<<

arXiv:physics/0403146v1 [physics.comp-ph] 31 Mar 2004 † ∗ ehd hc a ecmr Padpoiea alternative inadequate. an is theoretic provide GP for and where GP need benchmark the can provides increases the which and This methods tune physics, to rich experimentally. way of interaction powerful source the a of as strength used [4 resonances been Fesbach successfully recently, More have 3]. performed sys [2, expected, weakly-interacting as well has, extremely For approach GP mean-field length. the tems scattering repuls or the attractive on either based potential, describe two-body well simple is a atoms interacting the by of among interaction consisting The real- gases atoms. dilute the kali are by ato These motivated ultracold [1]. devel- in was gases the condensation method Bose-Einstein in the this interest of systems. of Our ization many-boson use exact. and of principle opment state in is ground method the The study to method accurately more interaction of reliably. effect the metho describe computational can alternative that is develop It to increased. is necessary interactions fore particle especi of results, strength incorrect the wit to as lead approximate particl can only and of is approaches, these treatment effects correlation the or success, interaction remarkable their o fermion Despite bosons for (DFT) for theory equation density-functional (GP) Kohn-Sham the Gross-Pitaevskii the theo- as mean-field such simpler on ries based are studies quan mechanical computational tum Most theoreti Computational extract systems. to such on choice understanding of cal years. way the many been often have for methods field research challenging lcrncades [email protected] address: Electronic lcrncades [email protected] address: Electronic eea M ehd xs o acltn h properties the calculating for exist methods QMC Several (QMC) Carlo Monte quantum a present we paper this In very a been has systems quantum many-body of study The ueia tde.W hwrslsfrsseswt pto up with systems for results show fo We technique studies. back-propagation numerical the and soluti cap sampling non-interacting framework, importance of appro same ensemble GP the coherent the using stochastic, while and a GP method upon our improve between atically relation formal mean close (GP) Gross-Pitaevskii the corresponding the on with in interactions compare attractive with bosons ex untrapped with repu of compare tems We or attractive grid. an real-space a exa via basis principle interact single-particle in atoms is the method where the gas, and boson consisting used, permanents be Th of can basis space physics. particle the of cl in fields is walk several random and in branching applied approach, are field-theoretical which a methods on QMC based is method The unu ot al ehdfrtegon tt fmany-boso of state ground the for method Carlo Monte Quantum efruaeaqatmMneCro(M)mto o calculat for method (QMC) Carlo Monte quantum a formulate We .INTRODUCTION I. eateto hsc,TeCleeo ila n ay Will Mary, and William of College The Physics, of Department iaa Purwanto Wirawan Dtd a 1 2018) 21, May (Dated: there- ally mic and ive, hin al- ds al s. d a e ] - r - - ∗ n hwiZhang Shiwei and n.W ics aiu loihi sus including issues, algorithmic various discuss We ons. ieso,weeaayia ouin xs.W also We exist. solutions analytical where dimension, e emo ytm.I h otx faoi ae,Kat [7], Krauth gases, atomic of context Gruter the an In boson many- of systems. variety in a fermion fram to applied work first-quantized successfully the been which have in work, and methods, space [6] configuration particle (PIMC) path-integr finite-temperature Carlo diffu- the ground-state Monte and [5] The Carlo Monte systems. sion many-body interacting of ot al ehd[9 2 omn-oo ytm.We tr initial systems. an many-boson state from ground-state to boson many-body 22] the project [19, we paper, quant method this auxiliary-field Carlo In ground-state Monte nuclear theory. fermion 19], gauge the lattice 18, generalize and [17, 21], matter [20, condensed physics in systems fermion functions. ap- correlation and mean-field conven observables allows the to it of addition, relation In calculation later. formal discuss we close as proaches, many- with a particl method provides proper It body second- antisymmetry. the in or imposes Working symmetry permutation automatically reasons. is it several integral for quantization, frame- AFQMC appealing many-dimensional The is The means. work stochastic using 16]. computed one-body then over [15, integral many-dimensional auxil propagators a of into use fields, by transformed, iary are resultin field- interactions propagators a two-body is many-body from AFQMC where The method, 14]. theoretical [13, approach (AFQMC) Carlo be method. can QMC that fermion problem a spin-like with XXZ treated an into problem the transform of a tion allowed which hard-c assumed, a was calculation, potential repulsive latter the and In systems model. finite- spin Bose-Hubbard did the quantum [12] on Scalletar calculations and QMC Ulmke temperature als 11]. bosons, [10, trapped co- spheres of hard and state by Glyde ground two- the potential. studied the have hard-sphere workers modeling a by lengths, interactions scattering body positive with trapp of bosons properties finite-temperature study to PIMC ployed dnia igepril rias hc a rtsugges first was which orbitals, single-particle identical c eut nsalsses n rirrl-ie sys- arbitrarily-sized and systems, small in results act N sv otc w-oyptnil ecos sthe as choose We potential. two-body contact lsive t eilsrt hsmto ihatapdatomic trapped a with method this illustrate We ct. h FM ehdhsbe ieyepoe ostudy to employed widely been has method AFQMC The Monte quantum field auxiliary the on based is method Our optn bevbe,adilsrt hmwith them illustrate and observables, computing r fil acltosfrtapdaos n discuss and atoms, trapped for calculations -field ∼ fietclsnl-atceobtl.Aysingle- Any orbitals. single-particle identical of sl eae oeitn emo auxiliary-field fermion existing to related osely | 400 Ψ uigitrcinadcreainefcswith effects correlation and interaction turing rudsaepoeto sipeetda a as implemented is projection ground-state e c.Ormto rvdsawyt system- to way a provides method Our ach. tal. et T n h rudsaeo aybsnsystems. many-boson of state ground the ing i bosons. u hieof choice Our . 8,adHlmn n ruh[]hv em- have [9] Krauth and Holzmann and [8], asug igna23187 Virginia iamsburg, † | Ψ T i sapraetcnitn of consisting permanent a is systems n e in ted ient um ore ial ed N e- e- a- al o d g - 2 a model calculation by Sugiyama and Koonin [14]. The many- teristics of our method. We carry out GP calculations on the body is projected from ΨT with open-ended, same Hamiltonian and compare the results with those from branching random walks to sample the| auxiliaryi fields. We our QMC calculations. In section VI we comment on some formulate an importance sampling scheme, which greatly im- characteristics of the method, further discuss its relation to proves the efficiency of the method and makes possible sim- and differences from GP, and mention future directions and ulations of large systems. We also discuss in detail the back- some immediate applications of this method. Some comput- propagation technique which allows convenient calculationof ing issues will also be discussed. Finally, in the appendices virtually any ground-state observables. we provide additional technical details of the method. Our method retains all the advantages of AFQMC. It al- lows the use of any single-particle basis, which in this paper is chosen to be a real-space grid. As we discuss in Sec. VI, it II. BACKGROUND provides a means for true many-body calculations in a frame- work which closely relates to the GP approach. The approach A. Many-body Hamiltonian can be viewed as a stochastic collection of parallel GP-like calculations whose “coherent” linear combination gives the We use the second quantized formalism throughout this pa- interaction and correlation effects. per. We assume that an appropriate set of single-particle basis In this paper we present our QMC method for bosons and χi has been chosen, in terms of which the wave func- discuss its behavior and characteristics. We use a trapped tions{| i} will be expanded. For simplicity, we assume that the atomic boson gas as our test system, where the atoms inter- single-particle basis is orthonormal, although this is not re- act via an attractive or repulsive contact two-body potential. quired. The number of basis states is M. The operators ci† and A sufficiently detailed description of the method is given to ci, respectively, are the usual creation and annihilation oper- facilitate implementation. Compared to its fermionic counter- ator for the state χi . They satisfy the commutation relation part, our method here is formally simpler. It therefore also | i [ci,cj†] = δij . This automatically imposes the symmetriza- offers opportunities to study algorithmic issues. Because of tion requirement− of the many-body wave functions. the intense interest in methods for treating correlated systems We limit our discussion to a quantum-mechanical, many- (fermions or bosons) and the relatively early stage of this type body system with two-body interactions. The Hamiltonian Hˆ of QMC methods, a second purpose of the paper is to use has a general form of the bosonic test ground to explore, discuss, and illustrate the generic features of ground-state QMC methods based on aux- ˆ ˆ ˆ (1) iliary fields. An example is the case of repulsive interactions, H = K + V, where a phase problem appears in a bosonic system, which ˆ provides a clean test ground to study methods for controlling where K is the sum total of all the one-body operators (the this problem [22], which is crucial for applications in fermion kinetic and external potential energy), systems. The majority of the applications in this paper will be ˆ to systems where exact results are available for benchmark. K = Kij ci†cj ; ij These include small systems, which can be diagonalized ex- X actly, and the case of untrapped bosons with attractive inter- actions in one dimension, where analytical solutions exist. It and Vˆ contains the two-body interactions: is worth emphasizing that the method scales gracefully (sim- ˆ ilar to GP) and allows calculations for a large number (N) V = Vijklci†cj†ckcl . of bosons. We will show results for larger systems ( 1000 ijkl X sites and hundreds of particles) in one- and three-dimensio∼ ns to illustrate this. Our objective is to calculate the ground state properties of Our paper is organized as follows. In section II, we es- such a system, which contains a fixed number of particles, tablish some conventions and review the basic ground-state N. projection and auxiliary-field method. In section III, we introduce our new AFQMC implementa- tion for bosons, including the formulation of an importance- B. Ground state projection sampling scheme and the back-propagationtechnique for con- venient calculation of virtually any ground-state observables. The ground state Φ can be readily ex- | 0i In section IV, we describe the implementation of our method tracted from a given trial solution Ψ using the ground-state | Ti to study the ground state of a trapped Bose atomic gas, which projection we model by by a Bose-Hubbard Hamiltonian with an exter- ∆τHˆ ∆τET nal trapping potential. We also describe our implementation gs e− e , (2) of the GP approach to study the same Hamiltonian. In section P ≡ V, we present our computational results. We benchmark the where ET is the best guess of the ground-state energy, pro- method in systems where exact results are available. We also vided that ΨT is not orthogonal to Φ0 . Applying the oper- provide examples to illustrate the behavior and key charac- ator repeatedly| i to the initial wave| functioni Ψ would Pgs | Ti 3 exponentially attenuate the excited-state components of the where vˆ is a one-body operator: initial wave function, leaving only the ground state:

n n ( gs) ΨT →∞ Φ0 ; (3a) vˆ vˆ c†c . P | i −→ | i ≡ ij i j Φ Φ . (3b) ij Pgs| 0i −→ | 0i X Because of its resemblance to the real-time propagator, the operator is also called the imaginary-time propagator. In The hermiticity of Vˆ allows us to decompose it into a sum Pgs ground-state QMC methods, gs is evaluated by means of a of the square of one-body operators vˆi (see, for example, Monte Carlo sampling, resultingP in a stochastic representation Refs. 19 and 25): { } of the ground-state wave function.

