Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Impact of the initial fluctuations on the dissipative dynamics of interacting Fermi systems: a model

case study
Ibrahim Ulgen,1, 2 Bulent Yilmaz,1, ∗ and Denis Lacroix3
1 Physics Department, Faculty of Sciences, Ankara University, 06100 Ankara, Turkey
2 Physics Department, Faculty of Sciences and Arts, Siirt University, 56100 Siirt, Turkey
3 Institut de Physique Nucléaire, IN2P3-CNRS, Université Paris-Sud,
Université Paris-Saclay, F-91406 Orsay Cedex, France
(Dated: August 16, 2019)
Standard methods used for computing the dynamics of a quantum many-body system are the mean-field
(MF) approximations such as the time-dependent Hartree-Fock (TDHF) approach. Even though MF approaches
are quite successful, they suffer some well–known shortcomings one of which is insufficient dissipation of
collective motion. The stochastic mean-field approach (SMF), where a set of MF trajectories with random
initial conditions are considered, is a good candidate to include dissipative effects beyond mean-field. In this
arXiv:1908.05520v1 [nucl-th] 15 Aug 2019

approach, the one-body density matrix elements are treated initially as a set of stochastic Gaussian c-numbers
that are adjusted to reproduce first and second moments of collective one-body observables. It is shown that
the predictive power of the SMF approach can be further improved by relaxing the Gaussian assumption for the
initial probabilities. More precisely, using Gaussian or uniform distributions for the matrix elements generally
lead to overdamping for long times whereas distributions with smaller kurtosis lead to much better reproduction
of the long time evolution.

I. INTRODUCTION The SMF approach and the f-TWA approach have also in com-
mon that they both assume Gaussian probabilities for the ini-
In many situations, the evolution of a quantum system can tial fluctuations. However, this assumption turns out to be
be replaced by a set of classical evolutions with random ini- more guided by practical arguments than from first principles.
tial conditions optimized to reproduce at best the initial quan- In the present work, we further explore the possibility to use
tum zero-point motion [1]. This quantum to classical map- alternative probability distribution for the initial conditions in
ping is particularly suitable in the absence of interference or the SMF approach. We show that, while the second moment
tunneling and can be exact in some cases, like the free wave of a one-body observable can be generally interpreted clas-
expansion or the quantum harmonic oscillator. This idea is sically, the fourth moment of a one-body observable is more
employed in many fields of physics to describe the out-of- problematic, since it can lead to negative values for the fourth
equilibrium motion of complex systems, like in quantum op- moments of the stochastic matrix elements of one-body den-
tics [2]. This also includes the possibility to describe many- sities when considering many-body Fermi systems. Such be-
body interacting systems. An illustration in bosonic systems is havior cannot be reproduced by a classical mapping. How-
the truncated Wigner approximation (TWA) [3] (see also [4]). ever, probability distributions with smaller kurtosis compared
For Fermi systems, as underlined in ref. [5], such mapping is to the Gaussian distribution turn out to be more efficient to de-
more tricky due to the absence of natural classical representa- scribe the time evolution in the SMF approach. Such finding
tion, contrary to bosonic systems. Despite this inherent diffi- is illustrated on a modified version of the Lipkin-Meshkov-
culty for fermionic systems, one can mention two attempts to Glick model [20].
map the complex many-body problem of interacting systems The paper is organized as follows. In section II, the third
as a set of "classical trajectories". The first one is the stochas- and fourth moments of the matrix elements of the stochastic
tic mean-field (SMF) approach [6,7] where the stochastic one- one-body densities are derived within the SMF approach and
body density matrices are treated as classical objects evolving it is shown that initial probability distribution functions with
through the time-dependent mean-field (TDMF) equation of small kurtosis can provide a better approximation to the fourth
motion. A second formulation was made more recently in central moment. In section III, the SMF dynamics is applied
ref. [5] based on the mapping between fermionic and bosonic to a modified version of the Lipkin-Meshkov-Glick model. Fi-
operators, leading to the fermionic-TWA (f-TWA). We actu- nally, the conclusions are given in section IV.
ally realize recently that the equation of motion obtained in the
f-TWA often coincides with the TDMF evolution and there-
fore these two approaches are relatively close to each other.
II. STOCHASTIC MEAN-FIELD APPROACH
In the last decades, the SMF approach has been successfully
applied to model cases [8–10] as well as realistic simulations
of dynamical phenomena in nuclear physics [11–18]. It has The Schrödinger equation of a many-body fermionic sys-
also been extended to describe superfluid systems in ref. [19]. tem can be cast into a hierarchy of differential equations
for reduced densities (one-body, two-body, etc.) known as
BBGKY (Bogoliubov-Born-Green-Kirkwood-Yvon) hierar-
chy. The truncation of BBGKY equations at lowest order
∗ bulent.yilmaz@science.ankara.edu.tr gives the mean-field equation for the dynamics of one-body
2

density, and

ih̄ ρ = [h[ρ], ρ], (1) 1
hA2 (0)i − hA(0)i2 = [ni (1 − n j ) + n j (1 − ni )] A ji Ai j ,
2∑
∂t
ij
where h[ρ] is the mean-field Hamiltonian. In the mean-field (8)
approximation, one-body density contains all the information
on the system hence the many-body state is restricted to be
a Slater determinant during the dynamical evolution. MF dy- respectively. Comparing Eqs. (4) and (6) with Eqs. (7) and (8),
namics (or TDHF) is known to provide good approximation to we see that the quantum averages match the classical averages
one-body observables but severely underestimate their quan- if we have,
tum fluctuations. Beyond mean-field approaches are neces-
sary to overcome these shortcomings and provide a more ac- ρiλj (0) =ni δi j , (9)
curate description of the dynamical evolution.
In the SMF approach [6], the initial one-body density ρ(0) 1
δ ρiλj (0)δ ρklλ (0) =δil δ jk [ni (1 − n j ) + n j (1 − ni )] . (10)
is replaced by an ensemble of stochastic initial one-body den- 2
sities ρ λ (0) where λ stands for event label. Each of these
densities evolves with its self-consistent mean-field equation, In the original formulation of the SMF approach [6], the
stochastic matrix elements δ ρiλj (0) are assumed to be Gaus-
∂ λ
ih̄ ρ = [h[ρ λ ], ρ λ ]. (2) sian random numbers satisfying Eqs. (9) and (10).
∂t
For each event λ , the ”event“ expectation value of a one-body
observable A is given by
A. Higher order moments of the stochastic matrix elements of
hAiλ = Tr(ρ λ A). (3) the one-body density
In the SMF approach, the expectation values are obtained by
ensemble averages. Hence, the expectation value of the one- Here, we derive higher order moments for the stochastic
body observable is defined as, matrix elements and test the Gaussian assumption of the orig-
inal formulation of the SMF approach when the initial state is
hAiλ =Tr(ρ λ A) a Slater determinant. For this purpose, similarly to the second
= ∑ ρiλj A ji , (4) central moment Eq. (6), we define the expectation value of the
ij mth central moment of a one-body operator A as

where overline stands for ensemble averaging explicitly given  m


(m)
by σA = hAiλ − hAiλ
N
 m
1 = Tr(δ ρ λ A) , (11)
ρiλj = lim
N →∞ N
∑ ρiλj . (5)
λ =1

