5
Imaginary-time nonuniform mesh method for solving the multidimensional Schrödinger equation: Fermionization and melting of quantum Lennard-Jones crystals Alberto Hernando * and Jiří Vaníček Laboratory of Theoretical Physical Chemistry, Institut des Sciences et Ingénierie Chimiques, École Polytechnique Fédérale de Lausanne, CH-1015 Lausanne, Switzerland (Dated: March 24, 2018) An imaginary-time nonuniform mesh method is presented and used to find the first 50 eigenstates and energies of up to five strongly interacting spinless quantum Lennard-Jones particles trapped in a one-dimensional harmonic potential. We show that the use of tailored grids reduces drastically the computational effort needed to diagonalize the Hamiltonian and results in a favorable scaling with dimensionality. Solutions to both bosonic and fermionic counterparts of this strongly interacting system are obtained, the bosonic case clustering as a Tonks-Girardeau crystal exhibiting the phe- nomenon of fermionization. The numerically exact excited states are used to describe the melting of this crystal at finite temperature. The multidimensional Schrödinger equation (MDSE) is undoubtedly one of the cornerstones of modern physics and much attention has been paid to developing effi- cient numerical methods for finding its solutions [1–18]. A very rich testing ground for such methods has been provided by the observation of new quantum phases at ultracold temperatures in finite and homogeneous sys- tems [18–23], and also by the development of optical lattices where ultracold atoms are trapped [24]. Due to their fascinating structural and dynamical proper- ties, special attention has been recently devoted to one- dimensional traps [25–30]. Indeed, in the strongly in- teracting (Tonks-Girardeau) regime of bosonic particles trapped in one-dimensional geometries, the repulsive na- ture of the atomic interaction at short distances gives rise to the phenomenon known as fermionization, the mech- anism of which is actively studied both theoretically and experimentally. Rigorous description and explanation of the new physics found in these well-controlled experiments re- quire accurate theoretical methods and constitute a formidable challenge [31], the main technical difficulty being the scaling of numerical algorithms with the num- ber of dimensions D. Indeed, standard algorithms for solving differential equations, such as the Finite Differ- ence method, scale exponentially with dimensions [32], making numerical solutions of many-dimensional prob- lems impracticable, if not impossible. Improved meth- ods addressing this difficulty in the case of station- ary states include the Discrete Variable Representation (DVR) [1], collocation method [2], phase-space method based on von Neumann periodic lattice [3], variational or diffusion quantum Monte Carlo (MC) methods [4, 18], Density Functional Theory (DFT) [5, 18], mean-field or pseudopotential interaction models [6–8], and many others. Some of these methods find only the ground state of the time-independent MDSE, using different ef- ficient techniques such as the imaginary time (IT) prop- agation [33] or the Variational Principle [34]. Meth- ods for real-time quantum dynamics include the Time- Dependent DVR [9], DFT [10], mean-field approaches [11], trajectory-based methods such as Bohmian dynam- ics [12], or time-dependent density matrix renormaliza- tion group (t-DMRG) method [13], which has proven to be very efficient in one-dimensional geometries. Despite many accomplishments in special cases, finding excited states and describing the real-time dynamics governed by a general high-dimensional Hamiltonian in the strongly interacting regime remains a difficult computational chal- lenge. In this paper we propose a novel general method, scal- ing favorably with dimensions, which is able to solve the time-independent MDSE numerically exactly and si- multaneously finds both its ground and excited states. Obviously, the proposed IT nonuniform mesh method (ITNUMM) is not intended to replace other well es- tablished approaches; instead we expect it to have a domain of applicability where other methods present more technical difficulties, such as in finding excited states of many-dimensional systems and where efficiency is more important than high accuracy. To show that ITNUMM achieves these goals, we apply it to find the wavefunctions of the first 50 states of an ensemble of up to five distinguishable Lennard-Jones (LJ) spinless particles trapped in a one-dimensional harmonic poten- tial in the Tonks-Girardeau regime. Once these states are obtained, we find, via symmetrization and anti- symmetrization, the solutions for the Bose-Einstein and Fermi-Dirac statistics, respectively, and observe fermion- ization in the bosonic case. We also show that the com- puted excited states can be used in a thermal average to describe the melting of the LJ clusters at finite tem- perature. As we use no other approximation than the numerical discretization of space and time, the obtained results are numerically exact. The derivation of our method starts by rewriting the time-dependent MDSE [34] i~ d dt |ψ(t)i = H|ψ(t)i, (1) arXiv:1304.8015v2 [quant-ph] 9 Jun 2013