Vˆ = 1 vˆ2 , − 2 i (6) C. Basic auxiliary-field method i X Two essential ingredients are needed in order to evaluate Because of this, we can always apply the Hubbard- gs within a reasonable computing time. The first is the Trotter-SuzukiP approximation [23, 24]. The propagator is Stratonovich transformation on a general two-body potential broken up into a product of exponential operators, which be- operator: comes exact in the limit ∆τ 0. The second-order form of → this approximation is 1 2 ∆τVˆ ∆τvˆi 2 e− = e 2 + (∆τ ) O ∆τ(Kˆ +Vˆ ) 1 ∆τKˆ ∆τVˆ 1 ∆τKˆ i e− = e− 2 e− e− 2 Y 1 2 (7) (4) 2 xi 3 ∞ e− xi√∆τ vˆi 2 + (∆τ ) . = dxi e + (∆τ ) . O √2π O i Z−∞ The second ingredient is the Hubbard-Stratonovich (HS) Y transformation [15, 16], which allows us to reduce the two- body propagator to a multidimensional integral involving only In general, the Trotter breakup incurs an additional systematic one-body operators, using the following identity: [25] error of (∆τ 2). O Applying these two procedures, we obtain an approximate 1 ∆τvˆ2 1 ∞ 1 x2 x√∆τ vˆ e 2 = dx e− 2 e , (5) expression of the ground-state projection operator: √2π Z−∞

1 1 ∆τE ∆τKˆ ∞ xi√∆τ vˆi ∆τKˆ 2 = e T e− 2 dx p(x )e e− 2 + (∆τ ) , (8) Pgs · i i O ( i ) Y Z−∞

where p(x) is the normalized Gaussian probability density Bˆv(~x) : a product of the exponential one-body op- 1 1 x2 • function with unit standard deviation: p(x) e 2 . erators arising from the auxiliary-field transformation. √2π − ≡ ˆ xi√∆τ vˆi This approach is applicable to both boson and fermion sys- From Eq. (8), Bv(~x) e . ≡ i tems. It enables us to compute the exact ground state of a quantum many-body system. To reduce the systematic er- Q Bˆ(~x) : the product of Bˆ (~x) with all other one-body ror from the finite timestep ∆τ, the so-called “Trotter er- • v ror”, small timesteps ∆τ are necessary. Often, calculations exponential operators that do not depend on the aux- are performed for several ∆τ values, then an extrapolation to iliary fields ~x, and all the necessary scalar prefac- ˆ ∆τET ∆τ 0 is made to remove the Trotter error. tors. For the projector in Eq. (8), B(~x) e → 1 ∆τKˆ 1 ∆τKˆ ≡ · For convenience we define the following notations: e− 2 Bˆv(~x) e− 2 .

~x x1, x2,... : the collection of all the auxiliary- • ≡ { } With these notations, gs takes a generic form of a high- fields. dimensional integral operator:P

p(~x) i p(xi) : a (normalized) multidimensional • probability≡ density function, which is the product of the gs d~xp(~x)Bˆ(~x) . (9) one-dimensionalQ probability density functions . P ≈ p(xi) Z 4

D. Wave function representation important virtue of this representation is that the exponen- tial of a one-body operator Aˆ transform a single-permanent We write our wave functions in terms of the basis functions wave function φ into another single-permanent wave func- | i χ . A single-particle wave function is written as tion φ′ : [26] | ii | i

Aˆ ϕ = ϕi χi = ϕici† 0 ϕˆ† 0 . e φ = φ′ . (12) | i | i | i≡ | i (10) i i | i | i X X A single-permanent, N-Bosons wave function is given by In particular, Bˆ(~x) in Eq. (12) transforms a single permanent φ into another single permanent φ′ . (In Appendix A we ˆ ˆ ˆ include| i a brief summary of properties| i of wave functions in φ = φ1† φ2† ... φN† 0 . (11) | i | i IOR.) In general, the exact ground state wave function is a superpo- sition of such permanents. Unlike the fermionic case, where the particles occupy mutually orthogonal orbitals, there is no such restriction on the orbitals here. We use this freedom in E. Metropolis AFQMC our method to have all the bosons occupy the same orbital in φ , which greatly simplifies the computation [14]. We will re- Standard AFQMC calculations [14] employ Metropolis fer| i to this as identical orbital representation (IOR). The most Monte Carlo to compute various ground-state observables,

Ψ Aˆ Ψ Aˆ = h T|Pgs ···Pgs Pgs ···Pgs| Ti h ig.s. Ψ Ψ h T|Pgs ···Pgs| Ti ( ~x , ~y ) P ( ~x , ~y ) Ψ Bˆ(~x ) Aˆ Bˆ(~y ) Ψ = D { m n} { m n} h T| m m n n | Ti ( ~x , ~y ) P ( ~x , ~y ) Ψ Bˆ(~x ) Bˆ(~y ) Ψ (13) R D { m n} { m n} h TQ| m m Qn n | Ti η( ~x ) Aˆ φ( ~y ) R ( ~x , ~y ) P ( ~x , ~y ) η( ~x Q ) φ( ~y )Qh { m} | | { n} i D { m n} { m n} h { m} | { n} i η( ~x ) φ( ~y ) = h { m} | { n} i , R ( ~x , ~y ) P ( ~x , ~y ) η( ~x ) φ( ~y ) D { m n} { m n} h { m} | { n} i R where are several advantages in implementing the Monte Carlo sam- pling as a random walk process. It is a true ground-state for- ( ~xm, ~yn ) m d~xm n d~yn , malisms with open-ended random walks which allow projec- D { } ≡ tion to long enough imaginary-times. The sampling process P ( ~xm, ~yn ) p(~xm) p(~yn) , { } ≡ Qm Q n can be made much more efficient than in standard AFQMC, and in the last line we have introducedQ theQ shorthand by virtue of importance sampling with ΨT to guide the ran- dom walks. It also leads to a universal approach for bosons ˆ and fermions, where it is necessary to use the random walk η( ~xn ) ΨT n B(~xn); h { } | ≡ h | formalism in order to implement a constraint to deal with the φ( ~y ) Bˆ(~y ) Ψ . | { m} i≡ m Q m | Ti sign and complex-phase problems [19, 22]. Q The Metropolis simulation is carried out by sampling the A key observation is that we can choose an IOR single- probability density function defined by the integrand in the permanent wave function as the initial wave function Ψ . | Ti denominator. Given the choice of ΨT in the identical-orbital At each imaginary timestep τ n ∆τ in the projection in representation, this readily applies to bosons, which is howthe Eq. (3), the wave function is stochastically≡ sampled by a col- model calculation by Sugiyama and Koonin [14] was done. (τ) lection of single-permanent wave functions φ , where The total length of the imaginary time is predetermined by i the index i (in Cursive letter) is different from{| thei} basis in- ˆ ∆τ and the number of B operators in the product. dex i. From Eqs. (9) and (12), we see that, with each walker (0) φi initialized to ΨT in IOR, the resulting projection will |leadi to a superposition| i of single-permanent wave functions, III. NEW METHOD FOR BOSONS all of which are in IOR.

In this paper we formulate a new approach for ground-state Each permanent evolves by the stochastic application of calculations of bosons with branching random walks. There , as follows: we randomly sample ~x from the probability Pgs 5 density function p(~x), then apply Bˆ(~x) on φ(τ) : (outside of the statistical errorbar) and converges to the exact | i i value as Nwlkr is increased. In practice, however, the use- (τ+∆τ) ˆ (τ) fulness of the ‘brute force” estimator is limited to smaller φi B(~x) φi , (14) | i ← | i systems. In general it will have large variances. Reducing We will call these permanents random walkers. The collection the variance is expensive because Aˆ scales as (N 2 ), h ibf O wlkr of these random walkers at each imaginary-time step is also where Nwlkr is the size of the population used to represent referred to as population. Φ0 . The population must first be equilibrated so that the ground- | Thei simplest approach to measuring the observables is the state distribution is reached. After equilibrium the ground mixed estimator, i.e. state is given stochastically by the collection of single- ψ Aˆ Φ permanent wave functions φi : ˆ T 0 (18) {| i} A mix = h | | i . h i ψ Φ0 . h T| i Φ = φi . | 0i | i (15) For example, to compute the ground-state energy, we can in- i X troduce the so-called local energy EL[ψT, φ]: Measurement of ground-state observables can then be carried ˆ out. ψT H φ EL[ψT, φ]= h | | i (19) The random walk process naturally causes the walker’s or- ψ φ h T| i bitals to fluctuate. In order to increase sampling efficiency, The ground state energy is obtained from the weighted sum of we may associate a weight factor wi to each walker φi . For the local associated with each walker: example, we can use the walker’s amplitude as the weight| i fac- tor: i ψT φi EL[ψT, φi] Emix = h | i (20) i ψT φi wi φi φi . P h | i ≡ h | i The local energy for each walkerP can be computed using the A better definition of the weightp will be introduced later when formula given in Appendix A. we discuss importance sampling. We duplicate a walker when The mixed estimator in Eq. (18) is exact only if the operator its weight exceeds a preset threshold. Conversely, walkers Aˆ commutes with the Hamiltonian. Otherwise, a systematic with small weight (lower than a predetermined limit) should error arises. Nonetheless the mixed estimator often gives an be removed with the corresponding probability. In this way, improvement over the purely variational estimator: the walkers will have roughly the same weight. This results in a branching random walk. Ψ Aˆ Ψ Aˆ h T| | Ti . (21) h iT ≡ Ψ Ψ h T| Ti A. Measurement: “brute force” and mixed estimators Two formulas are often employed to correct for the systematic error: The ground-state value of an observable ˆ is its expectation A Aˆ extrap1 2 Aˆ Aˆ ; (22) value with the ground-state wave function: h i ≡ h imix − h iT ˆ 2 ˆ A mix ˆ A extrap2 h i . (23) ˆ Φ0 A Φ0 h i ≡ Aˆ A g.s. = h | | i . (16) h iT h i Φ0 Φ0 h | i The second formula is useful for quantities such as density In principle, we can use the same Monte Carlo samples as both profile, where it must be nonnegative everywhere. These cor- Φ and Φ . A “brute force” measurement on population rections are good only if Ψ does not differ significantly h 0| | 0i | Ti φ(τ) at imaginary-time τ is then given by from Φ0 . In general, we need the back-propagation scheme {| i i} to recover| i the correct ground-state properties. We will de- (τ) (τ) scribe this method after introducing importance sampling. (τ) φ Aˆ φ Aˆ ijh j | | i i (17) h ibf ≡ (τ) (τ) P ij φj φi h | i B. Importance sampling ˆ P and the estimator A bf is the average of such measurements. The “brute force”h estimatori is not useful in real-space based In practice, the efficiency of the bare random walk de- QMC methods such as diffusion Monte Carlo, because the scribed earlier is very low, because the random walks “ran- overlaps between different walkers would lead to δ-functions. domly” sample the Hilbert space, and the weights of the walk- Here the walkers are non-orthogonal mean-field wave func- ers fluctuate greatly. This results in large statistical noise. We tions, and Eq. (17) is well defined in principle. The estima- formulate an importance sampling procedure [19, 22]—using tor is exact for all observables in the limit of large Nwlkr. the information provided by the trial wave function ΨT — The ground-state energy estimated in this way is variational, to guide the random walk into the region where the expected| i namely, the computed energy lies higher than the exact value contribution to the wave function is large. 6