Here, N is the number of events in the ensemble. The quan- where δ ρ λ = ρ λ − ρ λ . Hence, the third and fourth central
tum variance of the one-body operator is defined by moments can be written as,
 2 (3)
(2)
σA = hAiλ − hAiλ σA = ∑ δ ρiλj δ ρklλ δ ρmn
λ A A A ,
ji lk nm (12)
i jklmn
 2
= Tr(δ ρ λ A) (4)
σA = δ ρiλj δ ρklλ δ ρmn
λ δ ρλ A A A A . (13)
∑ pr ji lk nm rp
i jklmnpr
=∑ δ ρiλj δ ρklλ A ji Alk , (6)
i jkl
The corresponding quantum central moments, for an initial
where δ ρ λ = ρ λ − ρ λ . Slater determinant, are given by (see Appendix A for details),
In the SMF approach, the distribution of the stochastic ma-
trix elements, ρiλj (0), is chosen such that the initial mean and (3)
variances of the observables are the same with those obtained h(∆A)3 i = ∑ Λi jk Ai j A jk Aki , (14)
i jk
with the initial density ρ(0). If {|ii} stand for the natural basis
which are satisfying hi|ρ(0)| ji = ni δi j , the mean and variance (4a)
h(∆A)4 i = ∑ Λi jkl Ai j A jk Akl Ali
of A are the expectation values given by i jkl
(4b)
hA(0)i = ∑ ni Aii , (7) + 3 ∑ Λi jkl Ai j A ji Akl Alk , (15)
i i jkl
3

where ∆A = A − hAi and Λ terms can assume the following non-zero values,

(3) (3) (3) 1


(3) 1 Λ pph = Λ php = Λhpp = − ,
Λi jk = [ni (1 − 3n j )(1 − n j nk ) + nk (1 − 3ni )(1 − ni n j ) 3
3 1
(3) (3) (3)
+n j (1 − 3nk )(1 − nk ni )] , (16) Λ phh = Λhph = Λhhp = + ,
3
(4a) 1 (4a) (4a) (4a) (4a) 1
Λi jkl = [ni (1 − 4n j )(1 − 3nk )(1 − n j nk nl ) Λ ppph = Λ pphp = Λ phpp = Λhppp = + ,
4 4
+ nl (1 − 4ni )(1 − 3n j )(1 − ni n j nk ) (4a) (4a) (4a) (4a) 1
Λhhhp = Λhhph = Λhphh = Λ phhh = + ,
+ nk (1 − 4nl )(1 − 3ni )(1 − nl ni n j ) 4
+n j (1 − 4nk )(1 − 3nl )(1 − nk nl ni )] , (17) (4a) (4a) (4a) (4a) 1
Λ pphh = Λhhpp = Λhpph = Λ phhp = − ,
1 2
(4b) 
Λi jkl = ni nk [(1 − 2nl )(1 − nl n j ) + (1 − 2n j )(1 − n j nl )] (4a) (4a)
Λ phph = Λhphp = −1,
8
+ n j nk [(1 − 2nl )(1 − nl ni ) + (1 − 2ni )(1 − ni nl )] (4b) (4b) (4b) (4b) 1
Λ phph = Λhphp = Λhpph = Λ phhp = + , (23)
+ ni nl [(1 − 2nk )(1 − nk n j ) + (1 − 2n j )(1 − n j nk )] 4

+n j nl [(1 − 2nk )(1 − nk ni ) + (1 − 2ni )(1 − ni nk )] . where p stands for a particle and h stands for a hole state. The
(18) non-zero terms of Eq. (19) are given by
(3)
Comparing Eqs. (12) and (13) with Eqs. (14) and (15), we see δ ρi j δ ρki δ ρ jk =Λ jik . (24)
that the quantum and classical averages coincide if we have,
Similarly, the non-zero terms of the first and second terms of
Eq. (20), by taking into account the Kronecker delta functions,
(3)
δ ρiλj (0)δ ρklλ (0)δ ρmn
λ (0) =δ δ δ Λ ,
il jm kn jik (19) read as
λ (0)δ ρ λ (0) =δ δ δ δ Λ(4a)
δ ρiλj (0)δ ρklλ (0)δ ρmn (4a)
pr il kn rm j p jikr δ ρi j δ ρki δ ρrk δ ρ jr = Λ jikr , (25)
(4b)
+ δil δ jk δrm δ pn 3Λ jipr .
and
(20)
(4b)
δ ρi j δ ρ ji δ ρrp δ ρ pr = 3Λ jipr , (26)
By considering that all initial densities ρ λ (0) in the en- respectively.
semble are Hermitian (In the following, we drop the labels The form of the last three equations suggests that these
for brevity, i.e. δ ρi j ≡ δ ρiλj (0).), it immediately follows from equations can be satisfied if one considers that the stochas-
Eq. (10) that (δ ρii )2 = 0 which means that tic matrix elements are correlated with each other. However,
averages that contain terms with the same indices δ ρ j j in
(3)
δ ρii = 0 (21) Eqs. (24) and (25) lead to, for instance, δ ρi j δ ρ ji δ ρ j j = Λ ji j
(4a)
and δ ρi j δ ρii δ ρ ji δ ρ j j = Λ jii j , where left-hand side of these
and hence the matrix elements ρii are not fluctuating. Further- equations is zero due to Eq. (21) and the right-hand side of the
more, for i 6= j, we have equations is non-zero as seen from Eq. (23). Hence, Eqs. (24)
and (25) cannot be satisfied completely even with correlated

1 matrix elements. The reason for such a discrepancy is related
 2 , (i, j) = (p, h) to the fact that classical probability distributions cannot sim-
ri2j + s2i j = (22) ulate quantum mechanical systems. In conclusion, classical
0 , otherwise,

distributions can only approximately match higher quantum
mechanical moments. Here, we explore classical probability
where we introduced the real and imaginary parts of the ma- distributions that can at best approximate the third and fourth
trix element dispersions, δ ρi j = ri j + i si j . Here, (i, j) = (p, h) quantum mechanical moments.
means that one of the states is a particle state and the other In this study, we investigate the consequences of the sim-
a hole state. The variance cannot be negative and hence, ple case of uncorrelated matrix elements. Hence, we assume
from Eq. (22), we have ri2j = s2i j = 0 for (i, j) = (p, p) or that each stochastic matrix element is statistically independent
(i, j) = (h, h) which means that these matrix elements are from the others as well as that the real and imaginary parts of
zero. Note that the stochastic matrix elements satisfy the a matrix element are also statistically independent. Then, for
equations r ji = ri j and s ji = −si j due to Hermiticity. simplicity, we set δ ρi j δ ρ ji δ ρi j = ri3j + is3i j = 0 for the third
In order to deduce the properties required for δ ρ to fulfill central moments which means,
the moments Eqs. (19) and (20), we need to analyze the pos-
sible values of the Λ terms. In general, from Eqs. (16-18), the ri3j = s3i j = 0. (27)
4

