• Keine Ergebnisse gefunden

“Coherent Backscattering in Fock Space:

N/A
N/A
Protected

Academic year: 2022

Aktie "“Coherent Backscattering in Fock Space:"

Copied!
3
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Supplementary material for

“Coherent Backscattering in Fock Space:

a Signature of Quantum Many-Body Interference in Interacting Bosonic Systems”

Thomas Engl,1 Julien Dujardin,2 Arturo Arg¨uelles,2 Peter Schlagheck,2Klaus Richter,1and Juan Diego Urbina1

1Institut f¨ur Theoretische Physik, Universit¨at Regensburg, D-93040 Regensburg, Germany

2D´epartement de Physique, Universit´e de Li`ege, 4000 Li`ege, Belgium

The semiclassical propagator cannot be directly ob- tained in Fock state representation, since the Fock states form a discrete basis rather than a continuous one as re- quired by the path integral formalism. To solve this prob- lem, we first derive a semiclassical propagator in quadra- ture representation and then project the result on Fock states.

The quadrature eigenstates |qi ≡ |q1, . . . , qLi and

|pi ≡ |p1, . . . , pLiare defined as the eigenstates of linear hermitian combinations of the creation and annihilation operators associated with the single-particle orbitals χα

(α= 1, . . . , L),i.e.,

b ˆaα+ ˆaα

|qi=qα|qi (1)

−ib ˆaα−aˆα

|pi=pα|pi, (2) with an arbitrary but fixed scale b. These quadrature eigenstates obey the resolutions of unity

Z

dLq|qihq|= ˆ1 = Z

dLp|pihp| (3)

and their overlap matrix elements are given by

hq|qi=δ(q−q) , (4) hp|pi=δ(p−p) , (5)

hq|pi= 1 2b√

πexp ip·q

2b2

. (6)

Following the usual steps to derive the Feynman prop- agator (i.e., splitting up the exponential into a product ofN exponentials, inserting unity operators in terms of ˆ

q and ˆp between each pair of exponentials, and taking the limitN → ∞), we find for the Hamiltonian

Hˆ =X

αβ

hαβαˆaβ+1 2

X

αβµσ

Vαβµσˆaαβµˆaσ (7)

[see Eq. (2) in the Letter] a path integral representation of the propagator in quadrature representation as

K

q(f),q(i), t

= lim

N→∞

Z

dq(1)· · · Z

dq(N−1)

Z dp(1) 4πb2 · · ·

Z dp(N) 4πb2

N

Y

k=1

exp

"

i

2b2p(k)·

q(k)−q(k−1)

−iτ

~

L

X

α,β=1

hαβ(k)α )ψ(k)β − iτ 2~

L

X

α,βµ,σ=1

Vαβµσ(k)α )µ(k))ψ(k)β ψ(k)σ

#

, (8)

with 2bψ(k)≡q(k−1)+ ip(k), 2b(ψ(k))≡q(k−1)−ip(k), and τ ≡ t/N. Calculating the integrals in stationary phase approximation and finally taking the limitN → ∞ yields the van-Vleck-Gutzwiller propagator

K

q(f),q(i), t

= 1

(−2πi~)L/2

×X

γ

s

det ∂2Rγ

∂q(i)∂q(f)eiRγ/~ (9) where the sum runs over all possible classical trajectories

defined by the equation of motion

i~∂ψα(s)

∂s =

L

X

β=1

hαβψβ(s)

+

L

X

β,µ,σ=1

Vαβµσψµ(s)ψβ(s)ψσ(s) (10)

and the boundary conditions 2bRe[ψ(0)] = q(i) and 2bRe[ψ(t)] =q(f). Rγ is the classical action given by

(2)

2

Rγ

q(f),q(i), t

=

t

Z

0

ds ( ~

2b2p(s)·q(s)˙ −

L

X

α,β=1

hαβψα(s)ψβ(s)−1 2

L

X

α,β,µ,σ=1

Vαβµσψα(s)ψβ(s)ψµ(s)ψσ(s) )

(11)

withq(s)≡2bRe[ψ(s)] andp(s)≡2bIm[ψ(s)] evaluated along the trajectoryγ.

We now want to project the result onto Fock states.