École Polytechnique Fédérale de Lausanne, CH-1015 · PDF fileImaginary-time nonuniform mesh method for solving the multidimensional Schrödinger equation: Fermionization and melting

  • Upload
    vobao

  • View
    215

  • Download
    1

Embed Size (px)

Citation preview

Imaginary-time nonuniform mesh method for solving the multidimensionalSchrödinger equation: Fermionization and melting of quantum Lennard-Jones crystals

Alberto Hernando∗ and Jiří Vaníček†Laboratory of Theoretical Physical Chemistry, Institut des Sciences et Ingénierie Chimiques,

École Polytechnique Fédérale de Lausanne, CH-1015 Lausanne, Switzerland(Dated: March 24, 2018)

An imaginary-time nonuniform mesh method is presented and used to find the first 50 eigenstatesand energies of up to five strongly interacting spinless quantum Lennard-Jones particles trapped ina one-dimensional harmonic potential. We show that the use of tailored grids reduces drastically thecomputational effort needed to diagonalize the Hamiltonian and results in a favorable scaling withdimensionality. Solutions to both bosonic and fermionic counterparts of this strongly interactingsystem are obtained, the bosonic case clustering as a Tonks-Girardeau crystal exhibiting the phe-nomenon of fermionization. The numerically exact excited states are used to describe the meltingof this crystal at finite temperature.

The multidimensional Schrödinger equation (MDSE) isundoubtedly one of the cornerstones of modern physicsand much attention has been paid to developing effi-cient numerical methods for finding its solutions [1–18].A very rich testing ground for such methods has beenprovided by the observation of new quantum phases atultracold temperatures in finite and homogeneous sys-tems [18–23], and also by the development of opticallattices where ultracold atoms are trapped [24]. Dueto their fascinating structural and dynamical proper-ties, special attention has been recently devoted to one-dimensional traps [25–30]. Indeed, in the strongly in-teracting (Tonks-Girardeau) regime of bosonic particlestrapped in one-dimensional geometries, the repulsive na-ture of the atomic interaction at short distances gives riseto the phenomenon known as fermionization, the mech-anism of which is actively studied both theoretically andexperimentally.

Rigorous description and explanation of the newphysics found in these well-controlled experiments re-quire accurate theoretical methods and constitute aformidable challenge [31], the main technical difficultybeing the scaling of numerical algorithms with the num-ber of dimensions D. Indeed, standard algorithms forsolving differential equations, such as the Finite Differ-ence method, scale exponentially with dimensions [32],making numerical solutions of many-dimensional prob-lems impracticable, if not impossible. Improved meth-ods addressing this difficulty in the case of station-ary states include the Discrete Variable Representation(DVR) [1], collocation method [2], phase-space methodbased on von Neumann periodic lattice [3], variational ordiffusion quantum Monte Carlo (MC) methods [4, 18],Density Functional Theory (DFT) [5, 18], mean-fieldor pseudopotential interaction models [6–8], and manyothers. Some of these methods find only the groundstate of the time-independent MDSE, using different ef-ficient techniques such as the imaginary time (IT) prop-agation [33] or the Variational Principle [34]. Meth-ods for real-time quantum dynamics include the Time-

Dependent DVR [9], DFT [10], mean-field approaches[11], trajectory-based methods such as Bohmian dynam-ics [12], or time-dependent density matrix renormaliza-tion group (t-DMRG) method [13], which has proven tobe very efficient in one-dimensional geometries. Despitemany accomplishments in special cases, finding excitedstates and describing the real-time dynamics governed bya general high-dimensional Hamiltonian in the stronglyinteracting regime remains a difficult computational chal-lenge.

