Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of...

17
Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green 1, 2, * and Dmitrii V. Shalashilin 1, 1 School of Chemistry, University of Leeds, Leeds LS2 9JT, United Kingdom 2 Consiglio Nazionale delle Ricerche, Istituto di Biostrutture e Bioimmagini (CNR-IBB), via Mezzocannone 16, 80134, Napoli, Italy (Current Address) (Dated: July 1, 2019) Computer simulations of many-body quantum dynamics of indistinguishable particles is a chal- lenging task for computational physics. In this paper we demonstrate that the method of coupled coherent states (CCS) developed previously for multidimensional quantum dynamics of distinguish- able particles can be used to study indistinguishable bosons in the second quantisation formalism. To prove its validity, the technique termed here coupled coherent states for indistinguishable bosons (CCSB) is tested on two model problems. The first is a system-bath problem consisting of a tunnelling mode coupled to a harmonic bath, previously studied by CCS and other methods in distinguishable representation in 20 dimensions. The harmonic bath is comprised of identical oscil- lators, and may be second quantised for use with CCSB, so that this problem may be thought of as a bosonic bath with an impurity. The cross-correlation function for the dynamics of the system and Fourier transform spectrum compare extremely well with a benchmark calculation, which none of the prior methods of studying the problem achieved. The second model problem involves 100 bosons in a shifted harmonic trap. Breathing oscillations in the 1-body density are calculated and shown to compare favourably to a multiconfigurational time-dependent Hartree for bosons calcu- lation, demonstrating the applicability of the method as a new formally exact way to study the quantum dynamics of Bose-Einstein condensates. I. INTRODUCTION In the past two decades there has been significant inter- est in systems of indistinguishable bosons, due to exper- imentally produced Bose-Einstein condensates of ultra- cold alkali metal atoms [13]. These condensates, first posited by the eponymous Bose and Einstein in 1924- 25, have permitted macroscopic observations of quan- tum phenomena and lead to a wealth of experimental re- search in areas such as atomic interferometry [4], bosonic Josephson junctions [5, 6], quantum vortices [7, 8] and the generation of solitons [9, 10]. From the theoretician’s point of view, the Gross- Pitaevskii equation (GPE) [11, 12] has been the pre- dominant method used to study Bose-Einstein conden- sates, see for example Refs. [1318] and the review ar- ticles [19, 20]. However the GPE is a mean-field the- ory and as such cannot describe many-body effects in condensates. It also assumes that all bosons occupy a single state at all times, which is not the case dur- ing fragmentation. In recent years, the multiconfigura- tional time-dependent Hartree method for bosons (MCT- DHB) [21, 22] has been used to treat indistinguishable bosons from the standpoint of exact quantum mechan- ics [2333]. A multi-layer version of MCTDHB has also been developed (ML-MCTDHB) [34, 35] that exploits the multi-layer structure to study mixed bosonic systems (for example impurities in Bose-Einstein condensates [3640], * [email protected] [email protected] binary mixtures of Bose-Einstein condensates [41], and solitons [40, 4244]) and bosonic systems where different degrees of freedom may be separated (for example differ- ent spatial locations when bosons are residing in optical lattices [4552]). Before being used to treat indistinguishable bosons, standard MCTDH [53] and ML-MCTDH [54, 55] have been well established theories for treating distinguish- able particles. They are able to solve the time-dependent Schr¨ odinger equation (TDSE) exactly for multiple de- grees of freedom, albeit with basis sets that grow expo- nentially with increased dimensionality. Our own coupled coherent states (CCS) method has also demonstrated its propensity at solving the TDSE for distinguishable par- ticles, with basis sets that scale more favourably with di- mensionality [56, 57]. This is achieved by using randomly sampled trajectory guided coherent states as basis func- tions, although the trade-off for this favourable scaling is that random noise and slow convergence may be present. Noise can cause a decay in auto/cross-correlation func- tions and may be reduced by increasing the number of configurations, or applying a filter diagonalisation tech- nique to extract frequencies [58]. Importance sampling of coherent state basis set initial conditions is also key to the accuracy and efficiency of the CCS approach [59]. In this present work we extend the CCS method to looking at indistinguishable bosons in the second quan- tisation representation, and dub the method coupled co- herent states for indistinguishable bosons (CCSB). Due to the use of coherent states in CCSB and their relation to the creation and annihilation operators of second quan- tisation, together with the fact that systems with a large number of particles tend towards classical behaviour and arXiv:1811.06293v3 [quant-ph] 28 Jun 2019

Transcript of Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of...

Page 1: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

Simulation of the Quantum Dynamics of Indistinguishable Bosons with the CoupledCoherent States Method

James A. Green1, 2, ∗ and Dmitrii V. Shalashilin1, †

1School of Chemistry, University of Leeds, Leeds LS2 9JT, United Kingdom2Consiglio Nazionale delle Ricerche, Istituto di Biostrutture e Bioimmagini (CNR-IBB),

via Mezzocannone 16, 80134, Napoli, Italy (Current Address)(Dated: July 1, 2019)

Computer simulations of many-body quantum dynamics of indistinguishable particles is a chal-lenging task for computational physics. In this paper we demonstrate that the method of coupledcoherent states (CCS) developed previously for multidimensional quantum dynamics of distinguish-able particles can be used to study indistinguishable bosons in the second quantisation formalism.To prove its validity, the technique termed here coupled coherent states for indistinguishable bosons(CCSB) is tested on two model problems. The first is a system-bath problem consisting of atunnelling mode coupled to a harmonic bath, previously studied by CCS and other methods indistinguishable representation in 20 dimensions. The harmonic bath is comprised of identical oscil-lators, and may be second quantised for use with CCSB, so that this problem may be thought ofas a bosonic bath with an impurity. The cross-correlation function for the dynamics of the systemand Fourier transform spectrum compare extremely well with a benchmark calculation, which noneof the prior methods of studying the problem achieved. The second model problem involves 100bosons in a shifted harmonic trap. Breathing oscillations in the 1-body density are calculated andshown to compare favourably to a multiconfigurational time-dependent Hartree for bosons calcu-lation, demonstrating the applicability of the method as a new formally exact way to study thequantum dynamics of Bose-Einstein condensates.

I. INTRODUCTION

In the past two decades there has been significant inter-est in systems of indistinguishable bosons, due to exper-imentally produced Bose-Einstein condensates of ultra-cold alkali metal atoms [1–3]. These condensates, firstposited by the eponymous Bose and Einstein in 1924-25, have permitted macroscopic observations of quan-tum phenomena and lead to a wealth of experimental re-search in areas such as atomic interferometry [4], bosonicJosephson junctions [5, 6], quantum vortices [7, 8] andthe generation of solitons [9, 10].

From the theoretician’s point of view, the Gross-Pitaevskii equation (GPE) [11, 12] has been the pre-dominant method used to study Bose-Einstein conden-sates, see for example Refs. [13–18] and the review ar-ticles [19, 20]. However the GPE is a mean-field the-ory and as such cannot describe many-body effects incondensates. It also assumes that all bosons occupya single state at all times, which is not the case dur-ing fragmentation. In recent years, the multiconfigura-tional time-dependent Hartree method for bosons (MCT-DHB) [21, 22] has been used to treat indistinguishablebosons from the standpoint of exact quantum mechan-ics [23–33]. A multi-layer version of MCTDHB has alsobeen developed (ML-MCTDHB) [34, 35] that exploits themulti-layer structure to study mixed bosonic systems (forexample impurities in Bose-Einstein condensates [36–40],

[email protected][email protected]

binary mixtures of Bose-Einstein condensates [41], andsolitons [40, 42–44]) and bosonic systems where differentdegrees of freedom may be separated (for example differ-ent spatial locations when bosons are residing in opticallattices [45–52]).

Before being used to treat indistinguishable bosons,standard MCTDH [53] and ML-MCTDH [54, 55] havebeen well established theories for treating distinguish-able particles. They are able to solve the time-dependentSchrodinger equation (TDSE) exactly for multiple de-grees of freedom, albeit with basis sets that grow expo-nentially with increased dimensionality. Our own coupledcoherent states (CCS) method has also demonstrated itspropensity at solving the TDSE for distinguishable par-ticles, with basis sets that scale more favourably with di-mensionality [56, 57]. This is achieved by using randomlysampled trajectory guided coherent states as basis func-tions, although the trade-off for this favourable scaling isthat random noise and slow convergence may be present.Noise can cause a decay in auto/cross-correlation func-tions and may be reduced by increasing the number ofconfigurations, or applying a filter diagonalisation tech-nique to extract frequencies [58]. Importance samplingof coherent state basis set initial conditions is also key tothe accuracy and efficiency of the CCS approach [59].

In this present work we extend the CCS method tolooking at indistinguishable bosons in the second quan-tisation representation, and dub the method coupled co-herent states for indistinguishable bosons (CCSB). Dueto the use of coherent states in CCSB and their relation tothe creation and annihilation operators of second quan-tisation, together with the fact that systems with a largenumber of particles tend towards classical behaviour and

arX

iv:1

811.

0629

3v3

[qu

ant-

ph]

28

Jun

2019

Page 2: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

2

the basis in CCSB is guided by classical-like trajectories,suggest that the method will be particularly suited tosuch systems. Indeed, recent semiclassical coherent statework with the Herman-Kluk method on indistinguishablebosons demonstrates this hypothesis [60, 61]. CCSB isfully quantum however, as with standard CCS, and it haspreviously been shown that a local quadratic approxima-tion of the Hamiltonian into the CCS equations yieldsthe coherent state matrix of the Herman-Kluk propa-gator [62]. We anticipate that the CCSB method willprovide a description of many-body dynamics over andabove the mean-field Gross-Pitaevskii approach. Fur-thermore, as CCS has previously shown to be able to pro-vide a similar numerical picture to MCTDH with lowercomputational scaling with dimensionality [63], we antic-ipate that CCSB may be able to do the same with respectto MCTDHB.

