Next Article in Journal
Magnetic Force-Free Theory: Nonlinear Case
Previous Article in Journal
GERARD: GEneral RApid Resolution of Digital Mazes Using a Memristor Emulator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Path Integral Estimates of the Quantum Fluctuations of the Relative Soliton-Soliton Velocity in a Gross-Pitaevskii Breather

1
Alliance School of Applied Mathematics, Alliance University, Bengaluru 562 106, Karnataka, India
2
Department of Physics, University of Massachusetts Boston, Boston, MA 02125, USA
*
Author to whom correspondence should be addressed.
Physics 2022, 4(1), 12-20; https://doi.org/10.3390/physics4010002
Submission received: 11 October 2021 / Revised: 9 December 2021 / Accepted: 16 December 2021 / Published: 10 January 2022
(This article belongs to the Section Statistical Physics and Nonlinear Phenomena)

Abstract

:
In this paper, the quantum fluctuations of the relative velocity of constituent solitons in a Gross-Pitaevskii breather are studied. The breather is confined in a weak harmonic trap. These fluctuations are monitored, indirectly, using a two-body correlation function measured at a quarter of the harmonic period after the breather creation. The results of an ab initio quantum Monte Carlo calculation, based on the Feynman-Kac path integration method, are compared with the analytical predictions using the recently suggested approach within the Bogoliubov approximation, and a good agreement is obtained.

1. Introduction

A purely solitonic solution focusing on nonlinear Schrödinger equation (NLSE) consists of a finite number of solitons, each parameterized by the norm, velocity, initial position, and initial phase. A single stationary soliton is known as the fundamental soliton. When a solution consists of two or more solitons that are at rest relative to each other and have the same initial positions, it is commonly known as a breather. The name comes from the fact that the density profile of this kind of solution periodically oscillates in time (provided the constituent solitons have unequal amplitudes). Certain kinds of breathers can be produced by quenching the strength of the nonlinear interaction. Mathematically, this means the following: let ψ ( x ) be the fundamental soliton of an NLSE, at some point of time. Suppose this ψ ( x ) is used as the initial condition for time propagation under an NLSE whose interaction strength is four times that of the original NLSE. Then, one would find that the propagated solution is a 2-breather whose constituent solitons have norms that are in a ratio of 3:1. Experimentally, this means that a sudden increase of the interaction strength by a factor of 4 converts a fundamental soliton to a 3:1 breather. This was analytically predicted by Satsuma and Yajima [1] in 1974 and recently experimentally verified in dilute Bose-Einstein condensates (BECs) [2,3]. Now, in experiments, there is always a background trapping potential, so that the relevant equation for BEC experiments is the NLSE plus external harmonic confinement. The resulting equation is called the one-dimensional Gross-Pitaevskii equation (GPE). Corresponding to the NLSE breathers, there are Gross-Pitaevskii (GP) breathers. In this paper, a correlation function, associated with a GP breather, corresponding to the 3:1 NLSE breather, is computed using a path integral technique.
Yurovsky et al. [4] showed that quantum many-body effects cause an NLSE breather to disassociate into its constituent solitons. Recall that the 3:1 breather consists of two solitons of norms N / 4 and 3 N / 4 (where N is the number of atoms in the condensate), which, in the mean-field approximation (i.e., at the level of the GPE), sit on the top of each other and do not move. Using the Bethe ansatz, in [4], it was found found that quantization leads to a drift in the relative position of the constituent solitons. Unfortunately, this method cannot be applied to more than N = 23 atoms, and the prediction for experimentally relevant, larger N was an extrapolation. Nevertheless, it was predicted that the relative soliton-soliton velocity in a GP breather may become experimentally observable, for empirically realistic propagation times. Further investigations corroborated this prediction using two different kinds of approximation: the Bogoliubov approach [5] and the truncated Wigner approximation [6]. The predicted quantum fluctuations of macroscopic variables still to be observed but the breather itself is already observed experimentally. From the theory side, however, one can find that there is a need for a nonpertubative, fully quantum-mechanical approach for building a quantum counterpart of the GP breather, an approach that can be applied to a number of particles N that is substantially larger than 23.
In order to estimate the quantum fluctuations of the relative soliton-soliton velocity in a GP breather, we adopt a quantum Monte Carlo method based on the Feynman-Kac path integration [7]. An ab initio confirmation of the observability of the quantum fluctuations of the relative soliton-soliton velocities is presented here using this path integral method [7].
Metropolis and Ulam [8] were the first to exploit a relationship between the Schrödinger equation for imaginary time and the random-walk solution of the general diffusion equation. Let us consider the initial-value problem,
ψ ( x , t ) t = Δ 2 V ψ ( x , t ) , ψ ( x , 0 ) = f ( x ) ,
where x R d , d denotes the dimension, and t is the time. The solution of Equation (1) can be written in FK representation as
ψ ( x , t ) = E x e 0 t V ( X ( s ) ) d s f ( X ( t ) ) ,
where X ( t ) is a Brownian motion trajectory, E x is the average value of the exponential term with respect to these trajectories, and f is the initial value of ψ , the latter being the sought-after solution of the above Cauchy problem. This classical representation of the time-dependent solution to the Schrödinger equation involves a Wiener measure [9] (i.e., the probability measure on the space of continuous functions) and, unlike the ordinary path integration, provides a rigorous mathematical justification. The above representation is used to calculate the energies and any correlation properties associated with any particular solution.
It is straightforward to implement numerically and does not require a trial function. This method was first applied to calculate energy eigenvalues of simple systems by Donsker and Kac [10] (see also [11]) and was eventually extended to atomic systems by Korzeniowski et al. [12].
It is also known that the classical FK formula provides a rather slow rate of convergence due to the fact that the underlying diffusion process—Brownian motion (Wiener process)—is non-recurrent. Specifically, in dimensions higher than two, the trajectories of the process escape to infinity with the probability one [13]. Mathematically speaking, for d 3 , the probability P ( lim t R t = ) = 1 (where R t is the distance of the Brownian particle from the origin) and lim t P [ X ( t ) B ] = 0 . Here B is a Borel set, i.e., the set of all Brownian processes X ( t ) . As a result, the sampling within the quantum-mechanical region of intersection occurs only during a small fraction of the total simulation time and the rate of convergence becomes prohibitively slow.
Another path integral method, known as the generalized Feynman-Kac (GFK) method, was initiated by Soto-Equibar and Claverie [14] and was subsequently extended to the full GFK method by Caffarel and Claverie [15]. These procedures can be considered an application of importance sampling to the FK integral, along with the transformation of Equation (1) into a Wiener path integral over a distribution determined by both the diffusion and drift terms. This transforms the difficult-to-handle branching (potential energy) term into a more manageable path integral. The GFK method is mathematically more convenient because the limiting distribution exists, i.e., lim t P [ X ( t ) B ] = B ϕ T 2 ( x ) d x .
In this paper, the GFK method is used to calculate the two-body correlation functions as a measure of the quantum fluctuation of the soliton-soliton relative velocity in a GP breather. It is found that the numerical estimate for the quantum fluctuations compares favorably with the preliminary theoretical estimates and is consistent with the Bogoliubov prediction [5].