1. Importance-sampled random walkers that the walkers propagate in the following manner:

φ(τ+∆τ) φ(τ) An importance-sampled walker also consists of a perma- w(τ+∆τ) | i i w(τ) | i i . (26) nent and a weight, although the weight will be redefined ac- i (τ+∆τ) ←− i (τ) ΨT φi ΨT φi cording to the projected overlap of the permanent with the h | i h | i trial wave function. The purpose is to define a random walk From this requirement comes the second part of process which will lead to a stochastic representation of the the modified propagator, which is the overlap ratio ground-state wave function in the form (τ+∆τ) (τ) ΨT φi / ΨT φi . This factor is obtained by h | i h | (τi+∆τ) bringing the term ΨT φi in Eq. (26) to the right-hand . φi side. It depends onh Ψ| and thei specific path in auxiliary- Φ0 = wi | i , (24) | Ti | i ΨT φi field space, and will “guide” the random-walk toward the Xi h | i region where Ψ φi is large. h T| i where wi is the new weight of the walker. The overlap enters Combining the two parts gives an importance-sampled to redefine the weight factor such that walkers which have propagator of the form large overlap with ΨT will be considered “important” and will tend to be sampled| i more. Such walkers will also have gs[φ] d~xp(~x)W (~x, φ)Bˆ(~x ~x) , (27) greater contributions in the measured observables. Since the P ≈ − Z permanent now appears as a ratio φ / Ψ φ , its normal- i T i where e ization is no longer relevant and can| i beh discarded,| i unlike in the unguided random walk. The only meaningful information ˆ ΨT B(~x ~x) φ ~x ~x 1 ~x ~x in φi is its position in the permanent space. W (~x, φ) h | − | i e · − 2 · (28) | i ≡ Ψ φ h T| i is the aggregate of all the scalar prefactors in the modi- (τ) (τ) 2. Modified auxiliary-field transformation fied propagator. This propagator takes wi , φi and (τ+∆τ) {(τ+∆τ)| i} advances the population to wi , φi , both of Now we describe the random walk process for the modified which represent Φ in the form{ of Eq. (24).| i} | 0i walkers. The goal is to modify gs in Eq. (9) such that the Monte Carlo sampling of the new propagator is similar P gs random walk process leads to random walkers with the char- to the one without importance samping. We sampleP ~x from a acteristics described above in Eq. (24). The basic idea is the normal Gaussian distribution, and apply the operatore Bˆ(~x ~x) same as that in Ref. 19. The main difference is that here we (τ) − to the current walker φi . But now we accumulate an extra are dealing with bosons. In addition the HS fields in Ref. 19 | i (τ) are discrete Ising-like, which allowed simplifications in the multiplicative weight factor W (~x, φi ) every time we apply importance sampling, while here the auxiliary fields are con- Eq. (27): tinuous and thus a more general formalism will be developed. φ(τ+∆τ) Bˆ(~x ~x) φ(τ) (29a) Our mathematical derivation here follows that of Ref. 22. Up | i i ← − | i i (τ+∆τ) (τ) (τ) to now we have assumed that ΨT φi is real and positive. w W (~x, φ ) w . (29b) There is therefore no additionalh subtlety| i with the meaning of i ← i i importance sampling and the correct form of the overlap to Here we use the customary notation of vector dot product, e.g. use, which Ref. 22 addressed in the context of fermionic cal- ~x ~x x x . Note that the weight factor W (~x, φ(τ)) de- culations with general interactions. · ≡ i i i i pends on both the current φ(τ) and future φ(τ+∆τ) walker To derive the importance-sampled propagator, we plug P i i positions. Eq. (24) into Eq. (3b). We will focus on the two-body prop-   agator, which is evaluated stochastically and is therefore af- fected by importance sampling in a non-trivial way. 3. The optimal choice for auxiliary-field shift ~x The modified propagator, gs, consists of two parts. The first part is the transformationP introduced in Eq. (5), which we The optimal importance sampling is achieved when each now rewrite in the following form: e random walker contributes equally to the estimator. We there- fore choose ~x to minimize the fluctuation in the weight factor 1 ∆τvˆ2 1 ∞ 1 x2 xx 1 x2 √∆τ (x x)ˆv e 2 = dx e− 2 e − 2 e − , (25) wi. The fluctuation in wi will be minimized if we minimize √2π the fluctuation in the prefactor Eq. (28). We do so by requiring Z−∞ the partial derivatives of this prefactor to vanish with respect where we have added an arbitrary shift to the auxiliary field x to xi at its average (xi =0): x in the auxiliary-field operator. This is a change of variable in the integral on the right-hand side and does not alter the ∂ Ψ Bˆ(~x ~x) φi 1 T ~x ~x 2 ~x ~x result of the integral. The new propagator gs must preserve h | − | i e · − · =0 . P ∂xi " ΨT φi × # the representation of Φ0 in the form of Eq. (24); this dictates h | i xi=0 | i e

7

It is sufficient to expand the exponentials in terms of ∆τ and C. Measurement: back propagation require the term linear in xi to vanish, since this is the leading term, containing √∆τ. The others contain higher-order terms With importance sampling, the mixed estimator in Eq. (18) and are vanishingly small as ∆τ 0. The best choice for xi is given by: that satisfies this requirement is → ˆ ΨT A φi √ ΨT vˆi φi √ wi h | | i xi = ∆τ h | | i ∆τ v¯i . (30) Ψ φi i T − ΨT φi ≡− ˆ h | i (34) h | i A mix = X . This choice depends on the current walker position as well h i wi as Ψ , which is to be expected, since the objective for the i | Ti X shift is to guide the random walk toward the region where For example, the ground-state energy is (τ) ΨT φi is large. With ~x determined, the algorithm for the h | i wiE [ψ , φi] random walk, as given in Eq. (29), is now completely speci- E = i L T . mix w fied. P i i As mentioned earlier, the normalizationP of φi is irrelevant be- cause φi only appears in ratios in any formula that defines the 4. Local energy approximation algorithm: Eqs. (24), (28), (30), (33), and Eq. (34). We can (and should) normalize the permanent as needed, and discard We can furthermore approximate the prefactor W (~x, φ) in the resulting normalization factor. Eq. (28) to obtain a more elegant and compact expression. The mixed estimator is often inadequte for computing ob- After rewriting the prefactor in the form of an exponential, servables whose operators do not commute with the Hamilto- expanding Bˆ(~x ~x) in terms of ∆τ, and ignoringterms higher nian. In some cases the error due to this noncommutation is than (∆τ) in− the exponent, we obtain unacceptable. For example, the condensate fraction in the at- O 1 2 2 2 1 2 tractive trapped Bose-Hubbard model is greater than 100% if ∆τ(1 xi )(¯vi v ) ∆τv e 2 − − i e 2 i , (31) the Green’s function ci†cj is estimated using the mixed esti- i h i Y mator. Therefore we have to propagate the wave functions on where both the right- and the left-hand side of the operator: 2 ΨT vˆi φi ˆ 2 τbpH ˆ vi h | | i . (32) ΨT e− A Φ0 ≡ Ψ φi Aˆ = h | | i . (35) h T| i bp τ Hˆ h i Ψ e− bp Φ The product is over the basis index i, which should be dis- h T| | 0i tinguished from the walker index i. The latter is held fixed This estimator approaches the exact expectation value in here. The first exponential in Eq. (31) can be ignored by not- Eq. (16) as τbp is increased. Zhang and co-workers proposed 2 ing that the average value of xi with respect to the Gaussian a back-propagation technique [19] that reuses the auxiliary- 2 probability density function is unity. Setting xi 1, i.e., field “paths” from different segments of the simulation to ob- 2 → ˆ evaluating the exponential at the mean value x , is justi- bp τbpH 2 i tain Φ0 ΨT e− , while avoiding the Nwlkr scaling 2 2 h i h | ≡ h | fied because v¯i and vi do not change drastically within one of a brute-force evaluation with two separate populations for 2 timestep. We also note that v = Ψ Vˆ φ / Ψ φ , Φ0 and Φ0 . Here we give a more formal derivation and i i T i T i h | | i which is the mixed-estimator of the potential−h | energy| i h with| rei- description of the technique, and implement it to bosons. P (τ) spect to the walker φi . Combining this term with the similar At imaginary-time τ, the population is φ , which rep- | i {| i i} contribution from the kinetic propagator, we obtain a simple, resents Φ0 in the form of Eq. (24). The propagator in the approximate expression for Eq. (28): denominator| i can be viewed equivalently as operating on the

(τ) ∆τ(E E [Ψ ,φi]) left or the right. The latter view is precisely the “normal” W (~x, φ ) e T− L T , (33) i ≈ importance-sampled random walk from τ to the future time where EL[ΨT, φi] is the local energy of φi as defined in τ ′ τ + τbp, which consists of nbp τbp/∆τ steps. Eq. (19). Note that, contrary to Eq. (28), this form depends We≡ first assume that there is no branching≡ (birth/death of only on the current walker position and not the future, al- walkers), i.e., the weights are fully multiplied according to though in practice a symmetrized version can be used which Eq. (28). The random walk of each walker will generate a replaces the local energy by the average of the two. For a path in auxiliary-field space. For convenience we will denote good trial wave function, the local energy fluctuates less in the the path-dependent operator Bˆ[~x(τ) ~x(φ(τ))] by Bˆ(τ), and i − i i random walk. If the trial wave function is the exact ground- weight factor W (~x(τ), φ(τ)) by W (τ). Further we will denote state wave function, the local energy becomes a constant and i i i the time-ordered product of Bˆ(τ) from imaginary-time τ to the weight fluctuation is altogether eliminated. This bears a ′ i (τ :τ) (τ) close formal resemblance to the importance-sampled difus- τ ′ by Bˆ , and correspondingly the product of W by ′ i i sion Monte Carlo method. W (τ :τ). Each path defines a product The algorithm resulting from Eq. (33) is an alternative to i ′ ′ Eq. (28). The two are identical and exact in the limit ∆τ 0, 1 (τ :τ) (τ :τ) (τ) ′ W Bˆ Ψ φ . → (τ ) i i T i (36) but can have different Trotter errors. Ψ φ h | i h T| i i 8