To illustrate the suitability of CCSB to problems in-volving indistinguishable bosons, we apply the methodto two model problems. The first model problem con-sists of a bosonic bath with an impurity, demonstratingthat the method is capable of studying multi-componentbosonic systems and opening up the possibility of study-ing multi-atomic Bose-Einstein condensates [64], spinorBose-Einstein condensates [65], dark-bright solitons [66],and Bose-polarons [67]. The second model problem con-sists of a collection of indistinguishable bosons in a har-monic trap, demonstrating the propensity of the methodto study systems of bosons in optical lattices [68], forexample with the Bose-Hubbard model [60, 61, 69], andthe possibility to study bosons in a single well that isdeformed into a double well, such as that in Ref. [22],and observed in experimental bosonic Josephson junc-tions [70, 71].

II. NUMERICAL DETAILS

The CCSB method relies on the machinery of the CCSmethod, which has been derived and presented previouslywhen treating distinguishable particles [56, 57]. A de-scription of the CCS method will be presented below,before a discussion on how the method is modified totreat indistinguishable bosons in the second quantisationrepresentation in CCSB.

A. Coupled Coherent States Working Equations

In the CCS method, the wavefunction is represented asa basis set of trajectory guided coherent states, |z〉. Thecoordinate representation of a coherent state is given by

〈x|z〉 =(γπ

)1/4

exp

(−γ

2(x− q)2 +

i

~p(x− q) +

ipq

2~

),

(1)where q and p are the position and momentum centresof the coherent state, γ is the width parameter of the

coherent state, given by γ = mω/~, with m mass and ωfrequency. In atomic units (which are used throughoutthe paper) m = ω = ~ = 1, thus γ = 1. Coherent statesare eigenstates of the creation and annihilation operatorsrespectively

〈z| a† = 〈z| z∗ (2a)

a |z〉 = z |z〉 , (2b)

where the creation and annihilation operators are givenby

a† =1√2

(q − ip) (3a)

a =1√2

(q + ip) . (3b)

The eigenvalues of Eqs. 2a and 2b, z∗ and z, can be usedto label a coherent state, and from Eqs. 3a and 3b it canbe seen they are given by

z∗ =1√2

(q − ip) (4a)

z =1√2

(q + ip) . (4b)

An important consequence of the above is that one maywrite a Hamiltonian in terms of creation and annihila-tion operators rather than position and momentum op-erators. A normal ordered Hamiltonian may then be ob-tained when the creation operators precede the annihila-tion ones

H(q, p) = H(a, a†) = Hord(a†, a). (5)

From this, matrix elements of the Hamiltonian are simpleto calculate in a coherent state basis

〈z′|Hord(a†, a)|z〉 = 〈z′|z〉Hord(z′∗, z), (6)

where the overlap 〈z′|z〉 is given by

〈z′|z〉 = exp

(z′∗z − z′∗z′

2− z∗z

2

). (7)

The wavefunction ansatz in CCS is given by

|Ψ(t)〉 =

K∑k=1

Dk(t)eiSk(t) |zk(t)〉 , (8)

where the sum is over K configurations, Dk is a timedependent amplitude and Sk is the classical action. Theclassical action in coherent state notation is given by

Sk =

∫ [i

2(z∗k zk − z∗kzk)−Hord(z∗k, zk)

]dt. (9)

The wavefunction is propagated via the time-dependenceof the coherent state basis vectors, amplitudes and ac-tion. The coherent states are guided by classical trajec-tories, and evolve according to Hamilton’s equation

zk = −i∂Hord(z∗k, zk)

∂z∗k. (10)

Page 3: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

3

The time-dependence of the amplitudes may be found viasubstitution of Eq. 8 into the time-dependent Schrodingerequation and closing with a coherent state basis bra:

K∑l=1

〈zk|zl〉 eiSldDl

dt= −i

K∑l=1

〈zk|zl〉 eiSlDlδ2H ′ord(z∗k, zl),

(11)where the δ2H ′ord(z∗k, zl) term is

δ2H ′ord(z∗k, zl) = Hord(z∗k, zl)−Hord(z∗l , zl)− izl(z∗k− z∗l ).(12)

Finally, the time-dependence of the classical action isstraightforwardly calculated from Eq. 9.

B. Second Quantisation and CCSB

CCS works for Hamiltonians that can be expressed viacreation and annihilation operators in the normal orderedform as illustrated in Eq. 6. In second quantisation, theHamiltonian of a system of bosons also appears in a nor-mal ordered form, therefore no modifications of the CCSworking equations are required for treating indistinguish-able bosons with CCSB. The only difference is that thecoherent state basis functions are used to represent parti-cle number occupations of quantum states in the secondquantisation formalism, as opposed to individual parti-cles in the distinguishable first quantisation representa-tion.

In the second quantisation representation, multiparti-cle states are described in terms of an occupation numbern(α) that describes the number of particles belonging toa particular quantum state |α〉. A Fock state describesthe set of occupation number states

|n〉 =

Ω∏α=0

|n(α)〉 = |n(0), n(1), . . . , n(Ω)〉 , (13)

and may be generated by successive application of cre-ation operators on the vacuum state |0〉

|n(0), n(1), . . . , n(Ω)〉 =(a(0)†)n(0)

√n(0)!

(a(1)†)n(1)

√n(1)!

. . .

(a(Ω)†)n(Ω)

√n(Ω)!

|0(0), 0(1), . . . , 0(Ω)〉 .

(14)

In CCSB, the multidimensional version of the CCS wave-function representation is used as a basis set expansionfor Fock states

|n〉 =

K∑k=1

Dk(t)eiSk(t) |zk(t)〉 , (15)

which is exactly analogous to Eq. 8. The only differenceis the multidimensional coherent state |zk〉 is a product of

coherent states that describe occupations of each quan-tum state |α〉

|zk〉 =

Ω∏α=0

|z(α)〉 . (16)

Therefore any wavefunction in the basis of Fock statescan be equivalently represented in the basis of coherentstates. The Hamiltonian of a system of indistinguishablebosons can be second quantised and presented in terms

of 1-body h(Q), 2-body W (Q,Q′), and creation and an-nihilation operators as

H =∑α,β

〈α|h|β〉 a(α)†a(β)

+1

2

∑α,β,γ,ζ

〈α, β|W |γ, ζ〉 a(α)†a(β)†a(ζ)a(γ),(17)

where |α〉, |β〉, |γ〉, and |ζ〉 are quantum states. This con-veniently gives a second quantised Hamiltonian in normalordered form, which is required by CCSB. In the follow-ing sections CCSB is applied to two model problems toillustrate its ability to study fully quantum bosonic prob-lems and compare to numerically exact results.

III. APPLICATION 1: DOUBLE WELLTUNNELLING PROBLEM

The first application of CCSB is to an M -dimensionalmodel Hamiltonian that consists of an (M − 1)-dimensional harmonic bath, coupled to a 1-dimensionaltunnelling mode governed by an asymmetric double wellpotential. This a system-bath problem, which may alsobe thought of as a bosonic bath with an impurity, previ-ously studied in distinguishable representation with lin-ear coupling of the bath to the system by matching pur-suit split-operator Fourier transform (MP/SOFT) [72],standard CCS [73], a trajectory guided configuration in-teraction (CI) expansion of the wavefunction [74], anadaptive trajectory guided (aTG) scheme [75], Gaus-sian process regression (GPR) [76], and a basis expan-sion leaping multi-configuration Gaussian (BEL MCG)method [77]. It has also been studied with quadraticcoupling of the bath to the system by MP/SOFT [72],standard CCS [73], trajectory guided CI [74], aTG [75]and a 2-layer version of CCS (2L-CCS) [78]. A bench-mark calculation for the quadratic coupling case has alsobeen proposed in recent work [79], using a relatively sim-ple wavefunction expansion in terms of particle in a boxwavefunctions for the tunnelling mode, and harmonic os-cillator wavefunctions for the harmonic bath. The sizeof the calculation in Ref. [79] was greatly reduced byexploiting the indistinguishability of the bath configu-rations, the first time this had been considered, and awell converged result was achieved, prompting the ideaof CCSB. The quadratic coupling case is the one we con-sider in this application.

Page 4: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

4

The Hamiltonian is given in distinguishable represen-tation by

H =p(1)2

2− q(1)2

2+q(1)4

16η+

P2

2+

(1 + λq(1)

)Q2

2, (18)

where (q(1), p(1)) are the position and momentum op-erators of the 1-dimensional system tunnelling mode,and (Q, P) are the position and momentum operatorsof the (M − 1)-dimensional harmonic bath modes, with

Q =∑Mm=2 q

(m) and P =∑Mm=2 p

(m). The coupling be-tween system and bath is given by the constant λ, whilstη determines the well depth.

In previous work [72–75, 78, 79], the parameters λ =0.1 and η = 1.3544 have been used in a 20-dimensional(M = 20) problem, which we also consider. The ini-tial wavefunction |Ψ(0)〉 is a multidimensional Gaussianwavepacket, with initial position and momentum centresfor the tunnelling mode q(1)(0) = −2.5 and p(1)(0) = 0.0,

and for the bath modes q(m)(0) = 0.0 and p(m)(0) = 0.0∀ m.