In this paper we propose a novel general method, scal-ing favorably with dimensions, which is able to solvethe time-independent MDSE numerically exactly and si-multaneously finds both its ground and excited states.Obviously, the proposed IT nonuniform mesh method(ITNUMM) is not intended to replace other well es-tablished approaches; instead we expect it to have adomain of applicability where other methods presentmore technical difficulties, such as in finding excitedstates of many-dimensional systems and where efficiencyis more important than high accuracy. To show thatITNUMM achieves these goals, we apply it to find thewavefunctions of the first 50 states of an ensemble ofup to five distinguishable Lennard-Jones (LJ) spinlessparticles trapped in a one-dimensional harmonic poten-tial in the Tonks-Girardeau regime. Once these statesare obtained, we find, via symmetrization and anti-symmetrization, the solutions for the Bose-Einstein andFermi-Dirac statistics, respectively, and observe fermion-ization in the bosonic case. We also show that the com-puted excited states can be used in a thermal averageto describe the melting of the LJ clusters at finite tem-perature. As we use no other approximation than thenumerical discretization of space and time, the obtainedresults are numerically exact.

The derivation of our method starts by rewriting thetime-dependent MDSE [34]

i~d

dt|ψ(t)〉 = H|ψ(t)〉, (1)

arX

iv:1

304.

8015

v2 [

quan

t-ph

] 9

Jun

201

3

2

where |ψ(t)〉 is the quantum state at time t of theD-dimensional system described by Hamiltonian H, interms of the quantum propagator K(q,q′; t − t′) :=〈q|e−i(t−t′)H/~|q′〉 in the position basis |q〉:

ψ(q, t) =

∫dq′K(q,q′; t− t′)ψ(q′, t′). (2)

Hamiltonian H := H0 + H1 is now split into two com-ponents: H0 is any Hamiltonian that includes the ki-netic energy operator T and whose matrix elements inthe q-representation are known, while H1 ≡ H1(q) isany many-body potential depending only on q. Forvery short time intervals t − t′ = ∆t, the time evolu-tion operator can be split to first order as e−i∆tH/~ =e−i∆tH0/~e−i∆tH1/~ +O(∆t2) and one can write

ψ(q, t′ + ∆t) =

∫dq′K0(q,q′; ∆t)e−i∆tH1(q′)/~ψ(q′, t′)

+O(∆t2), (3)

where K0(q,q′; ∆t) := 〈q|e−i∆tH0/~|q′〉 is the propaga-tor of H0, which is assumed to be known explicitly.

The |q〉 basis is discretized as∫dq|q〉〈q| = lim

N→∞

N∑j=1

w(qj)|qj〉〈qj | (4)

where w(qj) is a weight function depending on a partic-ular realization of the N states |qj〉. Indeed, w is definedas w(q) := [Np(q)]−1, where p(q) is the density distribu-tion of the qj . With this discretization, Eq. (3) becomes

ψ(qj , t′ + ∆t) = (5)

limN→∞

N∑k=1

w(qk)K0(qj ,qk; ∆t)e−i∆tH1(qk)/~ψ(qk, t′).

Since our main interest is finding the stationary statesof H, in the following we will assume that (i) ψ(q, t) =e−itEn/~ϕn(q) where ϕn(q) and En are the nth eigen-state and eigenenergy of the HamiltonianH, and that (ii)the evolution is performed in IT (t → −iτ). Althoughthe density p(q) is arbitrary, below we show that Eq. (5)simplifies in the IT scheme if this density corresponds tothe classical Boltzmann distribution of H1, namely if

p(q) = Z−1H1e−∆τH1(q)/~, (6)

where ZH1 = Tre−∆τH1/~ is a normalization constant(called configuration integral) and ∆τ/~ plays the role ofthe inverse temperature β. Under these conditions, Eq.(5) reads

e−∆τEn/~ϕn(qj) = (7)

limN→∞

ZH1

N

N∑k=1

K0(qj ,qk;−i∆τ)ϕn(qk).