Collectively these products give a stochastic representation of (τbp) 3. the population ηi is then generated by back- τ Hˆ {| i} e− bp . propagation using Eq. (39); τ Hˆ Replacing the operator e− bp in the numerator and de- nominator of Eq. (35) with Eq. (36), and using the expression 4. this population is matched in a one-to-one manner to (τ) for Φ given by Eq. (24), we obtain φ , weighted by the weight at the later time, 0 i′ | i {|(τ ) i} ′ ′ wi , and the estimator is formed. 1 (τ :τ) (τ :τ) (τ) (τ) ′ ˆ ˆ i ΨT (τ ) Wi Bi A wi φi h | ΨT φi | i In the back-propagation the propagators are, as shown in Aˆ = h | i ′ ′ . bp 1 (τ :τ) ˆ(τ :τ) (τ) (τ) h i P Ψ τ′ W B w φ Eq. (39), idential to those in the forward direction, but in re- i T Ψ φ( ) i i i i h | h T| i i | i verse order in imaginary-time. As in the normal walk, the nor- (37) (τ ) P malization of η bp does not enter in the estimator. Similar | i i Using the propagation relation in Eq. (29), we can show that to the mixed estimator, this procedure can be repeated period- ically to improve statistics. Evidently this estimator is exact ′ ′ ′ ′ in the limit of large τ . Bˆ(τ :τ)W (τ :τ) w(τ) φ(τ) = w(τ ) φ(τ ) , (38) bp i i i | i i i | i i We have assumed that there is no branching within the in- ′ terval . In practice, a population control scheme is often (τ ) τbp i.e., the denominator in Eq. (37) reduces to i wi . This used which causes birth/death of walkers. This does not affect resultis to be expected,and can also beseen by completingthe the derivation above or the basic algorithm. The effect on the P nbp steps of the “normal” random walk we discussed above. implementation is that a list of ancestry links must be kept for With importance sampling, the Monte Carlo estimate of the the forward steps, which indicates the parent of each walker denominator is simply given by the weights at time τ ′. at each step in the imaginary-time duration τbp. As a resultof To simplify the numerator we associate a back-propagated branching, two or more η ’s may share the same segment of (τ) h | (τ) wave function with each walker φi the paths in their “past” and the same parent φ . The esti- | i | i i mator remains exact for large τbp. Branching or weight fluctu- (τ ) (τ+τ : τ) η bp Bˆ bp † Ψ . (39) ation does have a more serious practical implication, however. | i i≡ i | Ti As τbp is increased, more and more η ’s will be traced back h i (τ) h | Note that each of these η’s originates from the trial wave func- to the same parent φ . Or equivalently, fewer and fewer | i i tion Ψ , and is propagated by applying the Bˆ’s in reverse permanents in the set φ(τ) will contribute to the estimator. | Ti i order, as implied by the Hermitian conjugation. We may then This results in a loss{| of efficiencyi} or an increase in variance. write Eq. (37) in the following form: Better importance sampling will help improve the situation, often greatly, by reducing fluctuations in weights, although (τ ) (τ) ′ bp ˆ the problem will always occur at large enough τ . In ourap- (τ ) ηi A φi bp wi h | | i plications to date we have rarely encountered the problem and (τbp) (τ) i ηi φi (40) find that the computed observables convergequite rapidly (see Aˆ = X h ′ | i . h ibp (τ ) section V for illustrative results). wi Xi The estimators in Eqs. (35) and (40) parallel that of the stan- IV. TRAPPED BOSON GAS: MODEL AND dard AFQMC estimator in Eq. (13). The φ ’s and η ’s have IMPLEMENTATIONS OF QMC AND GP METHODS similar meanings. The only difference lies| i in howh the| paths are generated. Here an open-endedrandomwalk is used to ad- In this section we discuss the model we use to describe vance an ensemble of paths from τ to τ ′, which result in fluc- a single-species, Bose atomic gas with pair-wise contact tuating weights that represent the path distribution. In stan- interaction, confined in a harmonic trap in one- or three- dard AFQMC a fixed length path (corresponding to τbp + τeq, dimensions. We then describe the implementations of both with τeq being the minimum time for equilibriation or, failing our QMC method and the standard mean-field GP approach that, the maximum time that can be managed by the calcula- to study this model. Numerical results will be presented in the tion) is moved about by the Metropolis algorithm, which elim- following section, Sec. V. inates branching by the acceptance/rejection step. In other words, the estimators in Eq. (13) and Eq. (40) are the same except for the weights. A. Model Eq. (40) defines an algorithm for obtaining the estimate of Aˆ via the following steps: h ibp We use an effective potential characterized by the low en- ergy atom-atom scattering length, a . The two-body interac- (τ) s 1. A population is recorded as φ ; tion takes a simple form {| i i} 2. as the random walk continues, the path history is kept ~2 r r 4πas r r for a time interval ; U( 1 2)= δ( 1 2) . (41) τbp − m − 9

For this effective potential to be valid, several assumptions are tion, although care must be taken to validate the conditions. made; for example, the dominant effect is from s-wave scat- tering, and as is much smaller than the average inter-particle We now derive the Bose-Hubbard model from the standard spacing. For| more| details we refer the reader to Ref. 3. In the many-body Hamiltonian of the trapped boson problem in d- alkali gases these conditions are in general well met, and the dimension. In the continuous, real space, the Hamiltonian is model potential can be expected to give quatitative informa- given by:

~2 3 2 1 2 2 Hˆ = Kˆ + Vˆ = d r ψˆ†(r) r + mω r ψˆ(r) −2m∇ 2 0 Z   (42) ~2 1 4πas 3 3 + d r d r ψˆ†(r )ψˆ†(r )δ(r r )ψˆ(r )ψˆ(r ) . 2 · m 1 2 1 2 1 − 2 2 1 Z Z

d The first term is the one-body Hamiltonian Kˆ , which consists L, in a simulation cell of volume (2rb) . The lattice spacing is of the kinetic energy and the (external) confinement poten- therefore ς = 2rb/L. Further we will consider only a spheri- tial. Vˆ is the interaction Hamiltonian, which is the sum of all cally symmetric trap here for simplicity. We truncate the sim- the two-body potentials. The characteristic trap frequency is ulation cell accordingly and assume that the wave function is ω0, which is related to the so-called oscillator length scale by negligible outside the maximum sphere enclosed by the cell. ~ (Generalization to inhomogeneous traps is straightforward.) aho = /mω0. We introduce a real-space lattice, with a linear dimension of The discretized Hamiltonian corresponding to Eq. (42) is p

1 2 1 Hˆ = t c†c 2dc†c + κ ˜r ˜r c†c + U c†c c†c c†c , (43) − i j − i i 2 | i − 0| i i 2 i i i i − i i i j nn(i) i X  h ∈X i  X     where ci† and ci are the usual creation and annihilation opera- terparticle spacing, but larger than the scattering length: tors at site i. The Hubbard parameters t, U, and κ are related 1/d to the real, physical parameters as follows: a ζ ρ− . (45) | s| ≪ ≪

1 With negative as, the particles tend to “lump” together due to t = (44a) 2ς2 the gain in the interaction energy. This is a situation where we especially have to be aware of the validity of the effective 4πas U = (44b) potential. As mentioned we will do a consistency check at ςd 2 the end of the calculation to ensure that the occupancy of the ς lattice points are less than unity. κ = 4 , (44c) aho where for simplicity we have set ~ = m = 1. The lattice co- B. Implementation of QMC r r r ordinate ˜i is related to the real coordinate by ˜i = (L/2rb) i, r and ˜0 is the lattice coordinate of the trap’s center. Note that Implementation of our QMC method for this model is as is the true scattering length only in three-dimensional sys- straightforward. The number of basis M is equal to the num- tems. Nonetheless we will retain the symbol as in Eq. (44b) ber of lattice sites inside the truncated sphere of radius rb. The as a convenient measure of the interaction strength in any di- two-body term in Eq. (43) is in the desired form of Eq. (6). mension. With a negative U, the HS transformation in Eq. (7) leads to In the discretized model our resolution is limited by the lat- M auxiliary fields, with one-body propagators in the form of tice spacing. This is consistent with the conditions of validity exp( ∆τ U xinˆi), where nˆi ci†ci is the density opera- of the model interaction in Eq. (41), as it in a sense “inte- tor. Our trial| wave| function Ψ ≡is the Gross-Pitaevskii (GP) p | Ti grates out” the short-range dynamics. In this model our lattice wave function ΦGP, which we describe in the next subsection. constant ζ must be much smaller compared to the average in- We mention here a technical point in the implementation. 10

The ground-state projection in our method involves the appli- the non-interacting Hamiltonian (with U =0). The density is Aˆ ˆ (0) cation of one-body propagator in the form of e on a single- fed back to construct the initial Hamiltonian HGP in (48). Di- permanent wave function φ . This usually translates into a rect diagonalization of this one-body Hamiltonian yields its | i matrix-vector multiplication in the computer program, which ground state Φ(1) . We thus obtain an updated density n¯(1) 2 GP i generally costs (M ). Often there are special properties of | i (1) and a better Hamiltonian Hˆ . This procedure is iterated un- ˆ O GP A that can be exploited to evaluate the one-body propagator til the desired convergence criterion is satisfied. We choose more efficiently. In the Bose-Hubbard Hamiltonian, the only our convergence condition to be: non-diagonal part of the Hamiltonian in real space is the ki- netic operator in Kˆ . We can separate it from the other one- dr ϕ(t+1)(r) ϕ(t)(r) | − | < ǫ , (50) body operatorsand apply the kinetic propagatorin momentum 1 dr ϕ(t+1)(r)+ ϕ(t)(r) space. Wave functions are quickly translated between these 2R | | two representations using the Fast Fourier transform (FFT). R 13 where ǫ is a small number (usually on the order of 10− for 1 ∆τKˆ In this way, the actual application of e− 2 involves only double precision numbers). 1 ∆τKˆ diagonal matrices; thus the overall cost for each e− 2 op- The second method we use to solve Eq. (48) avoids the eration is reduced to (M log M). We observe in our calcu- diagonalization procedure. It is closely related to the QMC lations that the additionalO Trotter error is much smaller than method, both computationally and formally (see Sec. VI). We the error already introduced in the original breakup, Eq. (4). ∆τHˆ use the ground-state projector e− GP :

