48] C. D’Errico, M. Zaccanti, M. Fattori, G. Roati, M. In-guscio, G. Modugno, and A. Simoni, New J. Phys. 9,223 (2007).

    http://dx.doi.org/10.1103/PhysRevLett.92.207901http://dx.doi.org/10.1103/PhysRevLett.92.207901http://dx.doi.org/10.1016/0031-9163(64)91136-9http://dx.doi.org/10.1103/PhysRevLett.13.321http://dx.doi.org/10.1103/PhysRevLett.13.508http://dx.doi.org/10.1103/PhysRevLett.13.585http://dx.doi.org/10.1103/PhysRevLett.13.585http://dx.doi.org/10.1103/PhysRevD.15.2929http://dx.doi.org/10.1103/PhysRevD.16.1762http://dx.doi.org/10.1103/PhysRevD.23.347http://dx.doi.org/10.1016/0370-2693(91)90130-Ihttp://dx.doi.org/10.1103/PhysRevD.49.748http://dx.doi.org/10.1103/PhysRevD.58.083514http://dx.doi.org/10.1016/0370-2693(93)90379-Vhttp://dx.doi.org/10.1016/0370-2693(93)90379-Vhttp://dx.doi.org/10.1103/PhysRevD.50.7222http://dx.doi.org/10.1103/PhysRevD.50.7222http://dx.doi.org/10.1103/PhysRevA.70.063615http://dx.doi.org/10.1103/PhysRevA.70.063615http://dx.doi.org/10.1103/PhysRevLett.109.085304http://dx.doi.org/10.1103/PhysRevLett.109.085304http://dx.doi.org/10.1016/j.physletb.2012.08.020http://dx.doi.org/10.1016/j.physletb.2012.08.021http://dx.doi.org/ 10.1103/PhysRevA.84.021605http://dx.doi.org/ 10.1103/PhysRevA.84.021605http://dx.doi.org/10.1126/science.1224953http://dx.doi.org/10.1103/PhysRevLett.110.200406http://dx.doi.org/10.1103/PhysRevLett.110.200406http://arxiv.org/abs/1303.5062http://arxiv.org/abs/1303.5062http://dx.doi.org/10.1103/PhysRevA.73.013627http://dx.doi.org/10.1103/PhysRevA.73.013627http://dx.doi.org/10.1103/PhysRevB.75.174511http://dx.doi.org/10.1103/PhysRevB.75.174511http://dx.doi.org/10.1103/PhysRevA.65.063621http://dx.doi.org/10.1103/PhysRevA.65.063621http://dx.doi.org/10.1103/PhysRev.130.1605http://dx.doi.org/10.1103/PhysRevA.81.025602http://dx.doi.org/10.1103/PhysRevA.81.025602http://dx.doi.org/10.1103/PhysRevLett.98.050401http://dx.doi.org/10.1103/PhysRevLett.98.050401http://dx.doi.org/10.1103/PhysRevA.81.063611http://dx.doi.org/ 10.1103/PhysRevLett.78.3594http://dx.doi.org/ 10.1103/PhysRevLett.78.3594http://dx.doi.org/10.1103/PhysRevA.71.011601http://dx.doi.org/10.1103/PhysRevA.71.011601http://dx.doi.org/10.1103/PhysRevA.80.011602http://dx.doi.org/10.1103/PhysRevA.80.011602http://dx.doi.org/10.1088/0953-4075/45/3/035004http://dx.doi.org/10.1088/0953-4075/45/3/035004http://arxiv.org/abs/1302.3304http://arxiv.org/abs/1302.3304http://arxiv.org/abs/1302.3304http://dx.doi.org/10.1103/PhysRev.40.749http://dx.doi.org/10.1017/S0305004100000487http://dx.doi.org/10.1209/0295-5075/21/3/005http://dx.doi.org/10.1209/0295-5075/21/3/005http://dx.doi.org/10.1103/PhysRevA.58.4824http://dx.doi.org/10.1103/PhysRevA.58.4824http://dx.doi.org/10.1088/0953-4075/35/17/301http://dx.doi.org/10.1088/0953-4075/35/17/301http://dx.doi.org/ 10.1080/00018730802564254http://dx.doi.org/10.1063/1.4801781http://dx.doi.org/10.1063/1.4801781http://dx.doi.org/10.1103/PhysRevA.78.023831http://dx.doi.org/10.1103/PhysRevA.78.023831http://dx.doi.org/ 10.1209/0295-5075/97/50003http://dx.doi.org/10.1088/0305-4470/13/7/018http://dx.doi.org/10.1088/0305-4470/13/7/018http://dx.doi.org/10.1103/PhysRevLett.58.1841http://dx.doi.org/10.1103/PhysRevLett.98.120402http://dx.doi.org/10.1103/PhysRevLett.98.120402http://dx.doi.org/10.1088/1367-2630/14/9/093012http://dx.doi.org/ 10.1088/1367-2630/11/10/103039http://dx.doi.org/ 10.1088/1367-2630/9/7/223http://dx.doi.org/ 10.1088/1367-2630/9/7/223

  • 11

    [49] C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, Rev.Mod. Phys. 82, 1225 (2010).

    http://dx.doi.org/10.1103/RevModPhys.82.1225http://dx.doi.org/10.1103/RevModPhys.82.1225

    Quantum simulations of the early universeAbstractI IntroductionII Coupled Bose fieldsIII Stable and unstable vacuaA Linearized solutionsB Density-phase representationC Domain walls

    IV Sine-Gordon regimeV Computational SimulationsA Dependence on couplingsB Effects of condensate sizeC Single trajectory examplesD Two-dimensional behaviorE Three-dimensional behavior

    VI Summary Acknowledgements References