<<

Eur. Phys. J. B 78, 439–447 (2010) DOI: 10.1140/epjb/e2010-10176-y THE EUROPEAN PHYSICAL JOURNAL B Regular Article

Superfluid density in a two-dimensional model of supersolid

N. Sep´ulveda1,2,C.Josserand2,a, and S. Rica3,4 1 Laboratoire de Physique Statistique, Ecole´ Normale Sup´erieure, UPMC Paris 06, Universit´eParisDiderot,CNRS, 24 rue Lhomond, 75005 Paris, France 2 Institut Jean Le Rond D’Alembert, UMR 7190 CNRS-UPMC, 4 place Jussieu, 75005 Paris, France 3 Facultad de Ingenier´ıa y Ciencias, Universidad Adolfo Ib´a˜nez, Avda. Diagonal las Torres 2640, Pe˜nalol´en, Santiago, Chile 4 INLN, CNRS-UNSA, 1361 route des Lucioles, 06560 Valbonne, France Received 28 February 2010 / Received in final form 27 October 2010 Published online 6 December 2010 – c EDP Sciences, Societ`a Italiana di Fisica, Springer-Verlag 2010 Abstract. We study in 2-dimensions the superfluid density of periodically modulated states in the frame- work of the mean-field Gross-Pitaevskiˇı model of a quantum . We obtain a full agreement for the su- perfluid fraction between a semi-theoretical approach and direct numerical simulations. As in 1-dimension, the superfluid density decreases exponentially with the amplitude of the particle interaction. We discuss the case when defects are present in this modulated structure. In the case of isolated defects (e.g. dislocations) the superfluid density only shows small changes. Finally, we report an increase of the superfluid fraction up to 50% in the case of extended macroscopical defects. We show also that this excess of superfluid fraction depends on the length of the complex network of grain boundaries in the system.

1 Introduction models, that a perfect of helium atoms cannot ex- hibit a supersolid behavior, since local exchange does not Superfluidity is a crucial property at low temperature of produce a finite NCRI in the thermodynamic limit [14,15]. many quantum or such as Helium In fact, there is no insight yet of any microscopic mecha- or Bose-Einstein condensates. On the contrary, the ques- nism explaining supersolidity and the Bose-Einstein con- tion of superfluidity in crystalline structures remains more densation of vacancies, early proposed by Andreev and complicated since the varying density inhibits the quan- Lifshitz [1], is still questioned [16]. The unknown complex tum coherence. This phenomenon, called “super- structure of the solid helium ground state wave function solidity”, has been of large speculation for the makes this challenge difficult and therefore a general the- last forty years, since seminal works of Andreev-Lifshitz, ory of supersolidity of solid helium – if it is supersolidity Reatto, Chester and Leggett [1–4]. Today, after the break- that explains the experimental observations – is still far through experiments by Kim and Chan [5–7] supersolid- from reachable. In this sense, a substantially less ambi- ity appears as a new striking feature of solid Helium be- tious subject is the study of supersolidity in more sim- low 100 mK. As originally established [4], supersolidity is ple systems than solid helium, such as cold atoms. There, manifested through the loss of a fraction of the moment one replaces a crystal structure by a density modulated of inertia of a torsional oscillator filled by solid Helium. system. Density modulated Bose-Einstein condensates are This non classical rotational inertia (NCRI), which is an currently formed by the imposition of an external poten- important property of superfluid systems, has nowadays tial, creating the so-called optical lattices [17,18]. In these been confirmed by more than six different experimental cases, by varying the properties of the optical lattice, the groups [8–13], using different geometries and conditions. condensate is shown to exhibit a Mott /superfluid However, it seems that the whole story – from theoret- [17–20]. More recently, it has also been ical and experimental points of view – is far from over. suggested recently that supersolidity might be present for Essentially, a coherent theoretical framework that would Rydberg atoms in dipole-blockade regime [21,22]. In refer- account the experimental findings is still lacking. In fact ence [21] the authors take an artificial interaction potential many open questions need clarifications at least at a fun- which allow a ground-state computation, which happens damental level. to display the properties of supersolidity proving that, un- In particular: what are the features of the wave func- der some conditions, a quantum system of interacting par- tion to display or not superfluidity? It is well known that ticles exhibits a crystalline structure and the superfluid- Chester-Reatto wavefunction displays off diagonal long ity property. The present article concerns thus the study range order (ODLRO) [2,3], thus NCRI. However, it has of the superfluid density in the framework of cold atoms been suggested, using quantum Monte-Carlo numerical where a mean field theory can be applied.