By defining vector Φn := ϕn(qj)Nj=1, whose jth com-ponent is the wavefunction evaluated at position qj , andmatrix Kjk := K0(qj ,qk;−i∆τ)ZH1

/N whose elementsare proportional to the propagator K0 from qj to qk, onecan rewrite Eq. (7) as a matrix eigenvalue equation

e−∆τEn/~Φn = K · Φn. (8)

This equation, central to the ITNUMM, exhibits themain advantage of our method—the problem of findingthe spectrum and eigenfunctions of the original Hamil-tonian H is reduced to sampling the classical Boltz-mann distribution and diagonalizing K evaluated atthose points. Instead of the Hamiltonian, we diagonalizethe imaginary-time propagator, i.e., a matrix with an-alytically known and real-valued elements. Evaluationof, e.g., derivatives or Fourier transforms is not needed.Indeed, the implementation of the algorithm is rathersimple since it only requires standard methods for sam-pling from arbitrary probability distributions and diago-nalizing sparse real-valued matrices. The computationaleffort is also reduced by constructing a nonuniform gridin which more grid points are placed in areas where thewavefunctions exhibit more detailed features. In the spe-cial case of H0 ≡ T , H1(q) equals the classical potentialenergy, K0 is a free-particle propagator in D dimensions[34], and matrix elements Kjk assume the Gaussian form

Kjk =ZH1

N

( m

2π~∆τ

)D/2exp

[− m

2~∆τ(qj − qk)2

],

(9)where m is the mass, for simplicity assumed to be thesame for all degrees of freedom. In correlated sys-tems, where sampling the Boltzmann distribution is dif-ficult or unfeasible—as in the case of Coulomb interac-tion, we propose the splitting H1(q) = V1(q) + V2(q),where V1(q) is a sum of well-behaved one-body poten-tials and V2(q) is the remainder including all correlations.Here the sampling is performed with the weight p(q) =Z−1V1e−∆τV1(q)/~ and normalization ZV1

= Tre−∆τV1/~;the matrix to be diagonalized becomes

Kjk =ZV1

N

( m

2π~∆τ

)D/2× exp

[− m

2~∆τ(qj − qk)2 − ∆τ

~V2(qk)

]. (10)

We have found this method to be very efficient in one-dimensional problems with several very different poten-tials. Although an arbitrary sampling procedure can beused, we have employed a quadrature scheme: insteadof random sampling of p(q) by a MC procedure, the qjpoints are chosen with a deterministic algorithm. Themotivation for this approach is reducing to a minimumthe number of vector-elements needed for a given accu-racy, and thus reducing the computational cost of the di-agonalization of K. Specifically, we first consider a new

3

variable u, uniformly distributed in the interval [0, 1], anddefine an equidistant grid uj = (j − 1/2)/N . The Jaco-bian of the transformation from q to u is given by p(q)since p(q)dq = du, hence

u(q) =

∫ q

−∞dq′p(q′) = P (q), (11)

where P (q) is the cumulative distribution function. Next,the q-grid is obtained by inverting this equation for allvalues of uj , and once the q-grid is ready, the evaluationand diagonalization of the matrix K is performed withstandard numerical methods.

As the first application of ITNUMM, we solved (i)the 1D harmonic oscillator [34] H1(q) = mω2q2/2, us-ing natural units for energy and position (defined by~ω and

√~/mω, respectively), and (ii) two particles of

equal mass m interacting via a LJ potential H1(q) =VLJ(q) ≡ ε[(re/q)

12 − 2(re/q)6]. For the latter, we

used a de Boer quantum delocalization length [35] ofΛ = 21/6~/(re

√mε) = 0.16, corresponding to hypo-

thetical particles with properties between para-hydrogen—where quantum effects dominate—and neon—wherequantum effects are present but classical behavior dom-inates. In the Supplementary Material (SM), we showthe grid points, eigenvalues, and several eigenstates ob-tained with ITNUMM in both cases—we also include anotebook executable in the Wolfram Research’s Mathe-matica software, where the interested reader can explorethe technical details of the method.