This is in accordance with the original formulation of the SMF


0.8
approach since the third central moment of a mean-zero Gaus-
γ = 3.0
sian random number is zero. Note that Eq. (24) does not pro- 0.7 γ = 1.8
vide any restrictions for the δ ρi j δ ρ ji δ ρi j moments. However,
a condition for the fourth central moments can be obtained 0.6 γ = 1.0

F (χ, γ)
by assuming uncorrelated matrix elements. It is seen from
0.5
Eqs. (25) and (26) that one has,
(4a) (4b)
0.4
δ ρi j δ ρ ji δ ρi j δ ρ ji = Λ ji ji + 3Λ ji ji , (28)
0.3
which leads to the result,
0.2
0 0.25 0.5

 − 41 , if (i, j) = (p, h)
ri4j + s4i j + 2 ri2j s2i j = (29) χ
0 , otherwise,

FIG. 1. The plot of the function, Eq. (30), versus the parameter χ for
where Eq. (23) was used. three different kurtoses γ. γ = 3, γ = 1.8, and γ = 1 correspond to
No classical probability distribution can have a negative the Gaussian, uniform, and two-point distributions, respectively.
second and/or fourth central moment. Hence, the left hand
side of Eq. (29) can never be negative. However, the distribu-
tions that we want to obtain are quantum mechanical in nature. kurtosis values 3 and 1.8 which are the kurtoses of the Gaus-
Actually the non-commutativity of operators within quantum sian and the uniform distributions, respectively. The kurtosis
mechanics is the main reason for inadequecy of classical dis- value of 1 corresponds to the minimum possible value. The
tributions to simulate the quantum systems. In the phase- probability distribution function with the minimum kurtosis is
space formulation of quantum mechanics, we already know the so-called two-point distribution function which can take
such behaviors coming from Wigner distribution which can only two values σ and −σ with equal probabilities [22]. This
assume negative values and therefore is called a quasiproba- probability function can be written as,
bility distribution [21]. It is important to note that positive-
definite phase-space distributions such as Husimi distribu- 1 1
P2p (x) = δ (x + σ ) + δ (x − σ ), (32)
tion exist, however they are still quasiprobability distributions 2 2
since the averaging are taken by the Weyl symbol of the oper- where δ (x) is the Dirac delta function and σ is the standard
ators rather than the operators themselves [10]. deviation of x. It is seen from Fig. 1 that the function F de-
Based on the discussion above, since a classical probabil- creases as the kurtosis γ decreases. We show below that the
ity distribution cannot lead to a negative fourth central mo- SMF dynamics with a small kurtosis probability distribution
ment, we anticipate that an efficient approximation would be provides a better approximation to the exact dynamics than the
to consider a distribution which minimizes the left-hand side SMF dynamics with a large kurtosis distribution. In particular,
of Eq. (29) while satisfying Eq. (22). The equation (22) sets for the three distributions we considered here, the two-point
a condition on the sum of the variances and hence it does not distribution provides the best approximation and the Gaussian
impose any conditions on the weights to the variances of the distribution provides the worst approximation. The choice of
real and imaginary parts. In order to investigate the depen- the weight parameter χ also has an impact on the values of
dence to the weights, we define a parameter χ = ri2j . From the function F as seen from Fig. 1. The equal weight case,
Eq. (22), we have s2i j = 1/2 − χ where 0 ≤ χ ≤ 1/2. Sub- χ = ri2j = s2i j = 1/4 provides the minimal value of F for a
stituting these results into the left-hand side of Eq. (29) we fixed value of the kurtosis γ except for the two-point distribu-
obtain the function, tion which is constant. Hence, the case with equal weights for
the variances of the real and imaginary parts of the stochastic
γ
F(χ, γ) = 2(γ − 1)χ 2 − (γ − 1)χ + , (30) matrix elements is anticipated to provide a better approxima-
4 tion than any unequal weights case.
where we used the kurtosis of the distribution which is defined
as
III. APPLICATIONS
ri4j
γ =  2 . (31) In this section, we apply the SMF approach with the two-
ri2j point distribution as well as the Gaussian and uniform distri-
butions on an exactly solvable model and show that indeed the
The kurtosis is an invariant form of the fourth central moment SMF dynamics with the two-point distribution for the matrix
that only depends on the distribution and does not depend on elements gives a better agreement with the exact dynamics
the particular values of the variance. Figure 1 shows the plot than the SMF dynamics with the uniform or Gaussian distri-
of the function, Eq. (30), versus the weight parameter χ for the butions.
5

A. A Modified Lipkin-Meshkov-Glick Model In practice, there are several numerical methods to solve this
equation of motion. We use the iterated Crank-Nicolson
We consider a modified Lipkin-Meshkov-Glick model method for computing the exact dynamics [24].
(mLMG) that has been introduced in ref. [20]. It has a pair- The mean-field equation of motion (or TDHF equation) can
ing Hamiltonian structure and the single particle energies are be derived by using Ehrenfest theorem,
stochastically distributed over the levels allowing for a de-
scription of dissipation. The Hamiltonian of this model reads dραβ dha†β aα i
ih̄ = ih̄
H = T +V (33) dt dt

= h[aβ aα , H]i, (37)
with
sα εα †
T =∑ a aα , (34) where the expectation value is taken with respect to a Slater
α 2 α determinant. By using Eq. (33), the last equation gives
V = v0 S+ S− , (35)
dραβ 1
where α = (sα , mα ), sα ∈ {−1, +1}, mα ∈ {− j, − j +

ih̄ = sα εα − sβ εβ ραβ
1, ..., j − 1, j}, S+ = ∑α>0 a†α a†α , and S− = S+†
. Here, α = dt 2  
(sα , −mα ) and α > 0 means mα > 0. v0 is the interaction + v0 ∑ ργα ργβ − ραγ ρβ γ
strength and εα /2 are the single particle energies which are γ>0
random numbers with some variance σε2 . The mean values of
 
− v0 ∑ ργα ργβ − ραγ ρβ γ . (38)
the single particle energies in the upper level sα = 1 and in the γ>0
lower level sα = −1 are ∆/2 and −∆/2, respectively. In or-
der to avoid any ambiguity, we should state that the stochastic The SMF equation is directly obtained from Eq. (38) by re-
single-particle energies are only set once and these values are placing all one-body densities with the stochastic ones, ρ →
used for the dynamics. Figure 2 shows a schematic illustration ρλ .
of the mLMG model. Two transitions are indicated on the fig- There are two different single particle bases that are used
ure. Note that, contrary to the original LMG model [23], there in this study. One of the bases is the fixed basis of the model
are transitions within the same level which is due to the fact for which we use the labels α, β , etc. The other basis is the
that the single particle energies are different. natural basis in which the initial one-body density is diagonal
and we use the labels i, j, etc. for the states in this basis. Note
that the MF equation (38) is written in the fixed basis whereas
the properties of the matrix elements of the initial stochastic
one-body densities in the SMF approach are derived in the
natural basis in section II. We solve the SMF equations of
motion in the fixed basis. Hence, for each event we sample
an initial stochastic one-body density in the natural basis, we
write it in the fixed basis and then evolve it with the SMF
equation which is just the MF equation.