a e-mail: [email protected] 440 The European Physical Journal B

A model for density modulated quantum system can be with the direct numerical calculation of the superfluid obtained in the general framework of the Gross-Pitaevskiˇı fraction. Here we find that the√ superfluid fraction de- (GP) equation [23,24], using a non-local [25,26]oranex- creases again with Λ,logfss ∼− Λ, at least for moderate ternal [17,18] potential. This mean-field GP equation can values of Λ, as in 1-dimension. For higher values of Λ,itis be deduced for Bose-Einstein condensates using the di- harder to draw a strong conclusions from our results. How- lute approximation which assumes that the N body ever, beside these needed technical but interesting aspects wave function, ΨN (r1, r2,...rN ), is factorized in terms of of comparing the semi-analytical macroscopic approach a single “condensate wave function” ψ(r), such that with the direct numerical calculations, 2-dimensions is crucial because it is the lowest relevant spatial dimension ΨN (r1, r2,...rN )=ψ(r1) × ψ(r2) ×···×ψ(rN ). to investigate and to disentangle the contribution of the crystal defects to the superfluid fraction. Indeed, although This approximation cannot apply formally to Helium the NCRI feature has always been experimentally noticed, atoms although such models have commonly been used the large variability of the amount of the NCRI fraction to describes the dynamics of Helium at low temperature. (NCRIF) as the geometries or the solid samples change, Superfluidity is then an inherent property of the GP model gives rise to a large scientific debate. The nature of the su- as it can be proved using the Leggett calculation of the persolidity is questioned as well as its dependence on the NCRIF lower bound [27]. Thus, the GP equation describes specific properties of the solid crystal structure (presence the evolution of a wave-function that exhibits, even with- of defects or vacancies in particular) [1,3,13,26,32–34]. For out an external potential, both crystallisation of the sys- instance, the macroscopic superflow can be dominated by tem (thus a solid) and a coherent quantum phase (showing multiple distinct paths that may avoid the absolute min- superfluidity): this is a supersolid showing in particular imum of the wave function. In the last section, we show the NCRI property [26]. As discussed above, the striking how the presence of defects enhance the supersolid frac- property of this model is not to exhibit the superfluid- tion in the model, with qualitative agreement with the ity, but rather the elastic behavior such as a solid [28]. experimental observations. In the same spirit, Anderson [29] has recently considered a Gross-Pitaevskii theory for a weakly interacting Bose- Einstein condensate of vacancies displaying the superfluid 2 The Gross-Pitaevskii equation property responsible of the non classical rotational inertia. Despite the fact that vacancies attract themselves [16], so with a nonlocal interaction potential that the superfluid is unstable in the long wave regime, the basic crystal structure of the solid is not considered in The original GP equation [24] for the complex wave func- this model so that the coupling between the condensate tion ψ(x,t)in2Dis: and the crystal displacements are not taken into account.  ∂ψ 2 By developing the long wave dynamics, we have ob- i = − ∇2ψ + ψ(x,t) U(|x − y|)|ψ(y)|2dy. (1) tained from the GP equation a macroscopic model of ∂t 2m a supersolid, coupling the elastic properties of the solid with the quantum phase [26,28]. We want to emphasize Equation (1) describes, by a mean-field approximation, ψ x,t U r here that the pertinence of this macroscopic approach the evolution of the wave-function ( )where ( )is goes beyond the details of the GP model and has been the two body potential between the atoms. proposed elsewhere with different methods [30]. In [31] The evolution (1) is Hamiltonian, we have studied this Gross-Pitaevskii model of supersolid δH in 1-dimension, using both direct numerical simulations i∂ ψ , t = ∗ and analytical estimates for the ground state solution. We δψ have compared the ground state characteristics such as the peak size, obtained via variational arguments, with di- with the energy or Hamiltonian   rect numerical calculations showing an impressively good 2  2 1 agreement. Consideringperiodicsystems,wehavecom- H |∇ψ| dx U |x − y| |ψ y |2|ψ x |2dydx. = m + ( ) ( ) ( ) puted the NCRIF using both the Leggett quotient [4], 2 2 (2) and direct numerical simulations, with full quantitative H N agreement. Thanks to such comparisons, we have shown This energy , the number of particles , that for the GP equation in 1-dimension the NRCIF de-  2 creases exponentially when increasing the dimensionless N = |ψ(x)| dx, (3) interaction parameter (Λ defined later in Eq. (7)). More√ precisely, the superfluid fraction follows log fss ∼− Λ and the linear momentum P , for large Λ because of the exponentially small dependence  of the profile of the ground state density with Λ. i P = − (ψ∗∇ψ − ψ∇ψ∗)dx, (4) In this paper, we present a detailed study of this model 2 in 2-dimensions. As for the 1-dimension case, the aim of the study is first to compare semi-analytical estimates of are conserved by the dynamical evolution of (1). Accord- the superfluid fraction using homogenization techniques ing to the energy dependence on the phase of the wave N. Sep´ulveda et al.: Superfluid density in a two-dimensional model of supersolid 441