As expected, we observed that the imaginary time ∆τmust be small enough to reduce the relative error σ in-troduced by the splitting of the propagator—which isσ ∼ O(∆τ3) since the second order term vanishes forstationary states—but large enough to avoid reducingthe Gaussian elements of the K matrix to delta func-tions and eventually obtaining a diagonal matrix. Thelatter condition is ensured by requiring 1 Kjj =

ZH1(m/2π~∆τ)

D/2/N , which imposes a lower bound on

∆τ for a given N . In the SM, we explore the dependenceof the relative error σ on ∆τ for a given number N ofgrid points, and also the dependence of σ on N in theharmonic oscillator. Remarkably, the relative error canbe fitted to σ(N) ' 0.18N−1.9, indicating a significantlyfaster convergence rate than the rate expected for a MCscheme [σ(N) ∼ N−1/2] [4, 18]. Regarding the excitedstates, we found that the error becomes large for stateswith the highest eigenenergies. Indeed, the number ofgrid points N becomes insufficient to reproduce the char-acteristic high frequency oscillations of wavefunctions de-scribing highly excited states. Yet, the agreement withexact results is very good for the first 150 states usingN = 500 grid points, as shown in Fig. 1 for the first 50states (the whole spectrum is shown in the SM).

As a more stringent test, we now apply the methodto D LJ particles in a one-dimensional harmonic trap.

D = 1

D = 2

D = 3

D = 4

D = 5

0 10 20 30 40 50

0

50

100

150

n

En

ÑΩ

-3 -2 -1 0 1 2 3

0

0.5

1

1.5

0.5

1

1.5

0.5

1

1.5

0.5

1

1.5

0.5

1

1.5

q re

Ρ 1

r e

Figure 1. (Color online) Left: Energy spectrum for D LJparticles in a 1D harmonic trap obtained with our method(circles). The exact results for D = 1 and 2 are shown asred solid lines. Energies are shifted to the minimum of thepotential min(H1) = −ε(D − 1)D/2. Right: One-body den-sities (normalized to the number of particles) of the groundstate for distinguishable (colored lines) and indistinguishableparticles (black lines).

Potentials V1 and V2 are defined by

V1(q) =

D∑λ=1

1

2mω2q2

λ, (12)

V2(q) =

D∑λ<µ

VLJ(|qλ − qµ|), (13)

the de Boer length has the same value as in the exampleabove, and ωre

√m/ε = 1/2. The problem is separable

only for D = 1 or 2, and so a multidimensional numeri-cal method is mandatory for D ≥ 3. In order to reducethe number of grid points in the numerical calculation,we first solve the problem for distinguishable particlesand construct a posteriori the eigenstates of indistin-guishable particles by symmetrizing or anti-symmetrizingthe wavefunction for spinless bosons or fermions, respec-tively. Thanks to the repulsive nature of the LJ potentialat short distances we only need to evaluate K in the sub-space defined by q1 > q2 + a, . . . , qD−1 > qD + a, wherea is the core radius of the LJ potential, within whichthe wavefunction is expected to be zero within numeri-cal accuracy (a = 0.63re in our calculations). The gridpoints are sampled from the classical Boltzmann distri-bution of the harmonic trap in this subspace, p(q) =

4

Z−1V1e−∆τmω2|q|2/2~ with ZV1

= (2π/∆τmω2)D/2/CD(a),where the normalization constant obeys CD(0) = D!. Allthe two-body interactions, contained in V2(q), are evalu-ated in the matrix elements of K. As mentioned above,only the low-lying eigenstates are accurate, so we haveused the Arnoldi algorithm [36] to obtain the first 50eigenstates. We have taken into account that many ofthe matrix elements are close to zero by using standardcomputational techniques for sparse matrices: instead ofstoring the N × N values of the matrix, only elementslarger than a certain threshold were stored. Parametersused in calculations with varying D were

D 1 2 3 4 5

N 500 6709 14 394 36 517 84 690

∆τ 0.0055 0.15 1.5 1.5 1.5

Note the relatively low total number of grid points neededto obtain results with reasonable accuracy (a relative er-ror of 0.002 for the D = 2 case). Figure 2 shows theground and 19th states for D = 2 and for the three statis-tics: distinguishable particles (in the above mentionedsubspace), bosons, and fermions (in the full space). Thespectrum of H as a function of D is shown in Fig. 1 (leftpanel). We find the same spectrum for the three cases,which is a consequence of the fermionization [30] mecha-nism due to the repulsive behavior of the LJ potential atshort distances. Indeed, the bosonic and fermionic sys-tems show the same one-body densities in position space,as shown in the right panel of Fig. 1. In all three cases thedensities show a well-defined structure, forming a quan-tum crystal. The displayed one-body densities, definedas [37]