As the bath oscillators have the same initial condi-tions and the same frequency, they can be thought ofas indistinguishable, and the bath part of the Hamilto-nian may be second quantised for use with CCSB. Asthe tunnelling mode is not part of this indistinguishablesystem, the portion of the Hamiltonian that describesit will not be second quantised. However, this will notpose a problem as the dynamical equations are identicalfor CCS and CCSB, the only subtlety is the interpreta-tion of the coherent state basis vectors |z〉 as will be dis-cussed below. Using the definition of a second quantisedHamiltonian in Eq. 17, and the definition of coherentstates as eigenstates of the creation and annihilation op-erators, Eq. 18 may be written in normal-ordered form asHord(a†, a), for which the coherent state matrix element〈zk|Hord(a†, a)|zl〉 = 〈zk|zl〉Hord(z∗k, zl), where

Hord(z∗k, zl) =− 1

2

(z

(m=1)∗2k + z

(m=1)2

l

)+

1

64η

(z

(m=1)∗4k + z

(m=1)4

l + 4z(m=1)∗3k z

(m=1)l + 4z

(m=1)∗k z

(m=1)3

l

+6z(m=1)∗2k z

(m=1)2

l + 12z(m=1)∗k z

(m=1)l + 6z

(m=1)∗2k + 6z

(m=1)2

l + 3)

+

Ω∑α=0

z(2α)∗k z

(2α)l ε(2α) +

λ

2

Ω∑α,β=0

z(2α)∗k z

(2β)l Q(2α,2β)2

(z

(m=1)∗k + z

(m=1)l

).

(19)

The quantum states |α〉 and |β〉 in Eq. 19 are those ofthe harmonic oscillator with α and β numbers of quanta,ε(α) is the eigenvalue for |α〉, and the position and mo-mentum operators of the tunnelling mode have explic-itly been labelled with (m = 1) to distinguish themfrom the α labelling scheme of the second quantised bathmodes. A full derivation of this, alongside evaluation of

the Q(2α,2β)2

matrix element is shown in Appendix A.Note that only even harmonic oscillator levels are re-quired due to all bath modes initially residing in theground level, as previously assumed [79], and the bathhaving quadratic coupling to the system meaning onlyeven harmonic oscillator levels will be occupied.

The multidimensional coherent state basis vector |z〉 isrepresented as

|z〉 = |z(m=1)〉 ×Ω∏α=0

|z(2α)〉 , (20)

where |z(m=1)〉 is a basis function for the tunnelling modeand |z(2α)〉 is a basis function for the second quantisedbath modes. The determination of initial conditions forthese coherent state basis functions, as well as the valuesof the initial amplitudes is shown in the following section.

A. Initial Conditions for Application 1

The initial coherent state basis functions for the tun-nelling mode are sampled from a Gaussian distributioncentered around the initial tunnelling mode coordinatesand momenta, as in previous works [73, 78]

f(z(m=1)) ∝ exp

(−σ(m=1)

∣∣∣z(m=1) − z(m=1)(0)∣∣∣2) ,

(21)where σ(m=1) is a parameter governing the width of thedistribution.

Sampling the initial coherent states for the bath can beperformed by obtaining a probability distribution fromthe square of the coherent state representation of theinitial bath Fock state. The initial bath Fock state isequal to

|n〉 =

Ω∏α=0

|n(2α)〉

= |n(2α=0), n(2α=2), . . . , n(2α=2Ω)〉= |(M − 1), 0, . . . , 0〉 ,

(22)

where there are M − 1 bath oscillators all in the groundharmonic oscillator state. Using the representation of a

Page 5: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

5

coherent state in a basis of Fock states

|z〉 = e−|z|2

2

∑n(α)

zn(α)√

(n(α)!)|n(α)〉 (23)

the following may be obtained

| 〈z(2α)|n(2α)〉 |2 =e−|z

(2α)|2 (|z(2α)|2)n(2α)

πn(2α)!, (24)

where the value of π has appeared to enforce normalisa-tion. This resembles a Poissonian distribution, however|z(2α)|2 is continuous so a gamma distribution is used

instead

f(|z(2α)|2) ∝(|z(2α)|2

)n(2α)

e−|z(2α)|2

σ(2α)

Γ(n(2α) + 1)(σ(2α)

)n(2α)+1, (25)

where σ(2α) is a compression parameter controlling thewidth of the distribution, and Γ is the gamma functionthat is calculated using n(2α) +1 because Γ(n) = (n−1)!.

The gamma distribution will be centred aroundσ(2α)n(2α), however | 〈z(2α)|n(2α)〉 |2 should be centredaround |z(2α)|2 = n(2α) as its maximum is found by

d| 〈z(2α)|n(2α)〉 |2

d|z(2α)|2=

1

πn(2α)!

(−e−|z

(2α)|2(|z(2α)|2

)n(2α)

+ n(2α)e−|z(2α)|2

(|z(2α)|2

)n(2α)−1)

= 0. (26)

Fortunately, this is not an issue, as when n(2α) = 0for states 2α > 0, the distribution will be centredaround 0 irrespective of the compression parameter,and for n(2α=0) = M − 1 a compression parameter ofσ(2α=0) = 1.0 is used. As we are not constrained by achoice of compression parameter σ(2α>0) for the states2α > 0, we are free to alter it to influence the accuracyof the calculation, and the final result presented in thefollowing section uses σ(2α>0) = 100. The affect of alter-ing this parameter is discussed in Sec. III C.

The initial amplitudes are calculated by projection ofthe initial basis onto the initial wavefunction with theaction set to zero

〈zk(0)|Ψ(0)〉 =

K∑l=1

Dl(0) 〈zk(0)|zl(0)〉 . (27)

The overlap of the initial coherent state basis with theinitial wavefunction can be decomposed to

〈zk(0)|Ψ(0)〉 = 〈z(m=1)k (0)|Ψ(m=1)(0)〉 〈

Ω∏α=0

z(2α)k (0)|n)〉 .

(28)The coherent state overlap with initial tunnelling mode

wavefunction 〈z(m=1)k (0)|Ψ(m=1)(0)〉 can be calculated

via a Gaussian overlap, Eq. 7, using the initial positionsand momenta for the tunnelling mode q(m=1)(0) = −2.5and p(m=1)(0) = 0.0. The coherent state overlap withinitial bath Fock state can be calculated by once moreusing the coherent state representation in a basis of Fockstates, Eq. 23

〈Ω∏α=0

z(2α)k (0)|n)〉 =

[Ω∏α=0

e−|z(2α)k

(0)|2

2

](z

(2α=0)∗k (0))M−1√

(M − 1)!.

(29)

B. Results and Comparison to Other Methods

The quantity of interest used to assess the performanceof CCSB and compare it to previous methods of study-ing the problem [72–75, 78, 79] is the cross-correlationfunction (CCF). This is the overlap between the wave-function at time t and the mirror image of the initialwavepacket, |Ψ(0)〉, i.e. 〈Ψ(0)|Ψ(t)〉. The mirror im-age of the initial state has coordinates for the tunnellingmode of q(1)(0) = +2.5 and p(1)(0) = 0.0, with bathmodes in the ground harmonic level. It is located inthe upper well of the asymmetric double well tunnellingpotential, therefore non-zero values of the CCF are in-dicative of tunnelling. The spectrum of the CCF is alsopresented via a Fourier transform (FT) of the real partof the CCF.

The results of the CCSB calculation compared withprevious methods of studying the 20D, λ = 0.1 case [72–75, 78, 79] is shown in Fig. 1 with the absolute values ofthe CCFs in Fig. 1(i), and FT spectra in Fig. 1(ii). As canbe seen from these two figures, the CCSB results com-pare extremely favourably to the benchmark calculation,with much closer agreement than prior methods. Previ-ously the trajectory guided CI expansion was the closestresult to the benchmark, due to its basis set expansionof time-independent basis functions used to represent ex-cited state configurations being similar to the benchmarkapproach. However the CCF still differed from the bench-mark, with noticeable differences occurring after 25 a.u.,possibly due to approximations used in sampling the po-tential energy surface, despite the FT obtaining splittingof the higher energy peaks that no prior method man-aged. For this present CCSB calculation, there is nosignificant degradation of the calculation at t > 25 a.u.as with the other methods, and the splitting of the highenergy peaks is very well reproduced. As was alluded toin Ref. [78], for this Hamiltonian a detailed description

Page 6: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

6

(i)

0

0.05

0.1

0 20 40 60 80 100 120

(f)

Abs(C

CF

)

Time (a.u.)

BenchmarkCCSB

0

0.05

0.1

(e)

Ab

s(C

CF

)

Benchmark2L-CCS

0

0.05

0.1

(d)

Abs(C

CF

)

Benchmark CCS

0

0.05

0.1

(c)

Abs(C

CF

)

BenchmarkAdaptive Trajectory Guided

0

0.05

0.1

(b)

Abs(C

CF

)

BenchmarkTrajectory Guided CI Expansion

0

0.05

0.1

(a)A

bs(C

CF

)

BenchmarkMP/SOFT

(ii)

0

0.5

1

1.5

7 8 9 10 11 12 13 14

(f)

I(ω

)

ω

BenchmarkCCSB

0

0.5

1

1.5

(e)

I(ω

)

Benchmark2L-CCS

0

0.5

1

1.5

(d)

I(ω

)

BenchmarkCCS

0

0.5

1

1.5

(c)

I(ω

)

BenchmarkAdaptive Trajectory Guided

0

0.5

1

1.5

(b)

I(ω

)

BenchmarkTrajectory Guided CI Expansion

0

0.5

1

1.5

(a)

I(ω

)

BenchmarkMP/SOFT

FIG. 1: Comparison of (i) absolute values of the cross-correlation functions and (ii) Fourier transforms of the realpart of the cross-correlation functions for different methods (red, solid) of studying Eq. 18 with M = 20. λ = 0.1

parameters relative to the benchmark [79] (black, dotted): (a) MP/SOFT [72], (b) Trajectory Guided CIExpansion [74], (c) aTG [75], (d) CCS [73], (e) 2L-CCS [78], (f) CCSB (present work).