    √ function, one notices that the ground state solution is al- With x = x/a, t = ma2 t,andψ = ψ/ n0,the ways real since any non uniform phase increases its energy. dimensionless GP equation for the Heaviside interaction Consider now a homogeneous solution ψ0 = (we drop the primes hereafter) reads: √ −i E0 t n e  n E  0  ,with 0 the mean density and 0 = ∂ψ 1 πn ∞ U r rdr i = − ∇2ψ + Λψ(x,t) Θ(|x − y|)|ψ(y)|2dy, (9) 2 0 0 ( ) . Then the perturbations around this ∂t 2 homogeneous solution give dispersive waves with a dis- 1 persion relation following the Bogoliubov spectrum rela- where Θ(|x − y|)= π θ(1 −|x − y|). tion [35]: In the following, we consider a square domain:  Ω =[0,L] × [0,L] with periodic boundary conditions 2 2 2 2 2 ψ(x, y, t)=ψ(x + L, y, t)andψ(x, y, t)=ψ(x, y + L, t). ωk = ( k /2m) +( k /m) n0Uˆk, (5) Note that here L is dimensionless. Uˆ We then define the energy and number of particles where k is the Fourier transform that has to be bounded densities of the system: for all k:     1 2 Λ ∞ 2π E = |∇ψ| dx + Θ(|x − y|)|ψ(y)|2 ikr cos θ Ω Ω Uˆk = U(r)e rdrdθ 2 Ω 2 Ω 0  0 ×|ψ(x)|2dydx, (10) ∞  =2π U(r)J0(kr) rdr. (6) 1 2 0 N = |ψ(x,t)| dx =1, (11) Ω Ω Remarkably, the spectrum (5) depends only on a single which are the relevant control parameters of the system. dimensionless parameter [25]: ma2 Λ n Uˆ 3 The ground state and macroscopic = 0 2 k=0 (7) equations denoting a the characteristic length scale of the potential. Λ To allow stability of the long wave modes, the potential For large enough values of ,equation(9) displays modu- U(r) has also to satisfy: lated structure solutions. We recall in this section the re- sults deduced for the dynamics near this modulate ground  ∞ state [25,26,28]. 0 < 2π U(r)rdr = Uˆk=0 < ∞. 0 3.1 Ground state Moreover, we shall require that Uˆk becomes negative at some kc to allow for the mecha- The ground state can be determined by minimizing the nism [36–38]. For instance in the numerics later on, we energy functional (10) together with the constraint (11): choose the soft core interaction, with no loss of generali- in general this ground state exists and provides a pure real ties (although in the case of Rydberg atoms [21,22]thein- and positive wave function. For low values of Λ the ground 3 teraction possesses a long range tail 1/r in 2-dimensions, state is homogeneous in space: that is a liquid without the roton spectrum (5) does not change substantially): positional order. As Λ increases there is a critical value where the liquid phase (the homogenous solution) is no U(|x − y|)=U0θ(a −|x − y|), (8) longer the lowest energy solution. There, the competition between the non-local interaction and the kinetic term in θ · with ( ) the Heaviside function. In 2D, the Fourier trans- equation (10) allows for the formation of density gradients form of this special interaction potential writes: such that the energy is then lower than the one of the homogenous solution. In reference [25] a weakly nonlinear J1(ka) Uˆk πU0a , analysis exhibited the formation of a crystalline structure =2 k for large enough Λ in two and three spatial dimensions. In 2-dimensions for the potential (8) the crystal appears where J1 is the Bessel function of the first kind. Uˆk=0 = 4 for Λ 41.6. For larger Λ, the ground state becomes a πa2U , Λ n πma U 0 so that = 0 2 0 in this special case. periodic structure of well defined tiny peaks (see Fig. 1). In We emphasize that the special choice of the poten- the large Λ limit, this ground state can even be described tial (8) is purely of practical interest because it is easy to as a solution of the 2-dimensional packing problem [39]. implement in numerical schemes. Other functions whose Fourier transform would be negative for a wave number domain (strictly bigger than zero), would show similar 3.2 Macroscopic equations properties. In particular, it is possible to find the poten- tial such that the Bogoliubov dispersion relation matches To obtain the dynamical properties of this crystal, we the Landau spectrum with the right values of the speed need to extract the long and slow variation of the solu- of sound c, and the three roton parameters [25]. tion from the periodic structure. In references [26,28], we 442 The European Physical Journal B

sponse of the system. If the phase Φ(x,t)canbene- glected, one recovers the usual equation for the elastic de- formation of macroscopical bodies. Finally, equation (14) is a Bernoulli like equation for the quantum phase, that gives the usual Landau equation for the superfluid veloc- s vs2  ity ∂tv = −∇ 2 + E (ρ) . Thus, this macroscopic set of equations shows that the GP equation (9)canmodela system that shows both superfluidity (if we find a non-zero ss ik ) and the usual elasticity of solid bodies.