To this end, we need the overlap hn|qi= 1

p2nn!√

2πbexp

−q2 4b2

Hn

q

√2b

(12) of a quadrature eigenstate|qi with a Fock state|nias- sociated with an individual single-particle orbital. For largen, we can employ the WKB approximation for the Hermite polynomialsHn, which yields

hn|qi ≃

p2/π (4b2n−q2)1/4

×cosh q 4b2

p4b2(n+ 1/2)−q2i

(13) within the oscillatory region |q| ≤ 2b√

n. Since hn|qi decreases exponentially for larger values of |q|, we can restrict the integration necessary for the projection onto Fock states to the oscillatory region|q| ≤2b√

n.

With this restriction, we substituteqα(i/f)7→θ(i/f)α (α∈ {1, . . . , L}) with θ(i/f)α ∈[−π, π] defined through

q(i)1 ≡2b q

n(i)1 + 1/2 cos θ(i)1

, (14)

q(f)1 ≡2b q

n(f)1 + 1/2 cos

θ(f)11(i)

, (15)

as well as qα(i/f)≡2b

q

n(i/f)α + 1/2 cos

θα(i/f)(i)1

(16) for α = 2, . . . , L. The integrations over θ(f)1 , . . . , θL(f), θ(i)2 , . . . , θ(i)L can now be performed in stationary phase approximation. This selects trajecto- ries that satisfy

ψ1(t) = q

n(f)1 + 1/2 expn i

θ1(f)1(i)o

, (17)

as well as ψα(0) =

q

n(i)α + 1/2 expn i

θα(f)(i)1 o

(18) ψα(t) =

q

n(f)α + 1/2 expn i

θα(f)(i)1 o

(19) forα= 2, . . . , L. Since the classical equations of motion preserve|ψ1(t)|2+. . .+|ψL(t)|2, these stationary phase conditions already imply |ψ1(0)|2 =n(i)1 + 1/2 provided the final state

n(f)i contains as many particles as the

initial state

n(i)i. Due to theU(1) gauge symmetry, a variation of the global phase θ(i)1 will trivially give rise to another solution ψ = ψexp(θ1(i)) that exhibits the same initial and final populations as ψ. Therefore, one cannot solve the integral over θ(i)1 in stationary phase approximation, but needs to do it exactly, which yields an additional factor 2π.

The propagator in Fock space then finally reads K

n(f),n(i), t

= 1

(−2πi~)(L−1)/2

×X

γ

r

det2Rγ

∂n(i)∂n(f)eiRγ/~, (20) whereγ indexes all classical trajectories that satisfy the boundary conditions|ψα(0)|2=n(i)α + 1/2 and|ψα(t)|2= n(f)α + 1/2 for α = 1, . . . , L with the global phase θ1(i) being fixed through the choice ψ1(0) = [n(i)1 + 1/2]1/2. The prime at the determinant

det2Rγ

∂n(i)∂n(f) ≡det ∂2Rγ

∂n(i)α∂n(f)β

!

α,β=2,...,L

(21) indicates that it is taken with respect to the matrix ob- tained by skipping the derivatives with respect to the first components.

The transition probability between Fock states at time tis then given by

P(n(f),n(i), t) =X

γγ

A(γ)A)ei(R(γ)−R)) (22)

with

A)(n(f),n(i), t) = s

det 1 2π

2R(γ)(n(f),n(i), t)

∂n(f)∂n(i)

. (23) A disorder average then suppresses quasi-random contri- butions from pairs of longer, unrelated pathsγ6=γ in the double sum in Eq. (22) and promotes contributions from those pairs of classical solutions that exhibit action de- generacies: R(γ)=R). Forn(f)6=n(i), the dominating contribution to the average transition probability is given by the incoherent (γ=γ) part of the double sum,

cl(n(f),n(i), t) =X

γ

|A(γ)|2. (24) This result is also expected for the case of backscattering, n(f)=n(i), if the dynamics of the system is not invariant under time reversal.

(3)

3

FIG. 1. (Color online) Sketch of a self-retracing (left column) and a non-self-retracing (right column) backscattered trajec- tiy in population space, for a Bose-Hubbard chain of five sites (marked by open black circles) containing altogether six par- ticles (marked by filled blue circles). The curved (red) ar- rows indicate the imminent hopping event along the chain, while the vertical (black) arrows indicate the time evolution.

The self-retracing trajectory shown in the left column involves two hopping events, with one particle hopping to an adjacent site and back. A non-self-retracing backscattered trajectory within the chain, on the other hand, requires at least four hopping events. In the example shown on the right column, two atoms undergo hoppings to adjacent sites and back in an alternating order.

In the presence of time-reversal invariance, the backscattering probability P(n(i),n(i), t) contains addi- tional contributions from pairs of backscattered trajec- tories γ, γ where γ = Tγ is the time reversal of γ.