Page 7: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

7

of the bath is required for accurate propagation, whichis achieved in CCSB by taking account of the symmetryof the Hamiltonian.

The CCSB calculation uses K = 4000 configurationsand Ω = 5 even harmonic oscillator levels in the bathbasis. The dimensionality of this problem has thereforebeen reduced from 20 to 6. The influence on the CCSBcalculation of altering these parameters, as well as thecompression parameter chosen of σ(2α>0) = 100 is shownin the following section.

C. Numerical Accuracy and Convergence

Using an approach first presented in Ref. [78] to illus-trate the accuracy and convergence of a method studyingApplication 1 with respect to the benchmark calculation,we define an error parameter χ as

χ =

∫|Abs(〈Ψ(0)|Ψ(t)〉)bench

−Abs(〈Ψ(0)|Ψ(t)〉)CCSB|dt,(30)

which indicates the cumulative error of the absolute valueof the cross-correlation function of the CCSB methodcompared to the benchmark. This is shown in Fig. 2(i)for different values of σ(2α>0), K and Ω, in panels (a),(b) and (c).

As with CCS, the CCSB method does not conservethe norm 〈Ψ(t)|Ψ(t)〉 by default due to the use of a basisconsisting of a superposition of coherent states [57]. How-ever, this can be a useful property as the extent of normconservation can be used to determine the accuracy andreliability of a propagation. Another important quantityfor CCSB to conserve is the total particle number

N = 〈Ψ(t)|Ω∑α=0

a(α)†a(α)|Ψ(t)〉

=

K∑k,l=1

Ω∑α=0

D∗kDlei(Sl−Sk) 〈zk|zl〉 z(α)∗

k z(α)l ,

(31)

which for Application 1 amounts to the number of os-cillators in the bath, N = (M − 1) = 19. Plots of thenorm and particle number conservation for different val-ues of σ(2α>0), K, and Ω, (as in Fig. 2(i)), are shownin Fig. 2(ii), with the value of the norm given by thedashed/dotted lines without circles and the particle num-ber by the dashed/dotted lines with circles. It can be seenthat the values of the norm and particle number followeach other closely for all calculations, and we will discussthe specific cases in the following.

Firstly, considering panel (a) in Figs. 2(i) and 2(ii),both K and Ω are held fixed whilst σ(2α>0) is varied.It can be seen that the quality of the calculation withrespect to the error term χ, and the conservation ofthe norm and particle number improves with increasing

σ(2α>0). Further increase of σ(2α>0) results in a numeri-cally unstable propagation, as the value of the norm andparticle number explodes as the basis is overcompressed.This suggests that appropriate choice of σ(2α>0) is nec-essary for the initial sampling of the coherent states, asa value that is too small leads to errors due to the co-herent states spreading too quickly, whilst a value thatis too large leads to numerical instability.

Secondly, considering panel (b) in Figs. 2(i) and 2(ii),the value of K is varied whilst Ω is held constant. Thevalue of σ(2α>0) was chosen based on the criteria pre-sented in the previous paragraph, with larger values ofσ(2α>0) for smaller values of K. This phenomenon hasbeen noted in previous studies with CCS, where largercompression parameters are necessary for basis sets withfewer configurations, see Ref. [59] for further details. Theerror χ decreases with increasing K, as Monte-Carlonoise causing decay of the cross-correlation function de-creases with increasing number of configurations. How-ever, χ does not remain at 0 for the duration of the calcu-lation, and the K = 4000 propagation is only equivalentto the K = 3000 calculation for the first 30 a.u. This in-dicates the slow convergence of the method as mentionedin the introduction. However, we can regard the accuracyof the K = 4000 calculation as sufficient for this applica-tion as we are able to obtain an accurate FT spectrumwith correct frequencies, alongside the good conservationof norm and particle number compared to the other cal-culations with fewer configurations.

Finally, considering panel (c) in Figs. 2(i) and 2(ii), thevalue of Ω is varied whilst K is held constant. The valueof σ(2α>0) was again chosen based on the criteria pre-sented above, with larger values possible with increasedΩ. It can be seen that altering the value of Ω has asmall effect on the accuracy of the calculation (note thedifference in y axis values for χ compared to panels (a)and (b)), and the value Ω = 5 was deemed to result in astable enough propagation.

IV. APPLICATION 2: INDISTINGUISHABLEBOSONS IN A DISPLACED HARMONIC TRAP

The second application of CCSB is to a system com-posed purely of indistinguishable bosons, with N inter-acting bosons placed in a harmonic trap displaced fromthe origin, with N = 100 used in the present applica-tion. The oscillations in the density are calculated andcompared to MCTDHB [21, 22] calculations (performedby the authors, using the MCTDHB package [80]). TheHamiltonian (in dimensionless units and distinguishablerepresentation) for this problem consists of a shifted har-monic potential and a 2-body interaction term

H =P2

2+

(Q− ξ)2

2+ W (Q,Q′), (32)

where Q and P are the position and momentum opera-tors of the N bosons, ξ = 2.1 is a parameter that shifts

Page 8: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

8

(i)

0

0.2

0.4

0.6

0.8

1

0 20 40 60 80 100 120

(c)

χ

Time (a.u.)

K=4000, Ω=3, σ(2α>0)=5K=4000, Ω=4, σ(2α>0)=20K=4000, Ω=5, σ(2α>0)=100

0

0.5

1

1.5

2

2.5

3

(b)

χ

K=1000, Ω=5, σ(2α>0)=50000K=2000, Ω=5, σ(2α>0)=1000K=3000, Ω=5, σ(2α>0)=200K=4000, Ω=5, σ(2α>0)=100

0 0.2 0.4 0.6 0.8

1 1.2 1.4

(a)

χ

K=4000, Ω=5, σ(2α>0)=10K=4000, Ω=5, σ(2α>0)=20K=4000, Ω=5, σ(2α>0)=50K=4000, Ω=5, σ(2α>0)=100

(ii)

0.99

1

1.01

1.02

1.03

0 20 40 60 80 100 120

18.8

19

19.2

19.4

19.6(c)

Norm

Part

icle

Num

be

r

Time (a.u.)

K=4000, Ω=3, σ(2α>0)

=5K=4000, Ω=4, σ

(2α>0)=20

K=4000, Ω=5, σ(2α>0)

=100

0.98

1

1.02

1.04

1.06

1.08

18.5

19

19.5

20

20.5(b)

Norm

Pa

rtic

le N

um

ber

K=1000, Ω=5, σ(2α>0)

=50000K=2000, Ω=5, σ

(2α>0)=1000

K=3000, Ω=5, σ(2α>0)

=200K=4000, Ω=5, σ

(2α>0)=100

1

1.02

1.04

1.06

1.08

1.1

1.12

1.14

19

19.5

20

20.5

21

21.5(a)

No

rm

Part

icle

Nu

mber

K=4000, Ω=5, σ(2α>0)

=10K=4000, Ω=5, σ

(2α>0)=20

K=4000, Ω=5, σ(2α>0)

=50K=4000, Ω=5, σ

(2α>0)=100

FIG. 2: (i) Cumulative error χ (defined in Eq. 30) of the CCSB method with respect to the benchmark [79] fordifferent values of: (a) compression parameter for coherent state sampling of the bath basis levels with zero initial

occupation σ(2α>0), (b) configurations K, and (c) even harmonic oscillator levels in bath basis Ω . (ii) Norm(dashed/dotted lines without circles) and particle number (dashed/dotted lines with circles) of CCSB calculations

with different values of: (a) σ(2α>0), (b) K, and (c) Ω. Note that in panels (b) and (c) for both (i) and (ii) the valueof σ(2α>0) changes as well as K and Ω. This is addressed in the text.

the harmonic potential from the origin, and W is the2-body interaction, given by the contact interaction

W (Q,Q′) = λ0δ(Q−Q′). (33)

The constant λ0 controls the strength of the interac-tion, with values of λ0 = 0.001 and λ0 = 0.01 used inthe present application, whilst δ(Q − Q′) is the Diracdelta function. As with Application 1, the Hamiltonianin Eq. 32 must be second quantised and normal-orderedbefore it can be used with CCSB, with

Hord(z∗k, zl) =

Ω∑α=0

ε(α)z(α)∗k z

(α)l −

Ω∑α,β=0

ξQ(α,β)z(α)∗k z

(β)l +

Ω∑α=0

ξ2

2z

(α)∗k z

(α)l +

1

2

Ω∑α,β,γ,ζ=0

λ0δ(α,β,γ,ζ)z

(α)∗k z

(β)∗k z

(ζ)l z

(γ)l .

(34)

The derivation of the above, and evaluation of the matrixelements Q(α,β) and δ(α,β,γ,ζ) is shown in Appendix B.The initial sampling of the coherent states and ampli-tudes is performed in a similar manner to the secondquantised bath of Application 1, and is shown in the fol-lowing section.

A. Initial Conditions for Application 2

The initial Fock state for the system includes all bosonsin the ground harmonic state

|n〉 =

Ω∏α=0

|n(α)〉 = |n(0), n(1), . . . , n(Ω)〉 = |100, 0, . . . , 0〉 .

(35)As with the second quantised bath of Application 1,

the coherent states are sampled via a gamma distributionlike in Eq. 25. The ground state with initial occupationn(α=0) = 100 is sampled with compression parameter

Page 9: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

9

σ(α=0) = 1.0 to ensure the distribution is centred in thecorrect place, whilst once more we are free to choose thecompression parameter for the excited states with initialoccupation n(α>0) = 0. Values of σ(α>0) = 109 for λ0 =0.001 and σ(α>0) = 107 for λ0 = 0.01 are used, withfull details for the determination of these compressionparameters shown in the following section.