3.3 Superfluid density In reference [28] we have obtained, using a homogeniza- Fig. 1. (Color online) Ground state for a 2D system with tion technique [40], an expression for the superfluid den- ss periodic boundary conditions. This ground state correspond sity matrix ik deduced directly from the ground state Λ . 1 L L |ψ x, y |2 dx dy 2 to =8236 an average density L2 0 0 ( ) =1. density profile of the crystal ρ0(x)=|ψ0(x)| .Thesuper- The total size of the system is L = 64 units and the range of fluid fraction tensor is: a  interaction is = 16. The system displays 30 peaks. ss  ρ0(x)(∇Ki · ∇Kk) dx f ss ik δ − V  , ik = ρ = ik ρ x dx (15) have described these low energy excitations around the V 0( ) ground state: expanding the solution as a slowly and large- where the auxiliary functions Kk (two in two space di- scale varying crystal structure, and using homogenization mensions) are defined only in the unitary lattice cell V techniques [40], we have obtained a set of macroscopic and satisfy: equations for the mean density ρ(x,t), the coherent phase Φ(x,t) and the elastic deformation u(x,t)ofthecrystal: ∂kρ0 + ∇ · (ρ0(x)∇Kk)=0 inV (16)   ∂ρ ∂ ss  ss with periodic boundary condition at the frontier of the + ik ∂kΦ +(ρδik − ik )˙uk = 0 (12) ∂t ∂xi m cell.    ρ x . ∂  If the crystal modulation is absent ( 0( )=const), ss K K m (ρδik − ik ) u˙k − ∂kΦ then x = y = 0 (up to constants) so that the superfluid ∂t m f ss → δ     fraction tensor ik ik. In general, there is no formal so- ∂ ss  lution to the problem (15, 16) for an arbitrary ρ0(x)inany +  (ρδil − il ) u˙ l − ∂lΦ ∂kΦ ∂xk m spatial dimension (the only exception is the 1-dimension ∂ case, see [31]). However, because of the elliptic character = (λiklmulm) (13) of equations (16), since ρ0(x) > 0, the symmetric matrix ∂x ss k f can easily be computed numerically. 2 ik ∂Φ  2  On the other hand, the superfluid fraction can also  + (∇Φ) + E (ρ)=0, (14) ∂t 2m be defined directly through the momentum carried by the 1 superfluid part. The total momentum, from (4), is where uik = 2 (∂iuk + ∂kui) is the usual strain tensor in   u ∂ u  ∂ Φ∂ u the theory of elasticity, ˙ i = t i + m k k i is the co- Pi =  ρ0(x)∇iφdx = mvk ρ0(x)(δik + ∂iKk) dx variant time derivative of the elastic deformation ui and ss il is the superfluid tensor. The set of equations (12–14) (17) and one gets the same definition of the superfluid fraction is in fact a simplified (the linearized) version of the gen- ss eral macroscopic model where some nonlinear terms have since: Pi = m ρ0(x) dxfik vk. been dropped for the sake of clarity. The parameters in these macroscopical equations, such as the elastic modulus  4 Numerical computations of the ground λiklm, the chemical potential E (ρ) and the superfluid den- ss state and the superfluid density matrix sity ik do depend on the local density, ρ,andcanbecom- puted ab-initio from knowledge of the underlying crystal 4.1 Ground state of the system structure [28]. Practically, these parameters are obtained by solving differential equations at the level of a unit cell The numerical results are obtained by minimizing the of the periodic structure, characterized by the wave func- energy equation (10) under the number of particles condi- tion equation (11). For this purpose, we use the Ginzburg- tion of the ground state ψ0(x). In this article, we focus ss Landau (GL) version of the dynamics which can be inter- on the behavior of the superfluid density ik for which we will show its resolution in details below. Equation (12) preted as the integration of the GP equation for imaginary time t = −ıτ: is the mass conservation equation for the total density ρ,  ss s ss with the current ji = ik vk +(ρδik − ik )˙uk and where ∂ψ 1 2 2 s  = μψ + ∇ ψ − Λψ(x,t) Θ(|x − y|)|ψ(y)| dy, we have defined the “superfluid velocity” v = m ∇Φ. ∂τ 2 Ω Equation (13) is the usual equation for the elastic re- (18) N. Sep´ulveda et al.: Superfluid density in a two-dimensional model of supersolid 443 where μ is the Lagrangian