C. The initial state

We consider the half-filled case for which there are N parti-


cles for 2N single particle states with initial states of the form,

|Ψ(0)i = eiµD |Φ(η) i, (39)

FIG. 2. Schematic illustration of the mLMG model. Two transitions where µ is a real parameter and |Φ(η) i is a Slater determinant
are indicated one within the same level (green arrows) and another given by
between the two levels (yellow arrows).
(η) (η) (η)
|Φ(η) i = (a†α1 )nα1 (a†α2 )nα2 ... (a†α2N )nα2N |0i. (40)

B. Exact, mean-field and stochastic mean-field dynamics (η)


Here, nα are occupation numbers with values 0 or 1 corre-
sponding to the single particle states indicated in Fig. 2. The
The exact dynamics, for an initial state |Ψ(0)i, is formally dipole operator D reads
given by
D = ∑(a†+1,mα a−1,mα + a†−1,mα a+1,mα ). (41)
|Ψ(t)i = e−iHt/h̄ |Ψ(0)i. (36) mα
6

Since D is a one-body operator the initial state |Ψ(0)i is also 6


a Slater determinant due to the Thouless theorem [25]. The (a)
effect of the exponential term in Eq. (39) is to introduce some 4
excitation by an instant dipole boost [20].
2
We consider two different initial states for N = 6 particles.
The first state corresponds to the ground state, |Φ(0) i, with

hDi
0
all the particles in the lower level (sα = −1) excited by the
dipole boost in Eq. (39) with µ = 0.8. The second state is −2
found by applying again the dipole boost with µ = 0.8 to the −4 exact
state |Φ(η) i for which MF
(b)
     
1 1 3 4
(sα , mα ) = +1, ± , −1, ± , +1, ± (42)
2 2 2
2
single particle states are occupied (see Fig. 2). The first and 0

hDi
second states are denoted by |Ψ(0)i1 and |Ψ(0)i2 , respec-
−2
tively.
Note that the schematic model in Fig. 2 is shown for −4 exact
SMF (G)
10 particles in 20 states. Here, we consider 6 par- −6
SMF (U)
ticles in 12 states, therefore mα can take the values −8 SMF (T)
{−5/2, −3/2, −1/2, 1/2, 3/2, 5/2}. 1.4
1.2 (c)
1
D. Results 0.8
S/N
0.6 exact
In the following computations, the units of energy will be 0.4 SMF (G)
written in terms of the level spacing ∆ and the time units in 0.2 SMF (U)
terms of ∆−1 . The single particle energies are chosen as 0 SMF (T)

ε(+1,±5/2) /2 =0.225 ∆, ε(−1,±5/2) /2 = −0.222∆, 0 20 40 60 80 100


time (∆−1 )
ε(+1,±3/2) /2 =0.697 ∆, ε(−1,±3/2) /2 = −0.593 ∆,
ε(+1,±1/2) /2 =0.578 ∆, ε(−1,±1/2) /2 = −0.685 ∆. (43)
FIG. 3. Comparison of the dynamical evolutions of the expectation
The mean of the single particle energies in the upper level value of the dipole operator D is illustrated for the exact (black line)
ε(+1,mα ) /2 and in the lower level ε(−1,mα ) /2 are ∆/2 and and MF solutions (orange line with circles) (a), and for the exact
−∆/2, respectively. The standard deviation of the εα /2 values and SMF solutions with Gaussian (G) (red line with boxes), uniform
(U) (green line with trangles), and two-point (T) distributions (blue
in each level is σε = 0.2 ∆. All the SMF results are obtained
line with diamonds) (b). The dynamics of the one-body entropy per
by considering N = 106 events. particle S/N for the exact and SMF solutions with the same three dis-
We follow the dynamics of the dipole operator D and the tributions is indicated in the lower panel (c). The interaction strength
fermionic one-body entropy given by is v0 = 0.05 ∆ and the initial state is |Ψ(0)i1 .
S = −Tr [ρ ln ρ + (1 − ρ) ln(1 − ρ)] , (44)
where ρ is the one-body density operator. The one-body en- by Eq. (44), per particle versus time for the weak coupling
tropy is a measure of departure from an uncorrelated (Slater) v0 = 0.05 ∆. The initial state is |Ψ(0)i1 which is explained in
state and a measure of thermalization [20]. For a Slater state, the subsection III C. In Fig. 3(a), it is observed that MF evo-
the entropy gets the value S = 0. Hence, during the MF dy- lution starts to deviate from the exact one at around t = 6 ∆−1
namics the entropy remains zero. The maximum value of the whereas in Fig. 3(b) the SMF evolutions start to deviate later
entropy for the half-filled case is obtained when all single par- at around t = 20 ∆−1 . Comparing the three SMF dynamics
ticle states are half-filled leading to the value, with the exact one in Fig. 3(b), we see that the SMF approach
S = 2N ln 2. (45) with the two-point distribution is much better than the other
two distributions at all times. A similar behavior is seen for
In the SMF approach, the one-body entropy is computed by the one-body entropy S in Fig. 3(c). The SMF evolution of the
the expression entropy with the two-point distribution follows the exact so-
h i lution very closely, whereas SMF solutions with the Gaussian
S = −Tr ρ λ ln ρ λ + (1 − ρ λ ) ln(1 − ρ λ ) . (46) and uniform initial distributions are underestimating the exact
entropy until t = 30 ∆−1 . After that time, the exact entropy de-
Figure 3 shows the expectation value of the dipole opera- creases which is followed by the SMF solution with two-point
tor D, given by Eq. (41), and the one-body entropy S, given distribution whereas the solutions with the Gaussian and uni-
7

6 4
4 2
2 0

hDi
hDi

0 −2 exact
MF
−2 −4
SMF (G)
−4 −6 SMF (U)
exact (a)
(a) SMF (T)
MF
1.4
4 1.2
1
2 0.8

S/N
exact
0.6
hDi

0 SMF (G)
0.4 SMF (U)
−2 exact 0.2 SMF (T) (b)
SMF (G) 0
−4 SMF (U) 0 1 2 3 4 5 6 7 8 9 10
SMF (T) (b)
time (∆−1 )
1.4
1.2
1 FIG. 5. Comparison of the dynamical evolutions of the expecta-
tion value of the dipole operator D is illustrated for the exact (black
S/N

0.8
0.6 exact line) and MF solutions (orange line with circles), SMF solutions with
SMF (G) Gaussian (G) (red line with boxes), uniform (U) (green line with tran-
0.4
SMF (U) gles), and two-point (T) distributions (blue line with diamonds) (a).
0.2
SMF (T) (c) The dynamics of the one-body entropy per particle S/N for the exact
0 and SMF solutions with the same three distributions is indicated in
0 100 200 300 400 500 the lower figure (b). The interaction strength is v0 = 0.5 ∆ and the
time (∆−1 ) initial state is |Ψ(0)i1 .