Initial amplitudes are calculated by projecting the ba-sis onto the initial Fock state in Eq. 35

〈zk(0)|n〉 =

K∑l=1

Dl(0) 〈zk(0)|zl(0)〉 , (36)

where

〈zk(0)|n〉 = 〈Ω∏α=0

z(α)k (0)|n)〉

=

[Ω∏α=0

e−|z(α)k

(0)|2

2

](z

(α=0)∗k (0))100

√100!

.

(37)

For the MCTDHB calculations, the initial orbitalswere constructed from eigenfunctions of the unshiftedtrap (ξ = 0), with the coefficient of one of the orbitals setto 1, whilst the rest were set to 0. This was chosen forthe initial conditions of the MCTDHB calculations ratherthan propagation in imaginary time to obtain the initialorbitals and coefficients of the ground state [21, 22], as wecurrently do not have an analogous procedure for CCSBdue to the instability of trajectories when propagatingin imaginary time [81]. This way we ensure the initialconditions for both methods are the same, and we aretesting the propagation accuracy of both methods. Infuture work we will look at the effect of initial conditionson CCSB, and its comparison to MCTDHB.

B. Results and Comparison to MCTDHB

The dynamics are followed by observing the evolutionof the density matrix over the course of the calculation,which in CCSB can be evaluated as

ρ(α,β) = 〈Ψ|a(α)†a(β)|Ψ〉

=

K∑k,l=1

D∗kDlei(Sl−Sk) 〈zk|zl〉 z(α)∗

k z(β)l .

(38)

As the creation and annihilation operators have differentinterpretations in CCSB and MCTDHB (acting on quan-tum states vs orbitals), the density matrix in this formalso has a different interpretation. Therefore, to comparethe two methods on the same footing, the 1-body densityis evaluated as a function of position, which for CCSB in

this application can be calculated by the following

ρ(Q) = 〈α|ρ(α,β)|β〉

=

Ω∑α,β=0

1√2αα!

(1

π

)1/4

e−Q2/2Heα(Q)ρ(α,β)

× 1√2ββ!

(1

π

)1/4

e−Q2/2Heβ(Q).

(39)

This 1-body density is shown as a function of positionand time in Fig. 3 for interaction strengths λ0 = 0.001in (i) and λ0 = 0.01 in (ii), with the MCTDHB cal-culations in panel (a) and CCSB calculations in panel(b). The MCTDHB calculations use 1 orbital for theλ0 = 0.001 case, and 3 orbitals for the λ0 = 0.01 case,labelled as MCTDHB(1) and MCTDHB(3), respectively.The CCSB calculations use K = 150 configurations forboth the λ0 = 0.001 and λ0 = 0.01 cases, with Ω = 26harmonic oscillator levels for the λ0 = 0.001 case andΩ = 25 harmonic oscillator levels for the λ0 = 0.01 case,with the compression parameters σ(α>0) as mentioned inthe previous section. It can be seen that the MCTDHBand CCSB calculations of the 1-body density comparewell to one another, illustrating the oscillations in thebosonic cloud due to the trap displacement.

As well as the density being subject to oscillations as afunction of the trap displacement, it also exhibits breath-ing oscillations, which may not be immediately apparentfrom Fig. 3. To illustrate these, we plot the variance inthe 1-body density as a function of time

〈Q〉 =

∫Q2ρ(Q) dQ−

(∫Qρ(Q) dQ

)2

, (40)

which is shown in Fig. 4, with the interaction strengthsλ0 = 0.001 in (i) and λ0 = 0.01 in (ii). It can immedi-ately be seen that the stronger interaction strength pro-duces larger breathing oscillations, which is not as easyto see in Fig. 3. Furthermore, the breathing oscillationsin Fig. 4 serve to determine the convergence of the meth-ods: MCTDHB with respect to the number of orbitals;and CCSB with respect to σ(α>0), K, and Ω. A portionof the peak of the density variance at ∼ 8 a.u. is high-lighted to clearly illustrate any discrepancies that maybe difficult to distinguish. As with Application 1, theaccuracy and convergence of the CCSB calculation mayalso be determined from the conservation of norm andparticle number, which is shown in Fig. 5.

For λ0 = 0.001, the MCTDHB calculations using 1, 3,and 4 orbitals demonstrate equivalent breathing oscilla-tions in panel (a) of Fig. 4(i), whilst there appears to bean anomalous result for MCTDHB with 2 orbitals. Theauthors are unsure of the reason for this, which may bebest left for a future MCTDHB study, however we regardthe MCTDHB(1) as being fully converged, and use thisto compare to the CCSB calculations in panels (b), (c),and (d), as well as in Fig. 3(i). An MCTDHB calcula-tion with 1 orbital is equivalent to the GPE [82], so thesedynamics are at the mean-field level.

Page 10: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

10

(i) (ii)

FIG. 3: Space-time representation of the evolution of the 1-body density for (a) MCTDHB and (b) CCSBcalculations of Application 2 with interaction strengths (i) λ0 = 0.001 and (ii) λ0 = 0.01.

In panel (b) of Fig. 4(i) we alter the value of the com-pression parameter σ(α>0), whilst keeping the numberof configurations fixed at K = 150, and the number ofharmonic oscillator levels fixed at Ω = 26. There ap-pears to be little difference in the breathing oscillationsfor different values of σ(α>0), although the highlightedpeak at ∼ 8 .a.u. demonstrates small discrepancies be-tween the MCTDHB calculation and σ(α>0) = 106 andσ(α>0) = 107, whilst σ(α>0) = 108 and σ(α>0) = 109 su-perimpose on the MCTDHB result. The conservation ofnorm and particle number for different values of σ(α>0)

is shown in panel (a) of Fig. 5(i), where σ(α>0) = 108 andσ(α>0) = 109 superimpose upon a value of the norm of1, and particle number 100, as should be expected. We

keep σ(α>0) = 109 for the remaining calculations. This ismuch larger than the compression parameter used in Ap-plication 1, however much fewer configurations are usedin this application, and the compression parameter nec-essary also depends upon the problem studied, and howthe dynamics affect the motion of the basis.

In panel (c) of Fig. 4(i) we alter the value of K whilstkeeping σ(α>0) and Ω fixed. Altering the value of thecompression parameter for different values of K was notnecessary like in Application 1, as a stable basis was ableto be formed. We observe very little discrepancy betweenthe different CCSB calculations and MCTDHB, and thenorm and particle number conservation for K = 150 andK = 200 are very similar in panel (b) of Fig. 5(i). We

Page 11: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

11

(i)

0.5

0.51

0.52

0.53

0.54

0 2 4 6 8 10

(d)

<Q

>

Time (a.u.)

MCTDHB(1)K=150, Ω=23, σ(α>0)=109

K=150, Ω=24, σ(α>0)=109

K=150, Ω=25, σ(α>0)=109

K=150, Ω=26, σ(α>0)=109

0.5

0.51

0.52

0.53

0.54

(c)

<Q

>

MCTDHB(1)K=100, Ω=26, σ(α>0)=109

K=150, Ω=26, σ(α>0)=109

K=200, Ω=26, σ(α>0)=109

0.5

0.51

0.52

0.53

0.54

(b)

<Q

>

MCTDHB(1)K=150, Ω=26, σ(α>0)=106

K=150, Ω=26, σ(α>0)=107

K=150, Ω=26, σ(α>0)=108

K=150, Ω=26, σ(α>0)=109

0.5

0.51

0.52

0.53

0.54

(a)

<Q

>MCTDHB(1)MCTDHB(2)MCTDHB(3)MCTDHB(4)

(ii)

0.5

0.6

0.7

0.8

0.9

0 2 4 6 8 10

(d)

<Q

>

Time (a.u.)

MCTDHB(3)K=150, Ω=20, σ(α>0)=107

K=150, Ω=22, σ(α>0)=107

K=150, Ω=24, σ(α>0)=107

K=150, Ω=25, σ(α>0)=107

0.5

0.6

0.7

0.8

0.9

(c)

<Q

>

MCTDHB(3)K=100, Ω=25, σ(α>0)=107

K=150, Ω=25, σ(α>0)=107

K=200, Ω=25, σ(α>0)=107

0.5

0.6

0.7

0.8

0.9

(b)

<Q

>

MCTDHB(3)K=150, Ω=25, σ(α>0)=106

K=150, Ω=25, σ(α>0)=107

K=150, Ω=25, σ(α>0)=108

K=150, Ω=25, σ(α>0)=109

0.5

0.6

0.7

0.8

0.9

(a)

<Q

>

MCTDHB(1)MCTDHB(2)MCTDHB(3)MCTDHB(4)

FIG. 4: Variance of the 1-body density 〈Q〉 for Application 2 with two-body interaction strength (i) λ0 = 0.001 and(ii) λ0 = 0.01. In the (a) panels MCTDHB calculations with different numbers of orbitals are shown, and the

converged result is used to compare to CCSB calculations with different values of (b) compression parameter forcoherent state sampling of the harmonic oscillator levels with zero initial occupation σ(α>0), (c) configurations K

and (d) harmonic oscillator levels in the basis Ω.

therefore regard K = 150 as being fully converged.

In panel (d) of Fig. 4(i) we alter the value of Ω andkeep K and σ(α>0) fixed. As with the above, altering thevalue of σ(α>0) was not necessary as a stable basis wasable to be formed each time. Larger discrepancies be-tween the CCSB calculations and the MCTDHB resultare seen with in this panel, indicating that the choiceof Ω has the largest influence on this calculation. Theresults for Ω = 25 and Ω = 26 superimpose, indicatingthat the calculation is converged by this point. The con-servation of norm and particle number for both of thesecalculations is also similar in panel (c) of Fig. 5(i).