multiplier introduced to satisfy the number of particles condition (11). Note that the total momentum of the ground state is zero. We use a pseudo- spectral numerical scheme, where the linear dynamics is computed in the Fourier space while the non-linear part is performed in real space. The Lagrange multiplier is found by a Newton method to reach the correct total number of particles. The Fourier transforms are computed using the FFTW library1. Figure 1 shows the ground state computed for Λ = (a) (b) 82.36 in a 64 × 64 periodic box. This ground state forms K x, y an hexagonal pattern of 30 peaks. Fig. 2. (Color online) Density plot of (a) x( )and(b) Ky(x, y), calculated directly from a cell of the ground state Figure 1. The parameters are a = 16, Λ =82.36. 4.2 Numerical computation of superfluid density following the solution of Kx and Ky In this section, we present the numerical calculation of the superfluid fraction through the formalism of the ho- mogenization described above. We need to calculate Ki(x) from (16) with periodic boundary conditions on the op- posite faces of the hexagonal unit cell V .Todothiswe use a finite element method: we first extract from the nu- merics the ground state ρ0(x) in the unit cell V .Next,we build a triangular mesh inside the unit cell and we have used nv = 100 vertices for the triangular mesh on which we have extrapolated the above ground state density ρ0 Fig. 3. (Color online) Plot of the different components of the obtained by minimizing the GP energy. Next, we solve superfluid matrix computed from formula (15) after calculation equation (16) in a variational form by multiplying (16) of the Kx and Ky, as a function of the control parameter Λ. ss ss ss ss with an arbitrary function v(x)periodicinV .Integrating The different plots represent: fxx(•), fxy(), fyx(), fyy(). over V , we obtain the two equations (for i =1, 2):  with f(x)=∂iρ0(x). The Ki(x) are finally computed by ρ x ∇K x ·∇v x − ∂ ρ x v x dx , [ 0( ) i( ) ( ) i 0( ) ( )] =0 (19) solving this linear system for the n vector (α1,α2, ..., αn). V Applying this technique, we plot Kx(x, y)andKy(x, y) valid for every periodic function v(x) defined on V .To in Figure 2 inside the unit cell of the ground state dis- solve the equation for each Ki, we expand them in a finite played in Figure 1.BothKx and Ky have a nodal line element basis compatible with the boundary conditions of that follows directly from (16). Indeed equations (16) n the problem, {vl(x)}l=1 (where n is the dimension of the may be seen as two Poisson-like equations with “elec- basis, typically n = nv = 100): tric charge distributions” ∂xρ0(x, y)and∂yρ0(x, y) respec- K x, y K x, y ρ x, y n tively for x( )and y( ). Since 0( )isasym- metric function, its gradient ∂xρ0(x, y)(resp.∂yρ0(x, y)) K(x)= αlvl(x) l=1 has a vertical nodal line (resp. horizontal) at the center of the cell. The function Kx (Ky) exhibits thus the follow- where we simplify the notations, K(x)=Ki(x), i =1, 2. ing anti-symmetry with this axis Kx(−x, y)=−Kx(x, y) Usually this basis is composed of Lagrange polynomial (Ky(x, −y)=−Ky(x, y)) where the origin is at the center functions. Inserting the expansion of K(x)intoequa- of the cell. Finally, because of the periodic boundary con- tion (19), we obtain a linear system of n different equations ditions there is also a nodal curve that matches with the by taking v(x)=vl(x)forl =1, 2,...n: contour of the cell. n Because of the accurate determination of the ground state solution ρ0(x, y), we are able to compute from the Ki Gjlαl = Mj the values of f ss via the formula (15), up to Λ of the order l=1 of 400. Every four entries of the 2×2 matrix are plotted in where  Figure 3. The results exhibit clearly the x − y symmetry f ss f ss Gjl ρ0 x ∇vj x ·∇vl x dx, whereas the off-diagonal terms ( xy and yx) are an order = ( ) ( ) ( ) ss ss ss V below the diagonal ones (fxx and fyy). In conclusion, f and  can be correctly approximated by a diagonal matrix: Mj = f(x)vj (x)dx, f ss ∼ f ss I V xx