Obviously, this requires that γ 6= Tγ, i.e., that γ is a non-self-retracing trajectory. In the limit of long evo- lution times t, self-retracing trajectories γ =Tγ gener- ally represent a negligible minority among the set of all backscattered trajectories. Then we can safely assume that all backscattered trajectories arenon-self-retracing, which yields the semiclassical prediction

P¯(n(i),n(i), t) = 2 ¯Pcl(n(i),n(i), t) = 2X

γ

|A(γ)|2 (25) for the average quantum backscattering probability. The right column of Fig. 1 shows an example of such a non- self-retracing backscattered trajectory within the on-site population space, for the specific case of a Bose-Hubbard chain containing five sites. It involves four inter-site hop- pings of particles along the chain.

For very short evolution times, on the contrary, the average quantum backscattering probability within such

a Bose-Hubbard chain is obviously dominated by con- tributions fromself-retracing trajectories, namely those trajectories in which one particle hops to an adjacent site and back, as shown in the left column of Fig. 1. Hence, for very short timest we obtain

P¯(n(i),n(i), t) = ¯Pcl(n(i),n(i), t) =X

γ

|A(γ)|2, (26) i.e., the classical and quantum backscattering probabili- ties are expected to agree with each other.

The crossover between the short-time behaviour given by Eq. (26) and the long-time behaviour given by Eq. (25) can generally be expressed as

P¯(n(i),n(i), t) = [1 +η(t)] ¯Pcl(n(i),n(i), t) (27) whereη(t) represents the relative frequency of non-self- retracing trajectories among the set of all backscattered trajectories at the evolution time t. This relative fre- quency depends on system-specific parameters, such as the relative interaction and disorder strengthsU/J and W/J, respectively, in our disordered Bose-Hubbard sys- tem. It can be quantitatively determined by means of the computation of classical trajectories within the Bose- Hubbard system under consideration, if a quantitative prediction of ¯P(n(i),n(i), t) is required in this crossover regime.

To provide an estimate for the relevant time scale that characterizes this crossover in the case of our Bose- Hubbard chain or ring systems, it is useful to note that at least four inter-site hoppings are required in order to accomplish a backscattered classical trajectory that is not identical to its time-reversed counterpart. In- deed, backscattered trajectories involving only two inter- site hoppings are necessarily self-retracing in population space, as the first hopping event has to be undone by the second hopping event in order to recover the ini- tial distribution of populations, and a trajectory with exactly three inter-site hoppings cannot recover the ini- tial distribution of populations at the end (except for a the special case of a three-site Bose-Hubbard ring) as those hoppings arise between adjacent sites only. De- noting by τ = ~/J the inverse Rabi frequency charac- terizing inter-site hoppings within the (non-interacting and disorder-free) Bose-Hubbard system, we can there- fore roughly estimate this crossover time scale as 4τ,i.e., Eqs. (26) and (25) are predicted to be valid fort ≪4τ andt≫4τ, respectively. This prediction is found to be in good agreement with our numerical findings, as shown in Fig. 2 of the Letter.

Abbildung

FIG. 1. (Color online) Sketch of a self-retracing (left column) and a non-self-retracing (right column) backscattered  trajec-tiy in population space, for a Bose-Hubbard chain of five sites (marked by open black circles) containing altogether six  pticles

Referenzen

ÄHNLICHE DOKUMENTE

enable two Soyuz spacecraft to rendezvous, and a docking system locks the two spacecraft together. A separate living compartment in each Soyuz is used for

Women interviewed for this report in Egypt, Libya, and Yemen felt increasingly concerned about their general safety and security, highlighting fears of crime and armed

Key words: source, sink, wise exploitation, brute force, education, emotions, fix cost, variable cost, productivity, game theory, cooperation, prisoners´ dilemma,

Key words: source, sink, wise exploitation, brute force, fear, education, emotions, hope, hate, fix cost, variable cost, productivity, game theory,

Key words: source, sink, wise exploitation, brute force, education, emotions, hope, hate, fix cost, variable cost, productivity, game theory, cooperation,

Key words: source, sink, ensemble, group selection, selfishness, wise exploitation, brute force, fear, education, emotions, hope, hate, fix cost, variable cost, productivity,

source, sink, ensemble, group selection, selfishness, altruism, exploitation, wise exploitation, brute force, fear, education, emotions, hope, hate, fix cost,

When the second moment is large at small spatial displacements (spatial aggregation), the negative inte- gral term in Equation (4) becomes greater, slowing down population growth,