Turning to the larger interaction strength of λ0 = 0.01,we follow the same approach as above in determining theaccuracy and convergence of the calculations. Initially,in panel (a) of Fig. 4(i) the MCTDHB calculations with1 and 2 orbitals exhibit minor differences to those with3 and 4 orbitals, shown in the highlighted portion of thefigure, so we regard the MCTDHB(3) calculation as our

fully converged reference point. At this level of interac-tion strength, an above mean-field description is thereforenecessary. We admit that the discrepancy between theseMCTDHB calculations is not very large, however a sim-ilar study in Ref. [31] also illustrated minor differencesin breathing dynamics for MCTDHB calculations withdifferent numbers of orbitals.

In panel (b) of Fig. 4(ii) we alter the value of the com-pression parameter σ(α>0), whilst keeping the numberof configurations fixed at K = 150, and the number ofharmonic oscillator levels fixed at Ω = 25. All the calcu-lations superimpose, and there is even less discrepancythan in the λ0 = 0.001 case. For the remaining calcula-tions we choose a compression parameter of σ(α>0) = 107

as this demonstrates the best norm and particle numberconservation in panel (a) of Fig. 5(ii).

In panel (c) of Fig. 4(ii) we alter the value of K whilstkeeping σ(α>0) and Ω fixed. There are minor differencesbetween the K = 100 calculation and the K = 150 and

Page 12: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

12

(i)

0.9998

1

1.0002

1.0004

0 2 4 6 8 10

99.98

100

100.02

100.04(c)

Norm

Part

icle

Num

ber

Time (a.u.)

K=150, Ω=23, σ(α>0)

=109

K=150, Ω=24, σ(α>0)

=109

K=150, Ω=25, σ(α>0)

=109

K=150, Ω=26, σ(α>0)

=109

0.9994

0.9996

0.9998

1

1.0002

1.0004

99.94

99.96

99.98

100

100.02

100.04(b)

Norm

Part

icle

Num

ber

K=100, Ω=26, σ(α>0)

=109

K=150, Ω=26, σ(α>0)

=109

K=200, Ω=26, σ(α>0)

=109

0.9992

0.9994

0.9996

0.9998

1

1.0002

1.0004

1.0006

99.92

99.94

99.96

99.98

100

100.02

100.04

100.06(a)N

orm

Part

icle

Num

ber

K=150, Ω=26, σ(α>0)

=106

K=150, Ω=26, σ(α>0)

=107

K=150, Ω=26, σ(α>0)

=108

K=150, Ω=26, σ(α>0)

=109

(ii)

0.999

0.9995

1

1.0005

1.001

1.0015

0 2 4 6 8 10

99.9

99.95

100

100.05

100.1

100.15(c)

Norm

Part

icle

Num

ber

Time (a.u.)

K=150, Ω=20, σ(α>0)

=107

K=150, Ω=22, σ(α>0)

=107

K=150, Ω=24, σ(α>0)

=107

K=150, Ω=25, σ(α>0)

=107

0.999

0.9995

1

1.0005

1.001

99.9

99.95

100

100.05

100.1(b)

Norm

Part

icle

Num

ber

K=100, Ω=25, σ(α>0)

=107

K=150, Ω=25, σ(α>0)

=107

K=200, Ω=25, σ(α>0)

=107

0.9995

1

1.0005

1.001

1.0015

1.002

99.95

100

100.05

100.1

100.15

100.2(a)

Norm

Part

icle

Num

ber

K=150, Ω=25, σ(α>0)

=106

K=150, Ω=25, σ(α>0)

=107

K=150, Ω=25, σ(α>0)

=108

K=150, Ω=25, σ(α>0)

=109

FIG. 5: Norm (dashed/dotted lines without circles) and particle number (dashed/dotted lines with circles) forCCSB calculations of Application 2 with two-body interaction strength (i) λ0 = 0.001 and (ii) λ0 = 0.01 with

different values of (a) compression parameter for coherent state sampling of the harmonic oscillator levels with zeroinitial occupation σ(α>0), (b) configurations K and (c) harmonic oscillator levels in the basis Ω.

K = 200 calculations, the latter of which superimposeon the MCTDHB result. The norm and particle numberconservation of the K = 150 and K = 200 calculationsare very similar in panel (b) of Fig. 5(ii), therefore weregard the K = 150 result as being fully converged.

In panel (d) of Fig. 4(ii) we alter the value of Ω andkeep K and σ(α>0) fixed. As with the λ0 = 0.001 cal-culations, this has the largest effect on the breathing os-cillations, with Ω = 20 and Ω = 22 being insufficient todescribe them accurately, whilst the Ω = 24 and Ω = 25cases superimpose upon the MCTDHB result. The normand particle number conservation of the Ω = 24 result,shown in panel (c) of Fig. 5(ii) is not as good as theΩ = 25 result, which is why we choose the latter as ourmost accurate calculation.

The above demonstrates that CCSB is able to repro-duce MCTDHB calculations in both the mean field andmulti orbital fully quantum regimes, with similar levelsof theory for the CCSB calculations in each regime. Wehave also shown that the method converges appropriatelywith respect to the K and Ω parameters, this it is stablewith respect to norm and particle number conservation,and that appropriate choice of the compression parame-ter σ(α>0) for initial sampling of the coherent state basisis necessary, like in Application 1.

V. CONCLUSIONS

In this work the CCS method has been straight-forwardly applied to investigation of indistinguishablebosons, as MCTDH and ML-MCTDH have been, andthe method dubbed CCSB. Instead of the coherent statebasis functions being used to represent individual par-ticles like in the standard distinguishable representationof CCS, in CCSB they are used as a basis for numberoccupation of quantum states in the second quantisationFock state formalism.

Two example model Hamiltonians have been studied,demonstrating the accuracy of the method by compar-ing to fully quantum benchmarks. In the first example,CCSB was applied to the system-bath asymmetric doublewell tunnelling problem previously studied in Refs. [72–75, 78, 79] in distinguishable representation. As the bathis comprised of oscillators of the same frequency, theywere treated as indistinguishable and the bath portionof the Hamiltonian second quantised. The system tun-nelling portion of the Hamiltonian was kept in distin-guishable representation, therefore this first applicationwas a hybrid of standard CCS and CCSB. This does notpose a problem however, as the working equations for tra-jectories and time-dependence of amplitudes are the samein each. This may also be thought of as a system with abosonic bath and an impurity, opening up the possibil-

Page 13: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

13

ity of the method studying multi-atomic Bose-Einsteincondensates [64], spinor Bose-Einstein condensates [65],dark-bright solitons [66], and Bose-polarons [67]. Thepreviously studied 20D, quadratic system-bath couplingwith constant λ = 0.1 case [72–75, 78, 79] was investi-gated, and the second quantised bath required Ω = 5harmonic oscillator levels in the basis, thus the dimen-sionality of the problem was reduced from 20 to 6. TheCCSB calculation was in much better agreement with abenchmark result [79] on the system than all other meth-ods that have studied the problem.

In the second example, a model Hamiltonian for a sys-tem of 100 bosons in a shifted harmonic trap was stud-ied, the 1-body density has been calculated, as well asits variance, to demonstrate the breathing oscillations ofthe density. Matrix elements of 2-body operators hadto be calculated, as is common for interacting conden-sates, and these may be computed analytically by CCSB.The density oscillations were calculated at two differenttwo-body interaction strengths and compared to MCT-DHB benchmark calculations. The weaker interactionstrength was able to be described by MCTDHB with 1orbital, such that it was equivalent to the GPE mean-fieldtheory, whilst the stronger interaction strength requiredMCTDHB with 3 orbitals, and was thus fully quantum,taking correlations into account and going beyond themean-field approach. CCSB was able to reproduce bothresults with similar levels of theory, providing motivationfor further study on more challenging Bose-Einstein con-densate systems. In particular, future avenues of researchfor CCSB in this vein include more complicated Bose-Einstein condensate problems, such as that in Ref. [22]of a condensate in a single well trap that is deformed intoa double well, like that observed in experimental bosonicJosephson junctions [70, 71]. We also wish to considercondensates in multi-well traps, such as a multi-site ver-

sion of the Bose-Hubbard model studied in Ref. [61], andother interesting systems that have previously been stud-ied by ML-MCTDHB [45–52].

Both applications have demonstrated that the CCSBmethod converges with the number of configurations Kand number of quantum states included in the basis Ω.We have also demonstrated that appropriate sampling ofthe initial coherent states via a compression parameterσ is necessary to ensure a reliable and accurate calcula-tion. Further developments of the method that we envis-age in future include development of methods to generateinitial conditions, as imaginary time propagation is un-stable with trajectories; incorporation of SU(n) coherentstates, as demonstrated in Ref. [83]; and the combinationof the method with one to treat identical fermions [84] tostudy Bose-Fermi mixtures, as has been carried out byMCTDH [85] and ML-MCTDH [86] previously.

Data generated in this work that is shown in the figuresmay be found at https://doi.org/10.5518/595, alongwith the program code used to generate the data.

VI. ACKNOWLEDGEMENTS

J.A.G. has been supported by EPSRC grantEP/N007549/1, as well as the University Research Schol-arship from the University of Leeds, and funding from theSchool of Chemistry, University of Leeds. D.V.S acknowl-edges the support of the EPSRC, grant EP/P021123/1.J.A.G. would like to thank A. Streltsov for demonstra-tion of the use of the MCTDHB program, alongside help-ful discussions. J.A.G would also like to thank L. Chenfor reading through the manuscript and suggesting mod-ifications. J.A.G. and D.V.S gratefully acknowledge V.Batista, S. Habershon, and M. Saller for providing theirdata. This work was undertaken on ARC2 and ARC3,part of the High Performance Computing facilities at theUniversity of Leeds, UK.