1 www.fftw.org where I is the identity matrix in 2-dimensions. 444 The European Physical Journal B

(a) Fig. 4. (Color online) Plot of the different components of the ss ss ss ss superfluid matrix: fxx(•), fxy(), fyx(), fyy() as a function of Λ.

4.3 Direct numerical computation of the superfluid density

An alternative and complementary way to compute the superfluid density is by direct numerical simulation of the GP equation (9). A natural way to mimic the superfluid experiment [41] consists of imposing a given boost v to the system. The resulting current is then related to the ss superfluid fraction fik . The stationary state is obtained by minimizing the free energy: (b) F = E + v · P + μ (N−1), Fig. 5. (Color online) Comparison between eigenvalues of the matrix f ss calculated by the homogenization formula (15) where P is the dimensionless linear momentum deduced (•, ) and by direct computation of the matrix (20)(, ). from (4). We have to keep in mind that the minimiza- (a) Plots the eigenvalues of f ss as function of Λ and (b) plots tion of the free energy is in the moving reference frame, the logarithm (log) of the eigenvalues. The solid line indicates ss 1/2 so that we need to substract from the total momentum the exponential fit log f ∝−Λ as suggested by the theory. the bulk momentum density mv ρ0(x) dx.Thesuper- fluid fraction follows then from the momentum computed in the stationary state: direct numerical simulation (20), we show a good agree- ∂P ss  1 i ment between the two methods in Figure 5. The agreement fik = δik − lim . (20) Λ ∼ |v|→0 ρ0(x) dx ∂vk is particularly good up to 150 while the fluctuations in the direct measurements affect the comparison above ss ss ss ss Figure 4 shows the four elements fxx, fxy, fyx, fyy of the Λ ∼ 150, although the trend is continued. Despite this lack superfluid fraction matrix as a function of the parameter of precision in the direct method, we have demonstrated Λ by this direct numerical simulation. As a consequence that both methods are equivalent, as already shown in of the imposed boost v = 0 in the numerics, the resulting 1-dimension [31]. In 1-dimension, we have exhibited an curves are noisier than those obtained previously in Fig- exponential decrease of the superfluid fraction with the ure 3. The same kind of numerical limitations were already square-root of the control parameter log f ss ∝−Λ1/2 [31], observed in 1-dimension [31]. Again, we observe that the as predicted by the Leggett estimates of superfluid den- results present the x − y symmetry and that the matrix sity [27,39]. This estimate has been shown to be valid to is dominated by the diagonal terms so that f ss can be higher spatial dimensions [39] but it corresponds then to approximated by a multiple of the identity matrix I.This a lower bound of the superfluid fraction only. In our nu- isotropic property of the matrix can be exhibited by mea- merics, we observe that this exponential behavior is still ss ss ss ss suring the anisotropy parameter (f1 − f2 )/(f1 + f2 ), valid in 2D at least for Λ<150 as shown in Figure 5b. ss ss where f1 and f2 are the eigenvalues of the superfluid ForhighervaluesofΛ, it is harder to draw strong con- fraction matrix. We have found that in all our results, clusions from our results and a best fit would rather give this anisotropy is always smaller than 10%, even in the a power law decrease f ss ∝ Λ−0.6 than the exponential large Λ limit where the error bars and fluctuations are law. At this stage we have no clear and definite explana- higher. We observe also that the anisotropy is lower for tion for this apparent change of behavior although a few larger systems. propositions can be discussed: it is likely that the fluc- Comparing the eigenvalues of the superfluid fraction tuations in our numerics do not allow us to discriminate matrix computed by formula (15) with those obtained by strongly between the two behaviors; on the other hand, the N. Sep´ulveda et al.: Superfluid density in a two-dimensional model of supersolid 445 exponential decrease only describes a lower bound based on the density profile of the ground state. It does not ac- count for the multiple paths that can follow the superflow in more than 1-dimension since the superflow could in- deed be enhanced by finding its way along the low density regions of the solid.

5 Defects-enhaced superfluid density

The suspected dependence of the supersolid fraction on (i) (ii) the specific structure of the crystal lattice is expected to be more pronounced in the presence of defects or vacancies Fig. 6. (Color online) Stationary state corresponding to two × in the model. The dynamics of the vacancies is in fact at random initial conditions (box size is 256 256), (i) represents the heart of the Andreev-Lifshitz model of a supersolid [1] a state with only two points defects of the crystal pattern. and it might be important to quantify it in our mean-field (ii) shows f a complex network (sketched roughly by hand) of defects that forms a randomly oriented boundary grain. approach. In order to tackle the role and the influence of defects for superfluidity, we will now investigate dif- ss ferent situations where defects are present. For that pur- Table 1. The eigenvalues of the matrix f for two random f ss pose, we cannot use the homogenization technique since initial conditions. We see an increment in the for the cases when a macroscopic network of defects in the are it applies only for quasi-periodic structures. As explained × above, we will rather in this case use the direct numerical presents. In these simulations the box size is 256 256 units, a range of interaction of a =8andΛ = 128.68. simulation imposing a galilean boost to the structure to ss ss measure the superfluid fraction. We focus on two distinct f1 f2 configurations that represent the two main types of de- (i) 0.052 0.052 fect structures. The first one is obtained by starting a GL (ii) 0.059 0.058 computation with random initial conditions so that only a Perfect crystal 0.051 0.051 few stable defects remains in the domain. The other case consists of a grain boundary crossing the whole domain. 5.2 Grain boundary

5.1 Random defects It is tempting to investigate a situation where an ordered grain boundary is introduced in the system. To realize We use a fixed range of interaction a =8foravalue such a situation, we consider a modulated structure made of Λ = 128.68. We start with the unstable homogenous by the adjunction of two monocrystals having different , , state to which a random perturbation of small ampli- orientational order, one with (1 0 0) and the second one , , tude is added: ψ = 1 + noise (this initial wave function with crystalline order (3 2 0), as shown in Figure 7.To is normalized to the unit average density). Under the dis- create these crystals, we consider the following different sipative dynamics (18), we converge to (quasi-)stationary wave functions: (metastable-)solutions that exhibit defects. These solu- √ ψ(1) x  ky k x y / tions are local minimizers of the Hamiltonian (the global 0 ( )=1+ cos( )+cos ( 3+ ) 2 one is free of defects and corresponds to the periodic crys- √ k x − y / tal structure) and are usually metastable. The defects can +cos ( 3 ) 2 disappear after very long dynamics (much slower than the √ ψ(2) x  kx k x y / time we use to measure the superfluid fraction) by double- 0 ( )=1+ cos( )+cos ( + 3) 2 annihilation or by adding noise. To measure NCRIF, the √ +cos k (x − y 3)/2 solution with defects can thus be considered as a steady state. Among the numerous initial conditions that we have (1) L (3) ψ0 (x) ∀ x, 0