∆τHˆ n (0) n (e− GP ) Ψ →∞ Φ . (51) C. Implementation of Gross-Pitaevskii self-consistent equation | i −→ | GPi The initial wave function is arbitrary and can be, for example, The Gross-Pitaevskii (GP) wave function ΦGP is the single- chosen again as the solution with U =0. The feedback mech- permanent wave function anism through the density profile n¯i remains the same. By using the same Fast Fourier transform for the kinetic propaga- r r r r r r ΦGP( 1, 2,... N )= ϕ( 1)ϕ( 2) ϕ( N ) , (46) tor as described in subsection IVB, a speed gain is obtained, ··· especially for large systems. In practice we have often found which minimizes the expectation value of the ground-state en- this method to be a simpler and faster alternative to the first ergy. Such a wave function satisfies the self-consistent Gross- method of diagonalization and iteration. Note that the scalar 1 N 1 2 Pitaevskii equation [27, 28, 29] term − U n¯ does not affect the projection pro- − 2 N i i cess, but with it Hˆ corresponds to the original many-body ~2 PGP 2 r 1 2 r r 2 r ˆ ˆ ϕ( )+ 2 mω0 0 ϕ( ) Hamiltonian in that ΦGP HGP ΦGP = ΦGP H ΦGP . − 2m∇ | − | (47) h | | i h | | i N 1 4πa ~2 + − s ϕ(r) 2ϕ(r)= µϕ(r) . N m | | V. RESULTS [We keep the prefactor (N 1)/N, since we will study both large and small values of N−.] In this section we present results from our QMC and GP To compare our QMC results to those of mean-field, we calculations in one-, two-, and three-dimensions. To validate carry out GP calculations on the same lattice systems. The our new QMC method and illustrate its behavior, the major- discretized GP Hamiltonian in the second-quantized form is: ity of the calculations will be on systems where exact results are available for benchmark. These include small lattices, which can be diagonalized exactly, and the case of attractive Hˆ = t c†c 2dc†c δ-function interactions in one dimension, where analytic solu- GP − i j − i i i j nn(i) tions exist. For the purpose of presenting the method to facili- X  ∈X  1 2 tate implementation, some numerical results and comparisons + κ ˜r ˜r c†c 2 | i − 0| i i (48) are shown in detail to illustrate the behavior and characteris- i X tics of the method. N 1 1 2 + − U n¯ c†c n¯ . Most of the results we present here will be for attractive N i i i − 2 i i interactions, where the method is exact and is free of any X   phase problem [22] from complex propagators (see subsec- Here n¯i is the expectation value of the density operator: tion VC). Such systems therefore provide a clean testground for our new method. In addition, with attractive interactions ΦGP ci†ci ΦGP the condensate in 3-D is believed to collapse beyond a critical n¯i h | | i . (49) ≡ ΦGP ΦGP interaction strength or number of particles. Mean-field cal- h | i culations [30] estimate the collapse critical point to be about We have implemented two methods for solving the GP Nas/aho = 0.575. The exact behavior of the condensate equation. The first is the usual self-consistent iterative ap- near the critical− point is, however, not completely clear, as (0) proach. We generate an initial density profile, n¯i , by solving many-body effects are expected to have an impact. At the end 11 of this section we will also show some preliminary results for to rather large ∆τ values. We see that both quantities converge larger systems with both attractive and repulsive interactions to the exact results as ∆τ 0. in 3-D. → We measure the ground-state expectation values of the fol- Trap Potential Energy lowing quantities: the ground-state energy, kinetic energy 0.850 QMC (τ = 4.0) Tˆ , external confining potential Vˆ , interaction energy BP h i h trapi Exact diag. Vˆ2B , density profile nˆ , and the condensate fraction (of- i 0.840 tenh abbreviatedi “cond.frac.”h i in the tables and figures). The condensate fraction is defined as the largest eigenvalue of the Total Energy diagonalized density matrix [3]. If we write the one-body 0.830 −0.98 Green’s function matrix c†c in terms of its eigenvalues h i ji −0.99 nα and eigenvectors χ (i) : 0.820 { } { α } −1.00 −1.01 c†c = nαχ† (i)χα(j) , 0.810 h i j i α −1.02 α X 0 0.05 0.1 0.15 0.2 0.800 then the largest eigenvalue divided by the total number of par- 0 0.05 0.1 0.15 0.2 ticles gives the condensate fraction. ∆τ

FIG. 1: Convergence of QMC observables with ∆τ. The system has A. Comparison with exact diagonalization: as < 0 the same parameters as in Table I. Exact results are shown as dotted lines. Lines connecting QMC data are to aid the eye. The many-body Hamiltonian (43) can be diagonalized ex- actly for small systems to benchmark our QMC calculation. To illustrate the convergence of observables in back- We compare our QMC results with exact diagonalization for propagation length, we show in Fig. 2 the various observables a one-dimensional lattice of 13 sites, and study its behavior computed by QMC as a function of τbp. Separate calculations for different values of the interaction strength as and number were done for different values of τbp. For all calculations, of particles N. a small ∆τ value of 0.01 was used. We see that all quanti- The first system we study has 5 bosons, with t = 2.676, ties converge to the exact results rather quickly, by τbp 2. , . These values correspond to the ∼ U = 1.538 κ = 0.3503 (The total energy H is of course exact for any τbp, including physical− parameters A˚ and h i aho = 8546 as = 5.292 τbp = 0.) As we see from the energy expectations, this is in 6 ˚ 1 − × 10− A− . (Recall that, by our definition, as in 1-D does fact a system with significant interaction effects. Alkali sys- not have the dimension of length, and is not the scattering tems at the experimental parameters often have significantly length itself.) Table I shows the comparison of the quantities weaker interaction strengths and the convergence rate is ex- computed using three methods: QMC, GP, and exact diag- pected to be even faster. onalization (ED). The statistical uncertainty of QMC results Our QMC method is exact and therefore independent of the are presented in parantheses. We see that the agreement be- trial wave function ΨT, except for convergence rate and sta- tween QMC and ED is excellent. GP makes significant errors tistical errors. In Fig. 3 we show QMC results obtained us- here because of the sizable interaction strength as well as the ing two different ΨT’s, the noninteracting solution and the small number of particles. GP wave function. The convergence of condensate fraction and trap energy are shown versus back-propagation time τbp for a system of 6 particles on 13 sites. The calculations lead TABLE I: Comparison of QMC calculation against exact diagonal- to the same results. The quality of , however, does af- ization (ED) and Gross-Pitaveskii (GP). The system has 13 sites, 5 ΨT particles, t = 2.676, U = −1.538, κ = 0.3503. In the QMC cal- fect the variances of the observables and their convergence rates with τ . For example, the noninteracting wave func- culation we use ∆τ = 0.01, τbp = 4.0, and the GP solution as the bp trial wave function. tion, which disregards the two-body interaction, is more ex- tended (in its density profile) than GP. Its mixed estimator is Type g.s.energy hTˆi hVˆtrapi hVˆ2Bi cond.frac. therefore worse than that with the GP trial wave function. The ED −1.009 4.278 0.8427 −6.129 95.59% mixed-estimator for the ground-state energy is exact in both, QMC −1.008(2) 4.279(3) 0.8423(5) −6.129(2) 95.59% but the variance is slightly larger with the former. GP −0.493 3.919 0.7504 −5.162 100% We now show results for different systems with N from2 to 9 bosons, and varying interaction strengths. We note that if we keep the product U (N 1) constant, the Gross-Pitaevskii To illustrate the convergence in imaginary-timestep ∆τ, we equation predicts the× same−per-particle energies and densities. show in Fig. 1 the total energy and the average trap energy For brevity, we shall refer to the curve in which U (N 1) × − Vˆtrap . The former can be obtained exactly from the mixed is constant as the GP isoline. Deviation from the GP isoline estimatorh i while the latter requires back propagation. To show is therefore an indication of the effect of many-body corre- the Trotter error, we have deliberately done the calculationsup lations. In order to show results on multiple systems at the 12

Total Energy Condensate Fraction (%) Total Energy Interaction Energy -0.40 0.0 -0.9 -0.50 100 -1.0 -0.1 -0.60 99 QMC -1.1 Exact result -0.70 98 GP -0.2 -1.2 -0.80 -1.3 97 -0.90 -0.3 -1.4 96 -1.00 QMC -1.5 QMC -0.4 Exact diag. Exact diag. -1.10 95 -1.6 GP GP 0 1 2 0 1 2 -0.5 -1.7 2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 Kinetic Energy Trap Energy Interaction Energy N N 4.4 0.86 -5.0 0.84 -5.2 FIG. 4: Comparison of QMC, GP, and ED results for different sys- 4.2 0.82 -5.4 tems. Calculations were done along a GP isoline U × (N − 1) = 0.8 -5.6 −2.30t for up to nine particles in 13 sites. The graphs show the to- 4.0 0.78 -5.8 tal and interaction energies per particle. QMC and exact results are 0.76 -6.0 indistinguishable. GP is accurate in the limit of weak correlation but -6.2 3.8 0.74 deviates more from the exact results as the system becomes more -6.4 0 1 2 0 1 2 0 1 2 correlated. The solid lines are to aid the eye. τ BP

FIG. 2: Convergence of the computed observables versus τbp. The system is the same as in Table I. The different panels show five 0.1605(2) for N = 2 and 9 particles, respectively, while the different observables. The horizontal axes are the back-propagation GP value is 0.1501. In GP, interaction energy is lowered by length. Exact results are shown as dotted lines, while GP results as increasing particle overlap, namely by shrinking the profile. dash-dotted lines. Solid lines are present only to aid the eye. In reality, the particles find a way to reduce interaction with- out statically confining to the central sites, resulting in a more Condensate Fraction (%) Trap Energy extended one-body profile. 1.3 QMC (Ψ = GP) 102 T 1.2 QMC (Ψ = nonint) 0.55 T 1.1 GP QMC, N = 2 100 1.0 0.50 QMC, N = 9 0.9 0.45 GP 98 0.8 0.40 0.7 96 0.6 0.35 0.5 0.30 94 0.4 0.25 0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4 τ τ 0.20 BP BP 0.15 FIG. 3: Independence of QMC results on trial wave functions (“GP” 0.10 for Gross-Pitaevskii, “nonint” for noninteracting solution). The sys- 0.05 tem is the same as in Table I, except that here we use 6 particles. The 0.00 horizontal axes are the back-propagation length. Lines connecting 2 3 4 5 6 7 8 9 10 QMC data points are present only to aid the eye. Lattice site FIG. 5: The normalized density profiles as an illustration of parti- cle correlation effects. Results are for 13-site systems along the GP same time we will scan GP isolines. Figure 4 shows the QMC isoline U × (N − 1) = −2.30t. The normalized GP curve is iden- and GP results as a function of the number of particles. In the tical for any number of particles along this line. QMC results are GP calculations the per-particle quantities are constants. The shown for N = 2 and N = 9. The QMC results have very small QMC results, on the other hand, capture the effect of corre- errorbars and are indistinguishable from ED (not shown). The QMC lation. Both the total energy and the interaction energy are density profiles are more extended, although the interaction energies lowered from the GP results. The exact results deviate from are lower than GP, as shown in Fig. 4. GP more as the system becomes more correlated along the GP isoline, i.e. when U is increased or when N is decreased. Al- though N is too small here because of the limitation of ED, the results are representative of the general trend in larger sys- B. Comparison with analytic results in 1-D: as < 0 tems (see below). Figure 5 further illustrates the effect of particle correla- The problem of an arbitrary number of untrapped bosons tion in this system. Although the exact interaction energy is interacting with an attractive δ-potential in one dimension can lower than that of GP, the exact density profile is more ex- be solved analytically [31], yielding analytic expressions for tended. This is also manifested in the average trap potential the total energy and density profile. In this section we carry energy Vˆ /N, where the QMC results are 0.1981(8) and out QMC and GP calculations and compare our results against h trapi 13 these analytic results, on systems of up to 400 bosons. The −200 −100 0 100 200 Analytic Hamiltonian in the continuous real space is 6.0 1.0e−02 GP QMC