Appendix A: Second Quantisation and Normal Ordering of Hamiltonian for Application 1

Using the definition of a second quantised Hamiltonian in Eq. 17 in the main text, Eq. 18 may be written as

H =p(m=1)2

2− q(m=1)2

2+q(m=1)4

16η+

Ω∑α,β=0

〈α| P2

2+

Q2

2|β〉 a(α)†a(β)

+λq(m=1)

2

Ω∑α,β=0

〈α|Q2|β〉 a(α)†a(β)

=p(m=1)2

2− q(m=1)2

2+q(m=1)4

16η+

[Ω∑α=0

〈α| P2

2+

Q2

2|α〉 a(α)†a(α)

]+λq(m=1)

2

Ω∑α,β=0

Q(α,β)2

a(α)†a(β)

=p(m=1)2

2− q(m=1)2

2+q(m=1)4

16η+

[Ω∑α=0

ε(α)a(α)†a(α)

]+λq(m=1)

2

Ω∑α,β=0

Q(α,β)2

a(α)†a(β)

.(A1)

The quantum states |α〉 and |β〉 are those of the harmonic oscillator with α and β numbers of quanta, and the equality

on the second line for 〈α| P2

2 + Q2

2 |β〉 follows because this is non-zero with eigenvalue ε(α) only when α = β. Thesums are from the ground level α = 0, to some upper level Ω. In principle, one should choose Ω =∞ for a completedescription of the bath, however in practice additional oscillator levels may simply be added on until a converged

Page 14: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

14

result is achieved. The position and momentum operators of the tunnelling mode have explicitly been labelled with(m = 1) to distinguish them from the α labelling scheme of the second quantised bath modes.

The matrix Q(α,β)2

is evaluated as

〈α|Q2|β〉 =

1

2

√(α+ 2)(α+ 1) if α = β − 2

1

2

√α(α− 1) if α = β + 2

ε(α) if α = β

0 otherwise.

(A2)

As this matrix is non-zero only for quanta α = β and α = β ± 2, and we may say that all bath modes are initiallyin the ground harmonic oscillator level (α = 0) as they are at the origin in distinguishable representation (previouslyassumed in the benchmark calculation [79]), only harmonic oscillator levels with even numbers of quanta will beincluded and the bottom line of Eq. A1 is written as

H =p(m=1)2

2− q(m=1)2

2+q(m=1)4

16η+

[Ω∑α=0

ε(2α)a(2α)†a(2α)

]+λq(m=1)

2

Ω∑α,β=0

Q(2α,2β)2

a(2α)†a(2β)

. (A3)

The relationship between the creation and annihilation operators and q and p given in Eq. 3 may then be used inEq. A3 alongside the relationships in Eqs. 2 and 6 to give Eq. 19.

Appendix B: Second Quantisation of Hamiltonian for Application 2

Using the definition of a second quantised Hamiltonian in Eq. 17 in the main text, Eq. 32 may be written as

H =

Ω∑α,β=0

〈α| P2

2+

Q2

2|β〉 a(α)†a(β) −

Ω∑α,β=0

〈α|ξQ|β〉 a(α)†a(β) +

Ω∑α,β=0

〈α|ξ2

2|β〉 a(α)†a(β)

+1

2

Ω∑α,β,γ,ζ=0

〈α, β|λ0δ(Q−Q′)|γ, ζ〉 a(α)†a(β)†a(ζ)a(γ)

=

Ω∑α=0

〈α| P2

2+

Q2

2|α〉 a(α)†a(α) −

Ω∑α,β=0

〈α|ξQ|β〉 a(α)†a(β) +

Ω∑α=0

〈α|ξ2

2|α〉 a(α)†a(α)

+1

2

Ω∑α,β,γ,ζ=0

〈α, β|λ0δ(Q−Q′)|γ, ζ〉 a(α)†a(β)†a(ζ)a(γ)

=

Ω∑α=0

ε(α)a(α)†a(α) −Ω∑

α,β=0

ξQ(α,β)a(α)†a(β) +

Ω∑α=0

ξ2

2a(α)†a(α) +

1

2

Ω∑α,β,γ,ζ=0

λ0δ(α,β,γ,ζ)a(α)†a(β)†a(ζ)a(γ).

(B1)

The relationships in Eqs. 2 and 6 may then be used with Eq. B1 to give Eq. 34. In Eq. B1, ε(α) is the eigenvalue ofthe harmonic oscillator for state |α〉, and Q(α,β) is a matrix given by

Q(α,β) = 〈α|Q|β〉 =

α2 α = β + 1√β2 β = α+ 1

0 otherwise.

. (B2)

Evaluation of the δ(α,β,γ,ζ) matrix is slightly more involved, as it is required to solve the integral

δ(α,β,γ,ζ) = 〈α, β|δ(Q−Q′)|γ, ζ〉

=

∫ +∞

−∞

∫ +∞

−∞

1√2αα!

(1

π

)1/4

e−Q2/2He(α)(Q)

1√2ββ!

(1

π

)1/4

e−Q′2/2He(β)(Q′)

× δ(Q−Q′)1√2γγ!

(1

π

)1/4

e−Q2/2He(γ)(Q)

1√2ζζ!

(1

π

)1/4

e−Q′2/2He(ζ)(Q′) dQdQ′

(B3)

Page 15: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

15

where He(α)(Q) is a Hermite polynomial of order α. However, an analytic solution is possible, and the above may besimplified using the relationship ∫ +∞

−∞f(x′)δ(x− x′) dx′ = f(x), (B4)

and like terms collated to obtain

δ(α,β,γ,ζ) =1

π√

2(α+β+γ+ζ)α!β!γ!ζ!

∫ +∞

−∞e−2Q2

He(α)(Q)He(β)(Q)He(γ)(Q)He(ζ)(Q) dQ. (B5)

This will only be non-zero if the integrand is an even function, so the product of Hermite polynomials can only haveeven powers of Q

δ(α,β,γ,ζ) =1

π√

2α+β+γ+ζα!β!γ!ζ!

∫ +∞

−∞e−2Q2

(α+β+γ+ζ)/2∑τ=0

c2τQ2τ dQ (B6)

where c2τ is a constant obtained from the product of Hermite polynomial coefficients. Using the following identity∫ +∞

−∞x2ne−

12ax

2

dx =

√2π

a

1

an(2n− 1)!! forn > 0 (B7)

combined with a Gaussian integral for τ = 0, Eq. B6 can be evaluated as

δ(α,β,γ,ζ) =1

π√

2α+β+γ+ζα!β!γ!ζ!

√π

2c0 +

(α+β+γ+ζ)/2∑τ=1

c2τ

√π

2

1

4τ(2τ − 1)!!

. (B8)

We note that an alternative method of calculating the δ(α,β,γ,ζ) matrix elements exists using Gauss-Hermite quadra-ture [87], however our approach was sufficiently efficient.

[1] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E.Wieman, and E. A. Cornell, Science 269, 198 (1995).

[2] C. C. Bradley, C. A. Sackett, J. J. Tollett, and R. G.Hulet, Physical Review Letters 75, 1687 (1995).

[3] K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. vanDruten, D. S. Durfee, D. M. Kurn, and W. Ketterle,Physical Review Letters 75, 3969 (1995).

[4] C. Orzel, A. K. Tuchman, M. L. Fenselau, M. Yasuda,and M. A. Kasevich, Science 291, 2386 (2001).

[5] M. Albiez, R. Gati, J. Folling, S. Hunsmann, M. Cris-tiani, and M. K. Oberthaler, Physical Review Letters95, 010402 (2005).

[6] S. Levy, E. Lahoud, I. Shomroni, and J. Steinhauer,Nature 449, 579 (2007).

[7] M. R. Matthews, B. P. Anderson, P. C. Haljan, D. S.Hall, C. E. Wieman, and E. A. Cornell, Physical ReviewLetters 83, 2498 (1999).

[8] K. W. Madison, F. Chevy, W. Wohlleben, and J. Dal-ibard, Physical Review Letters 84, 806 (2000).

[9] S. Burger, K. Bongs, S. Dettmer, W. Ertmer, K. Seng-stock, A. Sanpera, G. V. Shlyapnikov, and M. Lewen-stein, Physical Review Letters 83, 5198 (1999).

[10] J. Denschlag, J. E. Simsarian, D. L. Feder, C. W. Clark,L. A. Collins, J. Cubizolles, L. Deng, E. W. Hagley,

K. Helmerson, W. P. Reinhardt, S. L. Rolston, B. I.Schneider, and W. D. Phillips, Science 287, 97 (2000).

[11] E. P. Gross, Nuovo Cimento 20, 454 (1961).[12] L. P. Pitaevskii, Soviet Physics JETP-USSR 13, 451

(1961).[13] A. Smerzi, S. Fantoni, S. Giovanazzi, and S. R. Shenoy,

Physical Review Letters 79, 4950 (1997).[14] V. M. Perez-Garcıa, H. Michinel, J. I. Cirac, M. Lewen-

stein, and P. Zoller, Physical Review A 56, 1424 (1997).[15] S. Raghavan, A. Smerzi, S. Fantoni, and S. R. Shenoy,

Physical Review A 59, 620 (1999).[16] W. Bao, D. Jaksch, and P. A. Markowich, Journal of

Computational Physics 187, 318 (2003).[17] Z. X. Liang, Z. D. Zhang, and W. M. Liu, Physical

Review Letters 94, 050402 (2005).[18] D. Ananikian and T. Bergeman, Physical Review A 73,

013604 (2006).[19] A. J. Leggett, Reviews of Modern Physics 73, 307 (2001).[20] A. Minguzzi, S. Succi, F. Toschi, M. Tosi, and P. Vignolo,

Physics Reports 395, 223 (2004).[21] A. I. Streltsov, O. E. Alon, and L. S. Cederbaum, Phys-

ical Review Letters 99, 030402 (2007).[22] O. E. Alon, A. I. Streltsov, and L. S. Cederbaum, Phys-

ical Review A 77, 033613 (2008).