(i) (ii) Fig. 7. (Color online) Stationary state corresponding to the (3) (4) initial conditions: (i) ψ0 ; (ii) ψ0 (box size 128 × 128). The interaction range is a =8andΛ = 128.68. (i) k =0.65 which is about the highest unstable wave num- ber for Λ = 128.68 and we take  =0.1 to initiate the instability that creates the crystal. Finally, each of the (3) (4) wave functions ψ0 and ψ0 contains two grain bound- aries located on the x and y axes respectively (one is in the middle of the sample, the other at one of the borders because of the periodic boundary conditions). We compare now the matrices f ss obtained for the two (1) perfect crystals issued of the initial wave functions ψ0 (2) (3) and ψ0 with those of the composite crystals ψ0 and ψ(4) 0 . For the perfect crystals we used again the homoge- (ii) nous based techniques while only the boost method (20) could be used for the composite crystals. Since the su- Fig. 8. (Color online) Plot of the eigenvalues of the matrix f ss perfluid fraction matrices are again almost diagonal, we for the four initial condition considered and for a box size × × ψ1 ♦ show in Figure 8 the eigenvalues f ss and f ss only, for the (i) 128 128 and (ii) 256 256. Initial states 0 ( )( )and 1 2 ψ2 Λ 0 ( )( ) correspond to two perfect crystal with different ori- four different states for different values of and for two ψ3 • ◦ ψ4  different system sizes, (i) 128 × 128 and (ii) 256 × 256. entations. The states 0 ( )( )) and 0 ( )( ) correspond to two crystalline states with a grain boundaries parallel to the In all cases, the isotropy is conserved, although the exist- axe x and y respectively. ing grain boundaries are oriented. We also observe that the presence of the grain boundaries always increases the superfluid fraction. However, this effect is higher for the Remarkably, this argument would be also valid in small system (Fig. 8i) than for the large one (Fig. 8ii). One 3-dimensions. can argue that when the size of the mono-crystal domain We end this article by considering the effect of thermal increases, the superfluid fraction should converge to the fluctuations. The case of 2-dimensions is known to be crit- perfect crystal one. Δf ss ical with respect to long range order with continuous sym- The difference between the composite and the metry, which is the case of the present model because it perfect crystals appears to depend only slightly on the U Λ possesses an (1) symmetry [42,43]. The long-wave acous- value of the parameter so that it can represent up to tic modes are easily excited by thermal fluctuations in 50% of the superfluid fraction for the small system at high the low temperature regime and that would destroy the Λ. We propose a simple argument to explain the behavior Δf ss phase coherence. Indeed, the phase fluctuations diverge of the excess superfluid fraction on the sample size. logarithmically with the distance while the spatial cor- Based on the numerical results, we make the assumption Δf ss relations decay eventually algebraically. This is however that is proportional to the fraction of grain bound- not a fundamental objection here since we are consider- ary: ing finite size systems and T = 0 dynamics. Moreover, d a Δf ss ∼ ∼ , (22) 2-dimensions Bose-Einstein condensates with a superfluid L2 L state have been observed experimentally. For such dilute gas systems the same argument on the breakdown of phase where  is the length of the grain boundary in the system coherence at large should apply, but this shows that such of typical thickness d. Assuming  ∼ L while d is of the effects can in fact be neglected practically there. Finally, order of the crystal mesh a, we finally have: in 3-dimensions, the phase fluctuations grow quadratically a in temperature and are lower if compression (density) in- Δf ss ∼ . L (23) creases [25]. N. Sep´ulveda et al.: Superfluid density in a two-dimensional model of supersolid 447