ρn(qλ) =

∫|ϕn(q)|2

D∏µ 6=λ

dqµ, (14)

were obtained from the nonuniform mesh as follows:first,we computed its Fourier transform in a regularequidistant grid in momentum (k) space as

ρn(k) =

∫|ϕn(q)|2e−ikqλ

D∏µ=1

dqµ (15)

≈ Z

N

N∑j=1

e∆τV1(qj)/~−ikqλ |ϕn(qj)|2, (16)

and then Fourier-transformed ρn(k) back to qλ-space us-ing standard numerical methods.

The 50 states obtained in the course of the diagonal-ization are sufficient to study the behavior of the systemat finite temperatures. The (unnormalized) probabilitydistribution of the system pβ(q) at finite inverse temper-ature β is defined as the thermal average

pβ(q) =

∞∑n=1

e−βEn |ϕn(q)|2, (17)

-2 -1 0 1 2 -2 -1 0 1 2 -2 -1 0 1 2

-2

-1

0

1

2

-2

-1

0

1

2

q1re

q2

re

Figure 2. (Color online) Wavefunctions ϕn(q1, q2) of theground and 19th states for D = 2 LJ particles in a 1D har-monic trap (see text for details). Lighter (darker) color in-dicates positive (negative) values of the wavefunction. Left:distinguishable particles in the subspace q1 > q2; center: in-distinguishable bosons; right: indistinguishable fermions.

-3 -2 -1 0 1 2 3

0

0.5

1

1.5

2

q re

ΡΒ

r e

Figure 3. (Color online) One-body densities (normalized tothe number of particles) for D = 4 LJ particles in a 1D har-monic trap at three different temperatures: ~ωβ = 2.8 (dottedline), 0.7 (dashed line), and 0.4 (solid line). The crystal struc-ture disappears with increasing temperature, resulting in anunstructured total density as in a fluid.

and the corresponding one-body density ρβ(q) is obtainedsimilarly as for pure states. Figure 3 shows the one-bodydensity for D = 4 at three different temperatures, wherethe lack of structure at the highest temperature can beunderstood as the melting of the quantum crystal.

To summarize, we have presented compelling evidencethat the proposed method achieves the original goals.Indeed, (i) the only approximation used is the numeri-cal discretization of space and time; (ii) the ITNUMMonly requires standard methods for sampling from anarbitrary probability distribution and for diagonalizingreal-valued sparse matrices; (iii) both ground and ex-cited states are obtained in the course of the diagonal-ization; and (iv) due to the nonuniform nature of thegrid that uses the potential to guide the sampling, the

5

complexity of the algorithm is significantly reduced inhigh-dimensional systems. In particular, all our calcu-lations were performed on a single workstation with a64-bit 2.4 GHz Quad-Core Intel Xeon E5 processor and12 GB of memory. Yet, the algorithm can be easily ac-celerated by parallelization. The accuracy of ITNUMMcan be increased by using tailored grids, larger N values,or splitting methods of a higher order than in Eq. (3).In addition to computing thermal averages—as shownhere—the large set of excited states can be also used forsolving real-time quantum dynamics in a straightforwardfashion. As we have not found any a priori limitation tothe applicability of the method, other systems describedby the MDSE will be studied in the future.

Acknowledgments. The authors thank E. Zambrano,M. Wehrle, M. Šulc, and F. Mazzanti for discussions.This research was supported by the Swiss NSF NCCRMUST (Molecular Ultrafast Science & Technology) andby the EPFL.

[email protected][email protected]

[1] D. O. Harris, G. G. Engerholm, and W. D. Gwinn, J.Chem. Phys. 43, 1515 (1965).