N 2 N 5.0 1 ∂ 1 ) ˆ −3 1.0e−03 H = 2 g δ(xi xj ) . (52) −2 ∂x − 2 − 10 i=1 i i>j=1 × X X 4.0

The interaction constant ( ) is related to our Hubbard pa- 1.0e−04 g > 0 3.0 rameters by g U/√t . The ground state of this Hamilto- nian is an -boson≡ | bound| state. By fixing the center of mass N 2.0 at x = 0, we can eliminate the contribution from its overall Normalized density ( motion, which leads to the following analytic expressions for 1.0 the density profile [32],

0.0 N 1 2 gnN x /2 −300 −200 −100 0 100 200 300 − n(N!) e− | | ρ(x)= 1 g ( 1)n+1 , (53) Distance (lattice unit) 2 − (N + n 1)!(N n 1)! n=1 − − − X FIG. 6: Comparison of calculated density profiles from QMC and and the total energy, GP with analytical results. The densities are normalized. The QMC errorbars are displayed every five data points to avoid cluttering the E = 1 g2N(N 2 1) . (54) plot. The QMC profile is given by the dotted curve. The inset shows − 96 − the same curves with logarithmic vertical scale, indicating that at large distances the density is exponential. In our QMC calculations, we again put the system on a real- space lattice. The lattice size is chosen to be large enough so that discretization errors are comparable to or smaller than sta- tistical errors. As the ground state of the system is a droplet as N is decreased, mean-field results deviate more and more in the absence of the external confining potential, the center from the exact results. For example, as we go from g = 0.01 of mass can slide in the calculation due to random noise. We (N = 400) to 10 times the strength along the isoline, the sys- therefore need to subtract the center-of-mass motion. Tech- tematic error in the GP total energy increases roughly from nically, this can be accomplished conveniently in the random 0.5% to 5%. walk by treating the system with respect to its center of mass. In Appendix B, we describe our method for this correction, -0.164 which is applicable in any situation where the center of mass and relative motions need to be separated. In our calculations, -0.166 the correction affects the kinetic and total energies as well as the density profiles. The results shown below were all ob- tained with such a correction applied. -0.168 We first study a system of 20 particles with g = 0.154. Table II shows the energies, and Fig. 6 the density profiles. -0.170 This is a system where mean-field makes significant errors. Our QMC results are in excellent agreement with the exact Energy per particle -0.172 results. Analytic QMC -0.174 TABLE II: Comparison of QMC and GP results to available exact GP results. The system has 20 particles and g = 0.154. A lattice of 1024 0 50 100 150 200 250 300 350 400 sites was used, with ∆τ = 0.01 and τbp = 2.5. N Type g.s.energy Tˆ Vˆ cond.frac. h i h 2Bi FIG. 7: Comparison of the energy from QMC (crosses) with the Analytic result −1.971 - - - exact answer (dotted curve) for different number of particles. Energy QMC −1.964(8) 2.044(8) −4.007(4) 99.76% per particle is shown along the GP isoline g × (N − 1) = 4.0. The GP result is the flat, dash-dotted line. We use a lattice of 1024 sites, GP −1.784 1.776 −3.561 100% ∆τ = 0.01 and τbp = 4.0.

We next scan systems with various numbers of particles by We now study the system along a different line, holding the following the GP isoline g (N 1) = 4.0. The energy interaction strength g fixed while scanning the number of par- per particle is shown as a function× − of N in Fig. 7, for up to ticles, again up to N = 400 particles. Figure 9 shows the 400 particles. Fig. 8 shows the density profiles for up to 100 behavior of Hˆ /N 3 for up to 400 particles, with g =0.0403. particles. Again, the agreement between QMC and exact re- At large N,h thei total energy is roughly proportional to N 3. sults is excellent. As the interaction strength g is increased or Compared to Figs. 7 and 8, the interaction strength here is 14

10 -1.60 GP Analytic QMC, N = 5 QMC 9 QMC, N = 20 GP QMC, N = 100 8 ) -1.64 −5 7 ) × 10 −3 ( 3 10 N × 6 / 〉

H -1.68 5 〈

4 Normalized density ( 3 -1.72 0 50 100 150 200 250 300 350 400 2 N

1 FIG. 9: Comparison of computed ground-state energy for different numbers of particles N. The interaction strength is held constant at 3 0 g = −0.0403. The total energy divided by N is shown as a function 0 50 100 150 200 of N for QMC, GP and exact calculations. Conservative parameters Distance from the center−of−mass (lattice unit) were used, with τbp = 4.0 in all case, and ∆τ = 0.01 for N < 200 and ∆τ = 0.005 otherwise. FIG. 8: Comparison of the density profiles from QMC and GP with analytic results. The normalized densities are shown along the GP isoline g × (N − 1) = 4.0 for several N values. The system is the with 13 sites and 4 particles. The agreement between QMC same as that in Fig. 7. The GP density is the same for any N on the isoline, and is given by the dash-dotted line. and exact result is excellent. Results from GP are also shown. The GP and QMC density profiles have roughly the same size, as evident from the values of Vˆ . However, GP overesti- h trapi stronger at larger N and weaker at lower N, with the crossover mates the interaction energy because it does not take into ac- at N 100. Most of the calculations are therefore more count the particle-particle correlation. In the mean field pic- challenging∼ numerically. Again QMC was able to completely ture, expanding the density profile is the only way to lower recover the correlation energy missed by GP. At large N, the interaction energy, so that the particles overlap less with each other. (Note that Vˆ is indeed slightly larger for smaller timesteps were used and more computing was nec- h trapi essary to reduce the statistical errors. (Note that the errorbars GP.) In reality, particles can avoid each other more effectively appear larger at smaller N in the plot because of the division by means of many-body correlation. The QMC correctly re- by N 3.) covers this correlation, which lowers the total energy without spreading the density as much as GP does.

C. Comparison with exact diagonalization: as > 0 TABLE III: Comparison of QMC results against exact diagonaliza- tion (ED) and Gross-Pitaveskii (GP) in 1-D. Here we use 13 sites and We have shown that our new QMC algorithm is exact and 4 particles; t = 2.676, U = +1.538, κ = 0.3503; ∆τ = 0.01 and works well for a wide range of systems with attractive inter- τbp = 2.5. actions. If the interaction is repulsive (a > 0, or equivalently s Type g.s.energy hTˆi hVˆtrapi hVˆ2Bi cond.frac. U > 0) the one-body propagatorsresulting from the HS trans- ED 4.24 1.18 1.793 1.269 98.5% formation become complex,in theformof exp(i√∆τUxinˆi). The same algorithm applies in this case as well. In principle QMC 4.24(2) 1.18(2) 1.790(8) 1.273(8) 98.6% the complex one-body operator only requires a change to the GP 4.43 1.03 1.800 1.599 100% corresponding complex operations. But in practice a serious phase problem occurs, which causes the calculation to lose ef- ficiency rapidly at larger interaction strengths. We discuss this Table IV shows results for bosons in a two-dimensional problem and how to control it below. Our initial studies indi- trap, using a 4 4 lattice. The GP solution also exhibits the cate that, for moderate interaction strengths, the algorithm as same behavior× as in the 1-D calculation, in that the density is remains very efficient and gives accurate results, allowing profile is slightly more extended, and the interaction energy is reliable calculations for parameters corresponding to experi- overestimated. As in other cases, the QMC statistical errorbar mental situations in 3-D. on the condensate fraction was not computed directly, but we We benchmark our algorithm in one- and two-dimensional estimate it to be on the last digit. systems with repulsive interactions against exact diagonaliza- As mentioned earlier, the only modification necessary to tion. Table III shows results for a one-dimensional system, the algorithm in order to treat repulsive interactions (as > 0) 15

exact ground-state wave function. We see that this is indeed TABLE IV: Comparison of QMC calculations against exact diag- the case in Table V. The interaction energy is lowered in the onalization (ED) and Gross-Pitaveskii (GP) projection in a 4 × 4 lattice, with 4 bosons. t = 0.2534, U = +0.3184, κ = 3.700; many-body calculation as expected. Interestingly, the external potential energy is lower than in GP. Consistent with this, the ∆τ = 0.01 and τbp = 2.5. exact density profile is tighter than in GP, as shown in Fig. 10. ˆ ˆ ˆ Type g.s.energy hT i hVtrapi hV2Bi cond.frac. The trend here appears different from what we observed in ED 6.000 1.818 3.8326 0.350 97.8% small 1-D trapped systems in Fig. 5, but consistent with the QMC 6.005(6) 1.817(2) 3.8325(2) 0.355(5) 97.8% large untrapped systems in Fig. 8. We are presently carrying GP 6.067 1.763 3.8359 0.469 100% out more calculations to cover a wider range of parameters and study the role of dimensionality.