6Conclusion 12. M. Kondo, S. Takada, Y. Shibayama, K. Shirahama, J. Low Temp. Phys. 148, 695 (2007) In this work, we have presented a detailed study on the 13. B. Hunt, E. Pratt, V. Gadagkar, M. Yamashita, A.V. superfluid fraction in a model of a supersolid. We have de- Balatsky, J.C. Davis, Science 324, 632 (2009) veloped two different methods to compute this superfluid 14.D.M.Ceperley,B.Bernu,Phys.Rev.Lett.93, 155303 fraction applied to the Gross-Pitaevskiˇı mean-field theory. (2004) As shown in references [26,28] the superfluid fraction de- 15. N. Prokof’ev, B. Svistunov, Phys. Rev. Lett. 94, 155302 pends explicitly on the ground state wave function of the (2005) system. Since no simple analytical relation can be drawn 16. M. Boninsegni, A.B. Kuklov, L. Pollet, N.V. Prokofev, out of the ground state, we have performed a numerical B.V. Svistunov, M. Troyer, Phys. Rev. Lett. 97, 080401 (2006) method based on auxiliary functions. The advantage of 17. M. Greiner, O. Mandel, T. Esslinger, T.W. H¨ansch, I. this method is that it is local, and can be used easily if Bloch, Nature (London) 415, 39 (2002) the ground state is periodic in space. On the other hand, 18. O. Morsch, M. Oberthaler, Rev. Mod. Phys. 78, 179 (2006) the direct numerical computation of the superfluid frac- 19. K. Goral, L. Santos, M. Lewenstein, Phys. Rev. Lett. 88, tion using a galilean boost provides an alternative global 170406 (2002) method. We have demonstrated here that the two methods 20.T.Lahaye,C.Menotti,L.Santos,M.Lewenstein,T.Pfau, agree in 2D and that the semi-analytical method is more Rep. Prog. Phys. 72, 126401 (2009) accurate than the direct numerical simulation method. Us- 21. F. Cinti, P. Jain, M. Boninsegni, A. Micheli, P. Zoller, G. ing the second method, we have investigated the role of Pupillo, Phys. Rev. Lett. 105, 135301 (2010) defects in the superfluid fraction. For isolated random de- 22. N. Henkel, R. Nath, T. Pohl, Phys. Rev. Lett. 104, 195302 fects the superfluid fraction is only little increased but (2010) it is clearly enhanced when a grain boundary crosses the 23. E.P. Gross, J. Math. Phys. 4, 195 (1963) sample. Our results are in agreement with a slightly su- 24. L.P. Pitaevskiˇı, Sov. Phys. JETP 13, 451 (1961) perfluid increase inversely proportional to the length of 25. Y. Pomeau, S. Rica, Phys. Rev. Lett. 72, 2426 (1994) the mono-crystal present in the sample. 26. C. Josserand, Y. Pomeau, S. Rica, Phys. Rev. Lett. 98, 195301 (2007) 27. A.J. Leggett, J. Stat. Phys. 93, 927 (1998) It is our pleasure to thank Olivier Devauchelle and Peter Mason 28. C. Josserand, Y. Pomeau, S. Rica, Eur. Phys. J.S.T. 146, for their help. This work has been supported by the grant ANR- 47 (2007) 08-SYSC-004 of the Agence National de la Recherche, France 29. P.W. Anderson, Science 324, 631 (2009) (ANR COSTUME). N.S. acknowledges the financial support 30. D.T. Son, Phys. Rev. Lett. 94, 175301 (2005) of CONICYT (Chile) and Foundation Pierre-Gilles de Gennes 31. N. Sep´ulveda, C. Josserand, S. Rica, Phys. Rev. B 77, and S.R. also thanks the FONDECYT grant 1100289 (Chile). 054513 (2008) 32. A.J. Leggett, Science 305, 1921 (2004) 33. P.W. Anderson, W.F. Brinkman, D.A. Huse, Science 310, References 1164 (2005) 34. N. Prokof’ev, Adv. Phys. 56, 381 (2007) 1. A.F. Andreev, I.M. Lifshitz, Sov. Phys. JETP. 29, 1107 35. N.N. Bogoliubov, J. Phys. USSR 11, 23 (1947) (1969) 36. D.A. Kirzhnits, Yu.A. Nepomnyashchii, ZETF 59, 2203 2. L. Reatto, Phys. Rev. 183, 334 (1969) (1970) [Sov. Phys. JETP 32, 1191 (1971)] 3. G.V. Chester, Phys. Rev. A 2, 256 (1970) 37. Yu.A. Nepomnyashchii, Theor. Math. Phys. 8, 928 (1971) 4. A.J. Leggett, Phys. Rev. Lett. 25, 1543 (1970) 38. Yu.A. Nepomnyashchii, A.A. Nepomnyashchii, Theor. 5. E. Kim, M.H.W. Chan, Nature (London) 427, 225 (2004) Math. Phys. 9, 1033 (1971) 6. E. Kim, M.H.W. Chan, Science 305, 1941 (2004) 39. A. Aftalion, X. Blanc, R.L. Jerrard, Phys. Rev. Lett. 99, 7. E. Kim, M.H.W. Chan, Phys. Rev. Lett. 97, 115302 (2006) 135301 (2007) 8. A.S. Rittner, J.D. Reppy, Phys. Rev. Lett. 97, 165301 40. A. Bensoussan, J.L. Lions, G. Papanicolaou, Asymptotic (2006) Analysis in Periodic Structures (North-Holland, 9. A.S. Rittner, J.D. Reppy, Phys. Rev. Lett. 98, 175302 Amsterdam, 1978) (2007) 41. E.L. Andronikashvili, Zh. Eksp. Teor. Fiz. 16, 780 (1946) 10. Y. Aoki, J.C. Graves, H. Kojima, Phys. Rev. Lett. 99, 42. N.D. Mermin, H. Wagner, Phys. Rev. Lett. 17, 1133 (1966) 015301 (2007) 43. P.C. Hohenberg, Phys. Rev. 158, 383 (1967) 11. A. Penzev, Y. Yasuta, M. Kubota, J. Low Temp. Phys. 148, 677 (2007)