FIG. 4. Same as Fig. 3 for long time evolution


t = 50 ∆−1 . Figure 4(c) shows the similar behavior for the en-
tropy S. The SMF solutions with Gaussian and uniform distri-
form distributions continue to increase and approach the max- butions attain the maximum entropy value around t = 50 ∆−1
imum value 2 ln 2 = 1.38. The solution with the uniform dis- whereas the SMF solution with two-point distribution is in
tribution has a small decrease from t = 40 ∆−1 to t = 70 ∆−1 , better agreement with the exact entropy for longer times.
hence it is slightly better than the solution with the Gaussian The dynamics in the strong coupling case, v0 = 0.5 ∆, is
distribution. Note that as the kurtosis, Eq. (31), of the dis- indicated in Fig. 5. It is seen that the MF solution of hDi
tribution decreases the SMF dynamics becomes better. The deviates from the exact one at t = 0.7 ∆−1 and SMF solu-
best approximation is attained for the two-point distribution, tions start to deviate later at around t = 2 ∆−1 . In Fig. 5(a)
Eq. (32), with the minimum kurtosis value of 1. We clearly and 5(b), the difference between the SMF results with the
see in Fig. 3 that the use of the two-point distribution strongly three distributions is almost negligible. However, when the
increases the time scale over which the SMF approach is pre- plots at t = 2.1 ∆−1 are zoomed in, as seen from the inset fig-
dictive. ures, SMF solution with the two-point distribution turns out to
Figure 4 is the same with Fig. 3 except that the long time be the best approximation to the exact solution and the SMF
dynamics is shown. In Fig. 4(a), the exact solution exhibits solution with the Gaussian distribution is the worst approxi-
oscillations with decreasing amplitude for long times, on the mation. It is known that the validity time of MF as well as
other hand the MF solution has beating-like oscillations with beyond MF approximations decreases inversely proportional
almost constant amplitude. This result and the fact that the to the coupling strength, ∆tval ∝ v−1
0 , [9,26]. Here, we observe
one-body entropy in MF approximation remains zero at all the same behavior by comparing Figs. 3 and 5. Furthermore,
times clearly show the well-known shortcoming of MF dy- as the coupling strength increases the difference between SMF
namics, that is the underestimation of dissipation and hence evolutions with different initial distributions decreases.
thermalization in quantum many-body systems. In Fig. 4(b), In Fig. 6, we investigate the dependence of the SMF dynam-
the SMF solution with two-point distribution follows the am- ics on the parameter χ that controls the weights of the real
plitudes of the oscillation of the exact dynamics for a longer and imaginary parts of the stochastic matrix elements of the
time than the SMF solutions with Gaussian and uniform dis- one-body densities ρ λ . Three cases are considered: the first
tributions which become almost completely damped around one corresponds to the equal weights case, labeled as R+I, for
8

10
1.6 exact R+I R I
R+I
1.2
R
I
5
S/N

0.8 t=0

frequency (arb. units)


0.4
0
Gaussian distribution (a)
0 t = 2 ∆−1
1.6 exact R+I R I
5
1.2
S/N

0.8
0
0.4 t = 4 ∆−1
0 Uniform distribution (b)
5
1.6 exact R+I R I
1.2
S/N

0.8 0
-1 -0.5 0 0.5 -0.5 0 0.5 1
0.4
rαβ sαβ
0 Two-point distribution (c)
0 20 40 60 80 100 FIG. 7. Distributions of the real rαβ (the subfigures on the left)
time (∆−1 ) and imaginary sαβ parts (the subfigures on the right) of the matrix
λ of the stochastic one-body densities ρ λ at the times
elements ραβ
FIG. 6. SMF evolutions of the entropy per particle S/N are plotted t = 0 (upper subfigures), t = 2 ∆−1 (middle subfigures), and t = 4 ∆−1
for different weights of the variances to the real ri j and imaginary si j (lower subfigures). The single particle states are α = (+1, +1/2) and
parts of the stochastic density matrix elements. The SMF solutions β = (−1, +3/2). The initial distribution of the matrix elements is the
are obtained with the Gaussian (a), uniform (b), and two-point (c) two-point distribution. The same cases with those in Fig. 6 namely
distributions. The exact solutions are indicated by black lines. The R+I (blue solid line), R (red dotted line), and I (green dashed line)
SMF evolutions with two-point distribution are shown for: the equal are indicated. The interaction strength is v0 = 0.05 ∆ and the initial
weight case, ri2j = s2i j = 1/4, which is labeled as R+I (blue line with state is |Ψ(0)i1 .

diamonds); the full weight to the real parts, ri2j = 1/2 and s2i j = 0,
which is labeled as R (red line with boxes); and the full weight to the
imaginary parts, s2i j = 1/2 and ri2j = 0, which is labeled as I (green by Eq. (30), assumes smaller values. In Figs. 6(a) and 6(b),
line with triangles). The interaction strength is v0 = 0.05 ∆ and the The SMF solutions with the Gaussian and uniform distribu-
initial state is |Ψ(0)i1 . tions are consistent with this result. However, the weight pa-
rameter χ should not affect the quality of the SMF evolution
with the two-point distribution since F is constant for this dis-
tribution as seen from Fig. 1. In Fig. 6(c), it is observed that
which χ = ri2j = s2i j = 1/4; the second one corresponds to full the equal weight case is still much better than unequal weight
weight to the real parts, labeled as R, for which χ = ri2j = 1/2 cases. This can be explained by observing the distributions of
real and imaginary parts of the stochastic matrix elements. As
and s2i j = 0; and the third case is the opposite, full weight an illustration, in Fig. 7, we show the distributions of the real
to the imaginary parts, labeled as I, for which s2i j = 1/2 and rαβ and imaginary sαβ parts of a matrix element ραβ λ of the

χ = ri2j = 0. It is clearly seen that the equal weight case gives stochastic one-body densities ρ λ at the times t = 0, t = 2 ∆−1 ,
the best dynamics with all three distributions. When the imag- and t = 4 ∆−1 . In this figure, the single particle states are arbi-
inary matrix elements are set to zero, case R, the SMF entropy trarily chosen as α = (+1, +1/2) and β = (−1, +3/2). The
is overestimating the exact one and when the real matrix ele- initial distribution of the matrix elements is the two-point dis-
ments are set to zero, case I, the SMF entropy is underestimat- tribution. The same cases with those in Fig. 6 namely R+I,
ing the exact entropy. As discussed in section II, the quality of R, and I are indicated. It is observed that when the imagi-
the SMF approximation increases when the function F, given nary (real) parts of the matrix elements are set to zero, case R
9

SMF (G) SMF (G)


0.8 SMF (U) 0.8 SMF (U)
SMF (T) SMF (T)
0.6 0.6
t=0 t=0
0.4 0.4

0.2 0.2
(a) (a)
0 0
SMF (G)
frequency (arb. units)