TABLE V: Comparisons of QMC and GP calculations for 175 parti- is to allow complex arithmetic. A more serious problem can cles in a 3-D spherical trap, with as = −22.4 A˚ and aho = 8546 A.˚ occur, however. The orbitals and the walker weights become The energies are displayed as per-particle quantities. Both the QMC complex numbers. Asymptotically the phase of these weights and GP results are extrapolated to ∆τ → 0. will be uniformly distributed in the complex plane. The de- ˆ ˆ ˆ nomitors in Eqs. (34) and (40) will be dominated by noise, Type g.s.energy hT i hVtrapi hV2Bi cond.frac. causing the Monte Carlo sampling efficiency to decay and ul- QMC 16.979(6) 16.47(5) 6.54(1) −6.03(4) 99.73% timately destroying the algebraic scaling of QMC. This is the GP 17.115 15.60 6.77 −5.25 100% so-called sign or phase problem [19, 22]. In real-space meth- ods this problem is connected to fermions, but here we have a situation where a phase problem appears in the ground state of a bosonic system. Physically, it is easy to see why a phase 4.0 GP problem must occur. Our many-body wave function is being QMC 3.5 represented in IOR, with only one orbital in each walker. With ) −3 a repulsive interaction, the only way to reflect correlation ef- Å 3.0 −15

fects, i.e., particles avoiding each other, is to make the orbitals 10 2.5 complex. × As we see below, our algorithm remains efficient and gives 2.0 accurate results for large systems with scattering lengths cor- 1.5 responding to experimental situations in 3-D. As the interac- 1.0

tion strengths become much stronger, the phase problem will Normalized Density ( ultimately make the approach ineffective. We have done pre- 0.5 liminary calculations in which we control the phase problem 0.0 −20000 −10000 0 10000 20000 by applying a phaseless formalism described in Ref. 22. Our Distance (Å) results indicate that the systematic errors introduced by the phaseless approximation are small for moderate interaction FIG. 10: Comparison of density profiles from the QMC and GP for strengths. We expect to therefore be able to obtain accurate 175 particles. The system is the same as described in Table V. The and reliable results for scattering lengths well into the exper- QMC profile is more peaked and tighter than GP. imental ’strong-interaction’ regime achievable by Feshbach resonnance. We now turn to bosons with repulsive interactions in three- dimensional trap. We again use a 15 15 15 lattice, and sim- × × ˚ ulate 100 bosons. We choose a scattering length as of 80 A. D. Realistic calculations in three-dimensions This value is close to the experimental 39K singlet [33] or 87Rb triplet [34] scattering lengths. In Table VI we show the cal- In this section we present some test results on realistic culated energies and condensate fraction. For this interaction systems of trapped particles in three-dimensions. QMC re- strength, the impact of the phase problem on the statistical er- sults were obtained with back-propagation and conservative ror is small, and the QMC calculation is very efficient. The choices of ∆τ and convergence parameters. We expect the true condensate is, like in the 1-D repulsive case, tighter than QMC results to be exact. We also carry out the corresponding that predicted by GP, with lower interaction energy. Gross-Pitaevskii calculations, and make comparisons against our exact QMC results. Table V shows the result of a QMC calculation for 175 par- VI. DISCUSSIONS ticles in a three-dimensional trap. We choose a trap with a ˚ A. Connection between QMC and Gross-Pitaevskii projections characteristic length aho = 8546 A. The trap was discretized into a 15 15 15 lattice, in a range that corresponds to × × ˚ about 5 aho. The scattering length is as = 22.4 A. In The QMC method we have presented allows us to go be- this regime× the GP solution is a good approximation− to the yond mean-field and treat many-body effects. On the other 16

example, for Bose-Hubbard model the last two terms in the TABLE VI: QMC calculation of 100 particles in a three-dimensional exponent lead to the GP mean-field potential trap. A lattice of 15 × 15 × 15 was used. The parameters correspond to a = 8546 A˚ and as = 80 A.˚ The quantities displayed are for ho U n¯ nˆ 1 n¯2 . per particle. i i − 2 i (57) i Type g.s.energy hTˆi hVˆtrapi hVˆ2Bi cond.frac. X   Apart from the factor (N 1)/N which approaches unity in QMC 24.687(9) 9.573(9) 11.933(5) 3.181(3) 99.80% the limit of large N, we have− recovered the GP propagator. . . . . GP 24 922 9 281 12 028 3 612 100% The projection with Eq. (56) lowers the variational energy for any initial φ and is stationary when φ is the GP solution. This is why| GPi is the best variational| wavei function that has the form of a single permanent, and hence a reasonable trial hand, it has a deep connection with the GP mean-field ap- wave function to use for most of our QMC calculations. proach. Our approach uses an HS transformation which leads It is also clear from the discussion above that the impor- to integrals of single-particle operators over auxiliary-fields. tance sampling formalism allows us to have an optimal form The mean-field solution can be regarded as the leading term of HS transformation, in that the HS propagator y(ˆv v¯) in- in the stationary-phase asymptotic expansion of the exact so- e − volves only the difference . In other words, although lution [35]. Our method evaluates this exact solution, which vˆ v¯ in Eq. (7) we write the decomposition− for the bare inter- is in the form of many-dimensionalintegrals, by Monte Carlo. action term, the importance sampling transformation effec- In this section we further comment on the formal connection tively introduces a mean-field background based on the trial between our importance-sampled QMC and the GP as done wave function and allows the HS to deal with only a residual by projection (the second of the two GP methods discussed in quadratic interaction term, 2. subsection IV C). (ˆv v¯) To summarize, our QMC method− reduces to GP if we eval- Let us reconsider the two-body propagator in the modified uate the many-body propagator by the stationary-point ap- AF transformation Eq. (25). Let us suppose that we are now proximation, using only the centroid of the Gaussian. The taking our first Monte Carlo step, where our walker is φ , and full method evaluates the many-dimensional integral over we will also use the same wave function as Ψ . Following| i T auxiliary-fields exactly by Monte Carlo. It captures the in- the discussion of the optimal choice of ~x in the| samei section, teraction and correlation effects with a stochastic, coherent IIIB, we know that ~x = 0 is a stationary point with the choice ensemble of mean-field solutions. The structure of the cal- culation can be viewed as a superposition of the GP projec- tions that we have described. Our method therefore provides a φ vˆi φ x = √∆τ v¯i √∆τ h | | i . (55) way to systematically improve upon GP while using the same i − ≡− φ φ h | i framework. We can approximate the integral in Eq. (25) by the value of the integrand at ~x = 0, which can be justified in the limit of B. Computing small ∆τ. More explicitly, as ∆τ 0, the Gaussian function becomes the most rapidly varying→ term in the integrand. To exhibit the asymptotic behavior of this integral, we change the Because of the structure of QMC as a superposition of GP integration variable to ~y √∆τ ~x, so that the large parameter projections, our method scales gracefully with system size. 1/∆τ appears in the Gaussian’s≡ exponent: As discussed in Sec. IVB, the bulk of our method scales as (M log M), with the significant speedup from using Fast O y2/2∆τ Fourier transform. For example, the QMC calculation shown 1 ∆τvˆ2 ∆τ( 1 v¯2 v¯vˆ) ∞ e− y(ˆv v¯) e 2 = e− 2 − dy e − . in Table VI required less than 8 hours on a single Alpha EV67 √2π∆τ Z−∞ processor. The 1024-sites QMC calculation shown in Table II took about four hours to get good statistics, with very con- The dominant contribution to the integral comes from the servative choices of ∆τ and other convergence parameters. maximum of the Gaussian function at y =0. The asymptotic It required about 1.3 gigabytes of memory, largely because leading term of the importance-sampled many-body propaga- of back-propagation path recording. In contrast, treated fully, tor is therefore: the latter problem would mean the diagonalization of a sparse, 41 2 1 2 Hermitian matrix containing (8 10 ) elements. Although ∆τ Kˆ v¯ivˆi+ v¯i e− − i 2 i , (56) this can be reduced by exploiting× symmetries, exact diagonal- P P  ization of this problem is clearly not within reach with com- where Kˆ is the one-body term in the original Hamiltonian. puting capabilities in the foreseeable future. Under this approximation, our random walk becomes deter- We typically use hundreds of walkers in our calculation. ministic, needing only one walker. If for the next step we use The stochastic nature of QMC means the number of walkers the updated wave function φ′ to evaluate the new v¯i in fluctuates due to branching and killing of walkers with very Eq. (55), we obtain a self-consistent| i projection with one-b{ }ody large and very small weights (see subsection III). The popu- propagators. In fact, the one-body Hamiltonian in the expo- lation therefore must be controlled to ensure that it does not nent of Eq. (56) is precisely the mean-field Hamiltonian. For grow or decay too much, and that the walker weights have a 17 reasonable distribution. Our method to control the population different dimensions and under different conditions. In partic- is similar to that discussed in Ref. 25. ular, it would seem straightforward to generalize our present We comment on the effect of the number of particles, N, framework to study rotations and vortices, since we are al- on scaling. Because of the use of IOR, the number of parti- ready dealing with complex propagators and wave functions cles does not enter in the propagation. It would then seem as in the repulsive case. In addition, it will be interesting to though the algorithm might have a super-scaling in N. This treat boson-fermion mixtures with our approach. As men- ∆τHˆ is not true, of course, since the projector e− depends on tioned, the auxiliary-field method is already widely used to N. For example, the shift v¯i has a factor of N in front (see treat strongly interacting fermion systems. Appendix A), and the local energy scales with N. As a result, a smaller time-step must be used for larger N. The above ar- guement suggests a linear reduction in ∆τ as N is increased, Acknowledgments which we have used as a rough guideline in our calculations to select the range of ∆τ to use. Extrapolations with separate We thank D. M. Ceperley and H. Krakauer for stimulat- calculations using different ∆τ values are then carried out. ing discussions. Financial support from NSF (grant DMR- 9734041), ONR (grant N00014-97-1-0049),and the Research Corporation is gratefully acknowledged. SZ expresses his C. Conclusion and Outlook gratitude to Prof.’s Ceperley and Richard Martin for their hos- pitality during a sabbatical visit, where part of the work was In conclusion, we have presented a new auxiliary-field carried out. We also thank the Center of Piezoelectric by De- QMC algorithm for obtaining the many-body ground state sign (CPD), where part of our computing was performed. of bosonic systems. The method, which is based upon the field-theoretical framework and is essentially exact, provides a means to treat interactions more accurately in many-body APPENDIX A: IDENTICAL-ORBITAL REPRESENTATION systems. Our method shares the same framework with the GP approach, but captures interaction and correlation effects with In this appendix we show that the matrix representation of a stochastic ensemble of mean-field solutions. We have il- an N-boson wave function in AFQMC can be made particu- lustrated our method in trapped and untrapped boson atomic larly simple. In fermion calculations, we must use an M N gases in 1-, 2-, and 3-dimensions, using a real-space grid as matrix to represent a determinant, because the orbitals mus×tbe single-particle basis which leads to a Bose-Hubbard model mutually orthogonal. In the boson case, however, this restric- for these systems. We have demonstrated its ability to obtain tion is absent. The most general form of a many boson perma- exact ground-state properties. We have also carried out the nent is expensive to compute, having complexity of (NM!). GP mean-field calculations and compared the predictions with But we can choose to make all the orbitals identical.O In ma- our exact QMC results. Our method is capable of handling trix language, we will have only an M-row column vector. large systems, thus providing the possibility to simulate sys- We will term this representation identical-orbital representa- tem sizes relevant to experimental situations. We expect the tion—IOR. Each many-boson wave function in IOR has the method to complement GP and other approaches, and become form of a GP mean-field solution. Two conditions are neces- a useful numerical and theoretical tool for studying trapped sary for this choice to be viable in the QMC: that an initial atomic bosons, especially with the growing ability to tune the trial wave function of this form is allowed and that successive interaction strengths experimentally and reach more strongly projections preserve the form. The only requirement for the interacting regimes. former to hold is that the wave function in IOR not be orthog- From the methodological point of view, more work remains onal to the true many-body ground state, and it is straightfor- to be done with the repulsive case to deal with the phase prob- ward to show that Eq. (12) holds for a φ in this form. More lem. We have shown that our method as it stands can be complex wave functions can always be| generatedi by a linear very useful for moderate interaction strengths. For stronger combination of such wave functions. In fact, this is what we interactions, our preliminary study indicates that the phase- accomplish through our Monte Carlo simulation. less approximation [22], which eliminates the phase problem In operator language, a single N-boson wave function φ but introduces a systematic error, is very accurate for scatter- is given by | i ing lengths well into the Feshbach resonnance regime. We are N currently examining this more systematically to quantify the φ = φˆ†φˆ† φˆ† 0 = φˆ† 0 , extent of the bias. Because of the simplicity of these bosonic | i ··· | i | i systems compared to electronic systems, they provide an ideal N  testbed, where for small sizes the problem is readily solved by ˆ | {z } where φ† α cα† φα. In matrix form, φ would be M N exact diagonalization. matrix φ ≡whose columns are identical.| Thei overlap of× two A variety of applications are possible. The ground state such wave functionsP is given by of the Bose-Einstein condensates with both attractive and re- pulsive interatomic interactions can be studied for various in- ψ φ = per ψT φ h | i · teraction strengths, including the strongly interacting regime N = N!(ψ† φ) , reached by Fesbach resonance. They can also be studied in ·  18 where the bold-phased symbols ψ and φ represent the single- random-walk process, we are now free to correct for the CM column vectors for ψ and φ, respectively. Similarly, for any motion by shifting the walkers back to the origin whenever one-body operator Aˆ, necessary. For consistency, this correction must be applied both in the normal random walk and in the back-propagation N 1 ψ Aˆ φ = N! N(ψ† A φ)(ψ† φ) − , (A1) phase. h | | i · · · where A is the matrix for Aˆ. The matrix element of a quartic (two-body) operator is given by: 2. Separating the center-of-mass kinetic energy