Page 16: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

16

[23] A. I. Streltsov, O. E. Alon, and L. S. Cederbaum, Phys-ical Review Letters 100, 130401 (2008).

[24] A. I. Streltsov, O. E. Alon, and L. S. Cederbaum, Phys-ical Review A 80, 043616 (2009).

[25] A. I. Streltsov, O. E. Alon, and L. S. Cederbaum, Journalof Physics B 42, 091004 (2009).

[26] A. U. J. Lode, A. I. Streltsov, O. E. Alon, H.-D. Meyer,and L. S. Cederbaum, Journal of Physics B 42, 044018(2009).

[27] K. Sakmann, A. I. Streltsov, O. E. Alon, and L. S. Ceder-baum, Physical Review Letters 103, 220601 (2009).

[28] K. Sakmann, A. I. Streltsov, O. E. Alon, and L. S. Ceder-baum, Physical Review A 82, 013620 (2010).

[29] A. I. Streltsov, O. E. Alon, and L. S. Cederbaum, Phys-ical Review Letters 106, 240401 (2011).

[30] A. I. Streltsov, K. Sakmann, O. E. Alon, and L. S. Ceder-baum, Physical Review A 83, 043604 (2011).

[31] A. U. J. Lode, K. Sakmann, O. E. Alon, L. S. Cederbaum,and A. I. Streltsov, Physical Review A 86, 063606 (2012).

[32] A. I. Streltsov, Physical Review A 88, 041602(R) (2013).[33] R. Beinke, S. Klaiman, L. S. Cederbaum, A. I. Streltsov,

and O. E. Alon, Physical Review A 92, 043627 (2015).[34] L. Cao, S. Kronke, O. Vendrell, and P. Schmelcher, Jour-

nal of Chemical Physics 139, 134103 (2013).[35] S. Kronke, L. Cao, O. Vendrell, and P. Schmelcher, New

Journal of Physics 15, 063018 (2013).[36] J. M. Schurer, A. Negretti, and P. Schmelcher, Physical

Review Letters 119, 063001 (2017).[37] K. Keiler and P. Schmelcher, New Journal of Physics 20,

103042 (2018).[38] S. I. Mistakidis, A. G. Volosniev, N. T. Zinner, and

P. Schmelcher, ArXiv e-prints (2018), arXiv:1809.01889.[39] S. Mistakidis, G. Katsimiga, G. Koutentakis,

T. Busch, and P. Schmelcher, ArXiv e-prints (2018),arXiv:1811.10702.

[40] G. C. Katsimiga, S. I. Mistakidis, G. M. Koutentakis,P. G. Kevrekidis, and P. Schmelcher, Physical Review A98, 013632 (2018).

[41] S. I. Mistakidis, G. C. Katsimiga, P. G. Kevrekidis, andP. Schmelcher, New Journal of Physics 20, 043052 (2018).

[42] G. C. Katsimiga, G. M. Koutentakis, S. I. Mistakidis,P. G. Kevrekidis, and P. Schmelcher, New Journal ofPhysics 19, 073004 (2017).

[43] S. Kronke and P. Schmelcher, Physical Review A 91,053614 (2015).

[44] G. C. Katsimiga, S. I. Mistakidis, G. M. Koutentakis,P. G. Kevrekidis, and P. Schmelcher, New Journal ofPhysics 19, 123012 (2017).

[45] S. I. Mistakidis, L. Cao, and P. Schmelcher, Journal ofPhysics B: Atomic, Molecular and Optical Physics 47,225303 (2014).

[46] S. I. Mistakidis, L. Cao, and P. Schmelcher, PhysicalReview A 91, 033611 (2015).

[47] S. I. Mistakidis, T. Wulf, A. Negretti, and P. Schmelcher,Journal of Physics B: Atomic, Molecular and OpticalPhysics 48, 244004 (2015).

[48] S. I. Mistakidis and P. Schmelcher, Physical Review A95, 013625 (2017).

[49] G. M. Koutentakis, S. I. Mistakidis, and P. Schmelcher,Physical Review A 95, 013617 (2017).

[50] J. Neuhaus-Steinmetz, S. I. Mistakidis, andP. Schmelcher, Physical Review A 95, 053610 (2017).

[51] S. Mistakidis, G. Koutentakis, and P. Schmelcher,Chemical Physics 509, 106 (2018).

[52] T. Plaßmann, S. I. Mistakidis, and P. Schmelcher, Jour-nal of Physics B: Atomic, Molecular and Optical Physics51, 225001 (2018).

[53] H.-D. Meyer, U. Manthe, and L. Cederbaum, ChemicalPhysics Letters 165, 73 (1990).

[54] H. Wang and M. Thoss, Journal of Chemical Physics 119,1289 (2003).

[55] U. Manthe, Journal of Chemical Physics 128, 164116(2008).

[56] D. V. Shalashilin and M. S. Child, Journal of ChemicalPhysics 113, 10028 (2000).

[57] D. V. Shalashilin and M. S. Child, Chemical Physics 304,103 (2004).

[58] D. V. Shalashilin and M. S. Child, Journal of ChemicalPhysics 119, 1961 (2003).

[59] D. V. Shalashilin and M. S. Child, Journal of ChemicalPhysics 128, 054102 (2008).

[60] L. Simon and W. T. Strunz, Physical Review A 89,052112 (2014).

[61] S. Ray, P. Ostmann, L. Simon, F. Grossmann, and W. T.Strunz, Journal of Physics A: Mathematical and Theo-retical 49, 165303 (2016).

[62] M. S. Child and D. V. Shalashilin, Journal of ChemicalPhysics 118, 2061 (2003).

[63] D. V. Shalashilin and M. S. Child, Journal of ChemicalPhysics 121, 3563 (2004).

[64] G. Modugno, M. Modugno, F. Riboli, G. Roati, andM. Inguscio, Physical Review Letters 89, 190404 (2002).

[65] Y. Kawaguchi and M. Ueda, Physics Reports 520, 253(2012).

[66] C. Becker, S. Stellmer, P. Soltan-Panahi, S. Dorscher,M. Baumert, E.-M. Richter, J. Kronjager, K. Bongs, andK. Sengstock, Nature Physics 4, 496 (2008).

[67] M. Bruderer, A. Klein, S. R. Clark, and D. Jaksch, Phys-ical Review A 76, 011605(R) (2007).

[68] M. Greiner, O. Mandel, T. Esslinger, T. W. Hansch, andI. Bloch, Nature 415, 39 (2002).

[69] D. Jaksch and P. Zoller, Annals of Physics 315, 52(2005).

[70] R. Gati, M. Albiez, J. Folling, B. Hemmerling, andM. Oberthaler, Applied Physics B 82, 207 (2006).

[71] R. Gati and M. K. Oberthaler, Journal of Physics B:Atomic, Molecular and Optical Physics 40, R61 (2007).

[72] Y. Wu and V. S. Batista, Journal of Chemical Physics121, 1676 (2004).

[73] P. A. J. Sherratt, D. V. Shalashilin, and M. S. Child,Chemical Physics 322, 127 (2006).

[74] S. Habershon, Journal of Chemical Physics 136, 054109(2012).

[75] M. A. C. Saller and S. Habershon, Journal of ChemicalTheory and Computation 13, 3085 (2017).

[76] J. P. Alborzpour, D. P. Tew, and S. Habershon, Journalof Chemical Physics 145, 174112 (2016).

[77] T. Murakami and T. J. Frankcombe, Journal of ChemicalPhysics 149, 134113 (2018).

[78] J. A. Green, A. Grigolo, M. Ronto, and D. V. Shalashilin,Journal of Chemical Physics 144, 024111 (2016).

[79] J. A. Green and D. V. Shalashilin, Chemical Physics Let-ters 641, 173 (2015).

[80] A. I. e. a. Streltsov, “The multiconfigurationaltime-dependent hartree for bosons package,”Http://mctdhb.org.

[81] D. V. Shalashilin, M. S. Child, and D. C. Clary, Journalof Chemical Physics 120, 5608 (2004).

Page 17: Coherent States Method - arXiv · 2019. 7. 1. · Simulation of the Quantum Dynamics of Indistinguishable Bosons with the Coupled Coherent States Method James A. Green1,2, and Dmitrii

17

[82] K. Sakmann, A. I. Streltsov, O. E. Alon, and L. S. Ceder-baum, Physical Review A 89, 023602 (2014).

[83] A. Grigolo, T. F. Viscondi, and M. A. M. de Aguiar,Journal of Chemical Physics 144, 094106 (2016).

[84] D. V. Shalashilin, Journal of Chemical Physics 148,194109 (2018).

[85] O. E. Alon, A. I. Streltsov, and L. S. Cederbaum, Journal

of Chemical Physics 127, 154103 (2007).[86] L. Cao, V. Bolsinger, S. I. Mistakidis, G. M. Koutentakis,

S. Krnke, J. M. Schurer, and P. Schmelcher, Journal ofChemical Physics 147, 044106 (2017).

[87] M. Edwards, R. J. Dodd, C. W. Clark, and K. Burnett,Journal of research of the National Institute of Standardsand Technology 101, 553 (1996).