SMF (G)

frequency (arb. units)


0.8 SMF (U) 0.8 SMF (U)
SMF (T) SMF (T)
0.6 0.6
t = 2 ∆−1 t = 0.6 ∆−1
0.4 0.4

0.2 0.2
(b) (b)
0 0
SMF (G) SMF (G)
0.8 SMF (U) 0.8 SMF (U)
SMF (T) SMF (T)
0.6 0.6
t = 4 ∆−1 t = 1.2 ∆−1
0.4 0.4

0.2 0.2
(c) (c)
0
-10 -8 -6 -4 -2 0 2 4 6 8 10 -10 -5 0 5 10 15
hDiλ hDiλ

FIG. 8. Distribution of the ”event“ expectation values of the dipole FIG. 9. Same as Fig. 8 except that the system consists of N = 12
operator D in the ensemble are compared for the Gaussian (G) (red particles in 24 single particle states. The distributions of hDiλ are
dotted line), uniform (U) (green dashed line), and two-point (T) compared at the times t = 0 (a), t = 0.6 ∆−1 (b), t = 1.2 ∆−1 (c).
distributions (blue solid line) at the times t = 0 (a), t = 2 ∆−1 (b),
t = 4 ∆−1 (c). The interaction strength is v0 = 0.05 ∆ and the initial
state is |Ψ(0)i1 .
such as t = 2 ∆−1 for all the distributions. At this time, even
the MF solution agrees well with the exact solution as seen
(I), the real (imaginary) parts in the figure assume three val- from Fig. 3(a). Based on these observations and the fact that
ues with equal probability whereas the imaginary (real) parts the correlations between the real and imaginary parts of the
are zero at the initial time t = 0. After a very short time matrix elements of the one-body density are governed by the
interval, at t = 2 ∆−1 , both the real and imaginary parts de- MF equation, we think that the quality of the SMF approach
velop similar three-peak structures. However, after the same for different weights cases, such as the cases R and I, is un-
amount of time, at t = 4 ∆−1 , opposite distribution structures related to the choice of the initial distributions in the SMF
for the R and I cases are observed with respect to the dis- approach. Hence, the equal weight case R+I should always
tributions at time t = 0. At time t = 4 ∆−1 , the real (imag- be considered as a better approximation over unequal cases
inary) parts develop a narrow distribution around the value due to dynamical correlations, governed by the MF equation,
zero compared to the imaginary (real) parts that have a wide between the real and imaginary parts of the matrix elements.
three-peak distribution for the case R (I). The case R+I ex- Figure 8 shows the distribution of the ”event“ expectation
hibits almost symmetrical distributions for the real and imag- value, defined by Eq. (3), of the collective observable D for
inary parts at all times. These results clearly show how the the three probability distributions at three different times. The
dynamical correlations between the real and imaginary parts distributions of hDiλ are continuous and Gaussian for the ini-
of the matrix elements build up in time. In particlular, the real tial Gaussian and uniform distributions at all times. However,
and imaginary parts are balancing each other for the equal for the initial two-point distribution, the resulting distribution
weight case R+I whereas the distributions are changing be- of hDiλ is discrete. This result is directly related to the fact
tween narrow single-peak around zero and a wide three-peak that the two-point distribution is discrete. It is seen that in
distribution for the unequal weights cases R and I. In Fig. 6, a very short time interval 4 ∆−1 the result with the two-point
it is seen that the entropies for the cases R and I start to devi- distribution becomes also Gaussian since the correlations be-
ate from the exact and R+I case solutions at very short times tween the stochastic matrix elements of the one-body density,
10

2 2

1 1

0 0
hDi

hDi
−1 −1

−2 −2
exact MF (a) exact MF (a)
2 2

1 1

0 0
hDi

hDi
-1 -1

-2 -2
exact SMF (G) (b) exact SMF (G) (b)
0.8 0.8
SMF (U) SMF (T) SMF (U) SMF (T)
0.6 0.6
S/N

0.4 S/N 0.4

0.2 0.2
(c) (c)
0 0
0 20 40 60 80 100 0 100 200 300 400 500
time (∆−1 ) time (∆−1 )

FIG. 10. Same as Fig. 3 except that the initial state is |Ψ(0)i2 . FIG. 11. Same as Fig. 10 for long time evolution

governed by the stochastic version of the mean-field equation Figs. 10 and 11. Note that these figures are the same with
(38), build up very fast. Figure 9 is the same with Fig. 8 ex- Figs. 3 and 4 except that the initial state is different. The MF
cept that we consider a system with N = 12 particles in 24 evolution with this initial state results in a sinusoidal oscilla-
single particle states. Here, it is observed that the number of tion with constant amplitude for the dipole moment hDi and
discete peaks increases linearly with particle number and also the exact evolution demonstrates a beating oscillation with al-
that the time interval after which a Gaussian distribution is most equal amplitudes for the beatings as seen in Figs. 10(a)
obtained is even shorter than that for N = 6 particles. The and 11(a). We see that taking a different initial state leads
behaviors here are closely related to the central limit theo- to the same conclusion that the SMF evolution with the two-
rem which states that the sum of many independent random point distribution leads to a better agreement with the exact
variables tends toward a Gaussian distribution. Note that Fig- evolution compared to the SMF evolution with the Gaussian
ures 8 and 9 show how fast the discrete initial distribution of or uniform distributions. In Figs. 10(c) and 11(c) the system
hDiλ transforms to a Gaussian distribution and hence how fast does not reach the maximum entropy value 2 ln 2 due to the
the correlations between the matrix elements of the one-body symmetry of the initial state. This is also in accordance with
density build up. If the dynamical evolution of the distribu- observation of the beating oscillations of the exact evolution.
tion of hDiλ in Figs. 8 or 9 is followed for longer times (not SMF evolution with the two-point distribution is able to repro-
shown for brevity), it is observed that the Gaussian shape is duce the beatings with decreasing amplitudes in time whereas
preserved however centroids and widths of the distributions SMF evolutions with Gaussian and uniform distributions get
of hDiλ for the two-point, Gaussian, and uniform distributions damped in a very short time.
deviate from each other since the corresponding SMF means
and variances of D deviate from each other as well.
Similar results are reached for the SMF evolutions with dif- IV. CONCLUSIONS
ferent initial states. Here, we also demonstrate a result with
a different initial state |Ψ(0)i2 which is explained in the sub- In the SMF approach, expectation values of observables are
section III C. The corresponding solutions are illustrated in obtained by statistical averaging over an ensemble of stochas-
11