[2] R. Kosloff, J. Phys. Chem. 92, 2087 (1988).[3] A. Shimshovitz and D. J. Tannor, Phys. Rev. Lett. 109,

070402 (2012).[4] L. Pollet, Rep. Prog. Phys. 75, 094501 (2012).[5] R. G. Parr and W. Yang, Density-Functional Theory of

Atoms and Molecules, 2 ed. (Oxford University Press,1994).

[6] F. Dalfovo, A. Lastri, L. Pricaupenko, S. Stringari, andJ. Treiner, Phys. Rev. B 52, 1193 (1995).

[7] I. G. Savenko, T. C. H. Liew, and I. A. Shelykh, Phys.Rev. Lett. 110, 127402 (2013).

[8] J. Rogel-Salazar, Eur. J. Phys. 34, 247 (2013).[9] G. Billing and S. Adhikari, Chem. Phys. Lett. 321, 197

(2000).[10] E. Runge and E. K. U. Gross, Phys. Rev. Lett. 52, 997

(1984).[11] M. Abid, C. Huepe, S. Metens, C. Nore, C. T. Pham,

L. S. Tuckerman, and M. E. Brachet, Fluid Dyn. Res.33, 509 (2003).

[12] R. E. Wyatt and C. J. Trahan, Quantum Dynamics withTrajectories: Introduction to Quantum Hydrodynamics,

1st ed. (Addison-Wesley, 2005).[13] G. Alvarez, L. G. G. V. D. da Silva, E. Ponce, and

E. Dagotto, Phys. Rev. E 84, 056706 (2011).[14] A. Tkatchenko, R. A. DiStasio, R. Car, and M. Scheffler,

Phys. Rev. Lett. 108, 236402 (2012).[15] M. A. Martín-Delgado and G. Sierra, Phys. Rev. Lett.

76, 1146 (1996).[16] J. F. Corney and P. D. Drummond, Phys. Rev. Lett. 93,

260401 (2004).[17] G. Torres-Vega and J. H. Frederick, Phys. Rev. Lett. 67,

2601 (1991).[18] J. M. McMahon, M. A. Morales, C. Pierleoni, and D. M.

Ceperley, Rev. Mod. Phys. 84, 1607 (2012).[19] P. Kapitza, Nature 141, 74 (1938).[20] A. J. Leggett, Phys. Rev. Lett. 29, 1227 (1972).[21] S. Grebenev, B. Sartakov, J. P. Toennies, and A. F.

Vilesov, Science 289, 1532 (2000).[22] J. R. Anglin and W. Ketterle, Nature 416, 211 (2002).[23] S. Balibar, Nature 464, 176 (2010).[24] I. Bloch, Nature Physics 1, 23 (2005).[25] J. P. Ronzheimer, M. Schreiber, S. Braun, S. S. Hodg-

man, S. Langer, I. P. McCulloch, F. Heidrich-Meisner,I. Bloch, and U. Schneider, Phys. Rev. Lett. 110, 205301(2013).

[26] M. Panfil, J. D. Nardis, and J.-S. Caux, Phys. Rev. Lett.110, 125302 (2013).

[27] P. Vignolo and A. Minguzzi, Phys. Rev. Lett. 110,020403 (2013).

[28] 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).

[29] T. Kinoshita, T. Wenger, and D. S. Weiss, Nature 440,900 (2006).

[30] B. Paredes, A. Widera, V. Murg, O. Mandel, S. Folling,I. Cirac, G. V. Shlyapnikov, T. W. Hansch, and I. Bloch,Nature 429, 277 (2004).

[31] I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys.80, 885 (2008).

[32] K. Morton and D. Mayers, Numerical Solution of Par-tial Differential Equations, An Introduction, 2 ed. (Cam-bridge University Press, 2005).

[33] V. S. Popov, Phys. Atom. Nuclei 68, 686 (2005).[34] J. J. Sakurai, Modern Quantum Mechanics , 1 ed.

(Addison-Wesley, 1993).[35] J. Deckman and V. A. Mandelshtam, J. Phys. Chem. A

26, 7394 (2009).[36] W. E. Arnoldi, Q. Appl. Math 9, 17 (1951).[37] E. Lipparini, Modern Many-Particle Physics: Atomic

Gases, Nanostructures and Quantum Liquids (World Sci-entific, 2008).