N 2 ψ b† b† b b φ = N! N(N 1)ψ∗ ψ∗ φ φ (ψ † φ) − . h | α β γ δ| i − α β γ δ · The moving trial wave function, however, poses a problem (A2) for the calculation of the kinetic energy. Now the orbitals are free to slide, and the diffusive motion of the orbital’s CM is no longer suppressed in the LAB frame. When we use the APPENDIX B: DROPLET CENTER-OF-MASS usual t-term in the Hamiltonian in Eq. (43) to compute the ki- CORRECTION netic energy, we obtain the total Tˆ , in which T Tˆ h i cm ≡ h cmi and the desired Tˆ′ are mixed. This leads to a spurious in- 1. Correcting the density broadening crease in the estimateh i of the kinetic energy and consequently the total energy. For example, the uncorrected ground-state To handle the droplet system given by the translationally energy for the system shown in Table II would be 1.887(2) − invariant Hamiltonian in Eq. (52), an extra ingredient is nec- with Tˆ = 2.092(3); thus the total energy is overestimated essary in addition to the “basic” QMC algorithm that we have h i by 0.08 due to the contribution from Tcm. Since we know the described. In a deterministic calculation, for example in GP, nature of the CM motion, it is fairly straightforward to extract the motion of the center-of-mass (CM) can be simply elimi- Tcm and explicitly subtract it from the kinetic and total energy nated by fixing it at the origin, as in Eq. (53). In the QMC estimates. Allowing the droplet to freely slide in the calcula- calculation, however, the orbitals fluctuate as they are propa- ∆τTˆ tion is equivalent to having a spurious “propagator” e− cm , gated by Bˆ(~x ~x), where the random fields ~x are drawn from − whose effect on the wave function for the CM is described by a Gaussian probability density. Random noise will inevitably the diffusion equation cause the CM of the system to slide, undergoing a free diffu- sion whose average position is the origin. ∂Ψ (R, τ) cm = Tˆ Ψ (R, τ) . Left unchecked, this spurious CM motion will lead to an − ∂τ cm cm artificial broadening of the density profile. To correct for it in the density profile, we could simply shift the CM of every It is a well known property of such a diffusion process that the walker back to the origin. However, the importance-sampled averaged squared distance R2(τ) grows linearly with the propagator involves ratios of overlaps with the trial wave func- (imaginary) time τ: h i tion ΨT φi , which would have to be corrected in the random walkh whenever| i a shift is made. R2(τ) = bτ . h i Instead our solution to this diffusive motion is to let the trial wave function slide along with the walkers. In other words, We can obtain b by recording the quantity R2(τ) for a pe- h i we rewrite the kinetic energy operator as riod of time in the QMC simulation. The constant b is lin- early proportional to Tcm. More specifically, the center-of- Tˆ = Tˆcm + Tˆ′ , (B1) mass Hubbard hopping parameter tcm can be extracted from b: where Tˆcm represents the CM kinetic energy, and Tˆ′ the inter- nal kinetic energy in the CM frame. The total Hamiltonian is tcm = b/2 . (B4) given by This gives us the correct kinetic and total energies without the Hˆ = Tˆ + Tˆ′ + Vˆ Tˆ + Hˆ ′ . (B2) spurious center-of-mass motion: cm ≡ cm t The quantities that we wish to compute are governed by the Tˆ′ = 1 cm Tˆ ; (B5a) ˆ ˆ h i − t h i “internal” Hamiltonian H′. Since V involves only relative ˆ ˆ ˆ H′ = T ′ + V2B . (B5b) coordinates among the particles, it commutes with Tˆcm; or h i h i h i more generally, To conclude, there are two necessary modifications in the QMC algorithm in order to treat quantum droplets which are ˆ ˆ [Tcm, H′]=0 . (B3) not confined:

In this way, the importance-sampled QMC propagation is de- 1. We let the trial wave function effectively “follow” the termined by Hˆ ′. The motion of the CM in each walker is QMC orbitals, by defining its CM with that of each a separate free diffusion which is governed by Tˆcm. In the QMC orbital. 19

2. For each orbital, we keep track and accumulate all the These modifications in the QMC allows us to obtain the cor- applied CM shifts in order to estimate R2(τ) . This rect density profile and energies of a translationally-invariant gives us the fraction of CM kinetic energyh throughi the Hamiltonian. constant tcm.

[1] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, [20] D. J. Dean and S. E. Koonin, Phys. Rev. C 60, 054306 (1999). and E. A. Cornell, Science 269, 198 (1995). [21] S. Fantoni, A. Sarsa, and K. E. Schmidt, Phys. Rev. Lett. 87, [2] F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringari, Rev. 181101 (2001). Mod. Phys. 71, 463 (1999). [22] S. Zhang and H. Krakauer, Phys. Rev. Lett. 90, 136401 (2003). [3] A. J. Leggett, Rev. Mod. Phys. 73, 307 (2001). [23] H. F. Trotter, Proc. Am. Math. Soc. 10, 545 (1959). [4] S. L. Cornish, N. R. Claussen, J. L. Roberts, E. A. Cornell, and [24] M. Suzuki, Commun. Math. Phys. 51, 183 (1976). C. E. Wieman, Phys. Rev. Lett. 85, 1795 (2000). [25] S. Zhang, in Theoretical Methods for Strongly Correlated Elec- [5] W. M. C. Foulkes, L. Mitas, R. J. Needs, and G. Rajagopal, Rev. trons, edited by D. S´en´echal, A.-M. Tremblay, and C. Bourbon- Mod. Phys. 73, 33 (2001), and also the references therein. nais (Springer, New York, 2003), CRM Series in Mathematical [6] D. M. Ceperley, Rev. Mod. Phys. 67, 279 (1995), and also the Physics, pp. 39–74. references therein. [26] D. R. Hamann and S. B. Fahy, Phys. Rev. B 41, 11352 (1990). [7] W. Krauth, Phys. Rev. Lett. 77, 3695 (1996). [27] E. P. Gross, Nuovo Cimento 20, 454 (1961). [8] P. Gr¨uter, D. Ceperley, and F. Lalo¨e, Phys. Rev. Lett. 79, 3549 [28] E. P. Gross, J. Math. Phys. 4, 195 (1963). (1997). [29] L. P. Pitaevskii, Sov. Phys.–JETP 13, 451 (1961). [9] M. Holzmann and W. Krauth, Phys. Rev. Lett. 83, 2687 (1999). [30] P. A. Ruprecht, M. J. Holland, K. Burnett, and M. Edwards, [10] J. L. DuBois and H. R. Glyde, Phys. Rev. A 63, 023602 (2001). Phys. Rev. A 51, 4704 (1995). [11] J. L. DuBois and H. R. Glyde, Phys. Rev. A 68, 033602 (2003). [31] J. B. McGuire, J. Math. Phys. 5, 622 (1964). [12] M. Ulmke and R. T. Scalettar, Phys. Rev. B 61, 9607 (2000). [32] F. Calogero and A. Degasperis, Phys. Rev. A 11, 265 (1975), [13] R. Blankenbecler, D. J. Scalapino, and R. L. Sugar, Phys. Rev. and also the references therein. D 24, 2278 (1981). [33] J. L. Bohn, J. P. Burke, C. H. Greene, H. Wang, P. L. Gould, [14] G. Sugiyama and S. E. Koonin, Ann. Phys. 168, 1 (1986). and W. C. Stwalley, Phys. Rev. A 59, 3660 (1999). [15] J. Hubbard, Phys. Rev. Lett. 3, 77 (1959). [34] J. Weiner, V. S. Bagnato, S. Zilio, and P. S. Julienne, Rev. Mod. [16] R. D. Stratonovich, Dokl. Akad. Nauk. SSSR 115, 1907 (1957). Phys. 71, 1 (1999). [17] J. E. Hirsch, Phys. Rev. B 28, 4059 (1983). [35] J. W. Negele and H. Orland, Quantum Many-Particle Sys- [18] S. R. White, D. J. Scalapino, R. L. Sugar, E. Y. Loh, J. E. Gu- tems, Advanced Book Classics (Perseus Books, Reading, Mas- bernatis, and R. T. Scalettar, Phys. Rev. B 40, 506 (1989). sacusetts, 1998). [19] S. Zhang, J. Carlson, and J. E. Gubernatis, Phys. Rev. B 55, 7464 (1997).