tic one-body densities. It is through matching these expecta- ACKNOWLEDGMENTS


tion values to the corresponding quantum expectation values
of collective observables at initial time that provides informa-
tion about the properties of the initial distribution of the matrix
B.Y. gratefully acknowledges useful discussions with
elements of the stochastic one-body densities. In the original
Phuong Mai Dinh. This work was supported in part by
formulation of the SMF approach, the matching of the expec-
the Scientific and Technological Research Council of Turkey
tation values is performed for the quantum means and vari-
(TUBITAK).
ances of collective observables which provides relations for
the means and (co)variances (first and second moments) of
the stochastic one-body density matrix elements that are then
assumed to be Gaussian random numbers [6]. In the present
work, we have investigated the properties of the stochastic
one-body densities within the SMF approach by considering
higher order moments. An expression for the fourth central
moments of the matrix elements is derived which suggest that
a probability distribution with minimal kurtosis for the ini-
tial density matrix elements provides the best approximation Appendix A: Quantum moments of a one-body operator
to the exact dynamics when the stochastic matrix elements
are assumed to be statistically independent random numbers.
The distribution with minimal kurtosis is the so called two-
Let us consider that a many-body system is described by a
point distribution which can take only two values and hence
Slater determinant and that the corresponding one-body den-
is discrete. The SMF approach with the two-point distribu-
sity is ρ. In terms of the natural basis {|ii}, which are sat-
tion as well as the Gaussian and uniform distributions is ap-
isfying hi|ρ| ji = ni δi j where the occupation numbers ni can
plied to an exactly solvable model that is inspired from the
take values 0 or 1, a one-body operator can be written as
Lipkin-Meshkov-Glick model that is familiar from nuclear
A = ∑i j Ai j a†i a j . The first moment is the expectation value
physics [20]. It is shown that indeed the two-point distribu-
given by
tion is the best approximation for the matrix elements of the
stochastic one-body densities in the weak as well as strong
coupling strengths. However, in the strong coupling case, the
difference among the SMF evolutions with different distribu-
tions is almost negligible. hAi =Tr(ρA)
= ∑ ni Aii . (A1)
i

MF dynamics in it its various forms such as TDHF and DFT


is widely used in many fields of physics. However, MF dy-
namics is known to contain some drawbacks. Quantum corre- The second moment is obtained as
lations in collective observables are mostly missing hamper-
ing the use of the MF dynamics for long time evolutions. Fur-
thermore, the dissipation of the collective motion is severly
underestimated. The SMF approach cures these drawbacks to
some extent by taking into account the initial quantum fluctu- hA2 i =Tr1 (ρ1 A21 ) + Tr12 (ρ12 A1 A2 )
ations. In this work, it is observed that the long time correla- = ∑ ni (1 − n j )Ai j A ji + ∑ ni n j Aii A j j , (A2)
tions are much better reproduced with the two-point distribu- ij ij
tion than the Gaussian or uniform distributions which result in
overdamping and hence faster thermalization.

where ρ12 = ρ1 ρ2 (1 − P12 ). Hence, the second central mo-


ment is given by
The discrete form of the two-point distribution results in a
discrete initial distributions of the collective observables with
the tips of the discrete values having a Gaussian shape. Due to
correlations between the matrix elements supplied by the self-
consistent SMF equation, the distribution of the collective ob- h(A − hAi)2 i =hA2 i − hAi2
servables reaches a Gaussian distribution very fast which is in = ∑ ni (1 − n j )A ji Ai j
accordance with the central limit theorem. The discrete nature ij
of the physical observables and the matrix elements of the ini- 1
tial stochastic one-body densities opens up the possibility to = ∑ [ni (1 − n j ) + n j (1 − ni )] A ji Ai j . (A3)
2 ij
perform SMF dynamics with small number of events.
12

The third moment of A reads The fourth moment of A reads

hA3 i =Tr1 (ρ1 A31 ) + 2Tr12 (ρ12 A21 A2 ) + Tr12 (ρ12 A1 A22 )
hA4 i =Tr1 (ρ1 A41 ) + Tr12 ρ12 (4A31 A2 + 3A21 A22 )
 
+ Tr123 (ρ123 A1 A2 A3 )
= ∑ ni Ai j A jk Aki + 6Tr123 (ρ123 A21 A2 A3 )
i jk + Tr1234 (ρ1234 A1 A2 A3 A4 )

+ 2 ∑ ni n j Aik Aki A j j − A jk Aki Ai j = ∑ ni Ai j A jk Akl Ali
i jk i jkl

+ ∑ ni n j Aii A jk Ak j − A ji Aik Ak j + 4 ∑ ni n j Aik Akl Ali A j j − A jk Akl Ali Ai j

i jk i jkl
+ ∑ ni n j nk (Aii A j j Akk − Aii A jk Ak j + 3 ∑ ni n j (Aik Aki A jl Al j − A jk Aki Ail Al j )
i jk i jkl
− Ai j A ji Akk − Aik A j j Aki + 6 ∑ ni n j nk (Ail Ali A j j Akk − Ail Ali A jk Ak j
+ Ai j A jk Aki + A ji Ak j Aik ) i jkl

= ∑ ni Ai j A jk Aki − Akl Ali A j j Aik − A jl Ali Ai j Akk


i jk + 2Ai j A jk Akl Ali )

+ 3 ∑ ni n j Aik Aki A j j − A jk Aki Ai j + ∑ ni n j nk nl (Aii A j j Akk All − 6Ai j A ji Akk All
i jk i jkl
+ ∑ ni n j nk (Aii A j j Akk − 3Aii A jk Ak j + 8Ai j A jk Aki All + 3Ai j A ji Akl Alk
i jk
− 6Ai j A jk Akl Ali ), (A7)
+ 2Ai j A jk Aki ), (A4)

where ρ123 = ρ1 ρ2 ρ3 (1 − P12 )(1 − P13 − P23 ). The third cen- where ρ1234 = ρ1 ρ2 ρ3 ρ4 (1 − P12 )(1 − P13 − P23 )(1 − P14 −
tral moment of A reads P24 − P34 ). The fourth central moment of A reads

h(A − hAi)3 i =hA3 i − 3hA2 ihAi + 2hAi3


= ∑ ni (1 − 3n j + 2n j nk )Ai j A jk Aki h(A − hAi)4 i =hA4 i − 4hA3 ihAi + 6hA2 ihAi2 − 3hAi4
i jk
= ∑ ni (1 − 4n j − 3nk + 12n j nk − 6n j nk nl )
= ∑ ni (1 − 3n j )(1 − n j nk )Ai j A jk Aki i jkl
i jk
× Ai j A jk Akl Ali
= ∑ Λi jk Ai j A jk Aki , (A5)
i jk + 3 ∑ ni nk (1 − 2nl + nl n j )
i jkl
× Ai j A ji Akl Alk
= ∑ ni (1 − 4n j )(1 − 3nk )(1 − n j nk nl )
i jkl
where we used the fact that n2i
= ni for a Slater determi-
nant and Λi jk is the symmetrized version of the term ni (1 − × Ai j A jk Akl Ali
3n j )(1 − n j nk ) given by, + 3 ∑ ni nk (1 − 2nl )(1 − nl n j )
i jkl
× Ai j A ji Akl Alk (A8)
(4a)
=∑ Λi jkl Ai j A jk Akl Ali
(3) 1 i jkl
Λi jk = [ni (1 − 3n j )(1 − n j nk ) + nk (1 − 3ni )(1 − ni n j ) (4b)
3 + 3 ∑ Λi jkl Ai j A ji Akl Alk , (A9)
+n j (1 − 3nk )(1 − nk ni )] . (A6) i jkl
13

(4a) (4b)
where the symmetrized terms are Note that the symmetrized terms Λi jkl and Λi jkl are obtained
by realizing, from Eq. (A8), the following equalities,

Ai j A jk Akl Ali = Ali Ai j A jk Akl


= Akl Ali Ai j A jk
= A jk Akl Ali Ai j , (A12)
(4a) 1
Λi jkl = [ni (1 − 4n j )(1 − 3nk )(1 − n j nk nl )
4 and
+ nl (1 − 4ni )(1 − 3n j )(1 − ni n j nk )
+ nk (1 − 4nl )(1 − 3ni )(1 − nl ni n j ) Ai j A ji Akl Alk = Akl Alk Ai j A ji
+n j (1 − 4nk )(1 − 3nl )(1 − nk nl ni )] , (A10) = A ji Ai j Akl Alk
(4b) 1  = Akl Alk A ji Ai j
Λi jkl = ni nk [(1 − 2nl )(1 − nl n j ) + (1 − 2n j )(1 − n j nl )]
8 = Ai j A ji Alk Akl
+ n j nk [(1 − 2nl )(1 − nl ni ) + (1 − 2ni )(1 − ni nl )] = Alk Akl Ai j A ji
+ ni nl [(1 − 2nk )(1 − nk n j ) + (1 − 2n j )(1 − n j nk )] = A ji Ai j Alk Akl

+n j nl [(1 − 2nk )(1 − nk ni ) + (1 − 2ni )(1 − ni nk )] . = Alk Akl A ji Ai j . (A13)
(A11)

[1] M. J. Davis and E. J. Heller, “Comparisons of classical and ergies,” Phys. Rev. C 83, 064615 (2011).
quantum dynamics for initially localized states,” J. Chem. Phys. [14] B. Yilmaz, S. Ayik, D. Lacroix, and O. Yilmaz, “Nucleon ex-
80, 5036 (1984). change in heavy-ion collisions within a stochastic mean-field
[2] C. W. Gardiner and P. Zoller, Quantum Noise: A Handbook of approach,” Phys. Rev. C 90, 024613 (2014).
Markovian and Non-Markovian Quantum Stochastic Methods [15] Yusuke Tanimura, Denis Lacroix, and Sakir Ayik, “Micro-
with Applications to Quantum Optics (Springer–Verlag, Berlin- scopic Phase–Space Exploration Modeling of 258 Fm Sponta-
Heidelberg, 2004). neous Fission,” Phys. Rev. Lett. 118, 152501 (2017).
[3] Alice Sinatra, Carlos Lobo, and Yvan Castin, “The truncated [16] S. Ayik, B. Yilmaz, O. Yilmaz, A. S. Umar, and G. Turan,
Wigner method for Bose-condensed gases: limits of validity “Multinucleon transfer in central collisions of 238 U +238 U,”
and applications,” J. Phys. B: At. Mol. Opt. Phys. 35, 3599– Phys. Rev. C 96, 024611 (2017).
3631 (2002). [17] S. Ayik, B. Yilmaz, O. Yilmaz, and A. S. Umar, “Quantal dif-
[4] Anatoli Polkovnikov, “Phase space representation of quantum fusion description of multinucleon transfers in heavy–ion colli-
dynamics,” Ann. Phys. 325, 1790–1852 (2010). sions,” Phys. Rev. C 97, 054618 (2018).
[5] Shainen M. Davidson, Dries Sels, and Anatoli Polkovnikov, [18] S. Ayik, B. Yilmaz, O. Yilmaz, and A. S. Umar, “Quantal diffu-
“Semiclassical approach to dynamics of interacting fermions,” sion approach for multinucleon transfers in Xe+Pb collisions,”
Ann. Phys. 384, 128–141 (2017). Phys. Rev. C 100, 014609 (2019).
[6] S. Ayik, “A stochastic mean-field approach for nuclear dynam- [19] Denis Lacroix, Danilo Gambacurta, and Sakir Ayik, “Quan-
ics,” Phys. Lett. B 658, 174 (2008). tal corrections to mean-field dynamics including pairing,” Phys.
[7] Denis Lacroix and Sakir Ayik, “Stochastic quantum dynamics Rev. C 87, 061302(R) (2013).
beyond mean field,” Eur. Phys. J. A 50, 95 (2014). [20] L. Lacombe, E. Suraud, P.-G. Reinhard, and P. M. Dinh,
[8] Denis Lacroix, Sakir Ayik, and Bulent Yilmaz, “Symmetry “Stochastic TDHF in exactly solvable model,” Ann. Phys. 373,
breaking and fluctuations within stochastic mean-field dynam- 216–229 (2016).
ics: Importance of initial quantum fluctuations,” Phys. Rev. C [21] Adam Bednorz and Wolfgang Belzig, “Fourth moments reveal
85, 041602(R) (2012). the negativity of the Wigner function,” Phys. Rev. A 83, 052113
[9] D. Lacroix, S. Hermanns, C. M. Hinz, and M. Bonitz, “Ul- (2011).
trafast dynamics of finite Hubbard clusters:A stochastic mean- [22] J. J. A. Moors, “A quantile alternative for kurtosis,” The Statis-
field approach,” Phys. Rev. B 90, 125112 (2014). tician 37, 25–32 (1988).
[10] Bulent Yilmaz, Denis Lacroix, and Resul Curebal, “Importance [23] H. J. Lipkin, N. Meshkov, and A. J. Glick, “Validity of many-
of realistic phase-space representations of initial quantum fluc- body approximation methods for a solvable model. (I). Exact
tuations using the stochastic mean-field approach for fermions,” solutions and perturbation theory,” Nucl. Phys. A 62, 188–198
Phys. Rev. C 90, 054617 (2014). (1965).
[11] S. Ayik, K. Washiyama, and D. Lacroix, “Fluctuation and dis- [24] Saul A. Teukolsky, “Stability of the iterated Crank-Nicholson
sipation dynamics in fusion reactions from a stochastic mean- method in numerical relativity,” Phys. Rev. D 61, 087501
field approach,” Phys. Rev. C 79, 054606 (2009). (2000).
[12] S. Ayik, B. Yilmaz, and D. Lacroix, “Stochastic semi-classical [25] D. J. Thouless, “Stability conditions and nuclear rotations in the
description of fusion at near-barrier energies,” Phys. Rev. C 81, Hartree–Fock theory,” Nucl. Phys. 21, 225–232 (1960).
034605 (2010). [26] Anatoli Polkovnikov, “Quantum corrections to the dynamics of
[13] B. Yilmaz, S. Ayik, D. Lacroix, and K. Washiyama, “Nucleon interacting bosons: Beyond the truncated Wigner approxima-
exchange mechanism in heavy-ion collisions at near-barrier en- tion,” Phys. Rev. A 68, 053604 (2003).

You might also like