2. The Model

For calculating the two-body correlation function, let us first consider the ground state of a solitonic system consisting of bosonic atoms (for example, 7 Li ) with a negative scattering length. We assume that a 2-soliton breather is created at t = 0 by quenching the interaction strength of the initial Hamiltonian.
We would like to perform a time propagation of the time-dependent Schrödinger equation:
i t | ψ ( t ) = H 0 ^ | ψ ( t ) ,
with the initial condition given by the ground state of the initial Hamiltonian
H ^ 0 = 1 2 Δ + V 0 ( x ) ,
where
Δ = i = 1 N 2 x i 2
V 0 = V int , 0 + V trap V int , 0 = g ˜ 0 i < j δ σ ˜ ( x i x j ) ,
V trap = 1 2 i x i 2 .
Here and elsewhere below, the system of units, = m = ω = 1 , is used (unlike otherwise given), where is the reduced Planck constant, m is the atomic mass, ω is the frequency of the harmonic confinement. δ σ ˜ ( x i x j ) is the Dirac delta function, σ ˜ denotes the width of the Gaussian potential, and
g ˜ 0 | g 0 | m 3 2 ω
is the dimensionless form of the absolute value of the initial coupling constant, g 0 , and N is the number of bosonic particles. We assume that a 2-soliton breather is created at t = 0 . For that to happen, the pre- and post-quench values of the coupling constant have to satisfy the relation
g 0 = 1 4 g .
To estimate the quantum fluctuations of the relative velocity of the constituent solitons in a GP breather, a path integral approach [16,17,18] is adopted based on the FK integral formalism. To find the path integral solution, let us first consider the Cauchy problem related to the time-dependent Schrödinger equation (3):
i ψ ( x , t ) t = Δ 2 + V ψ ( x , t ) , ψ ( x , 0 ) = f ( x ) ,
with a Hamiltonian,
H = Δ / 2 + V ( x ) .
The solution of the above equation in the FK representation reads:
ψ ( x , t ) = E x e 0 t V ( X ( s ) ) d s f ( X ( t ) ) ,
where E x is the expectation value of the random variables. Numerical implementation of Equation (8) is given in Appendix A.
As was mentioned in the introduction, even though the FK formalism provides a basis for rigorous and accurate calculations of ground-state and excited-state properties of many-particle systems, it suffers from a slow convergence rate as soon as the underlying diffusion process—Brownian motion (Wiener process)—is non-recurrent. To speed up the convergence, one needs to apply the GFK formalism, as described below.
The GFK formalism employs an Ornstein-Uhlenbeck process Y ( t ) , which has a stationary distribution and the convergence becomes much faster. The solution in the FK representation holds for any potential V which belongs to the Kato class [19]. All the ordinary potentials fall under this category.
One can obtain the GFK formalism from the raw FK representation by allowing a large class of diffusions that, unlike Brownian motion, have stationary distributions. Specifically, for any twice-differentiable positive ϕ ( x ) , one defines a new potential U as a perturbation of the potential V:
U ( x ) = V ( x ) 1 2 Δ ϕ ( x ) ϕ ( x ) .
Then:
w ( x , t ) t = 1 2 Δ w ( x , t ) + ϕ ( x ) ϕ ( x ) w ( x , t ) U ( x ) w ( x , t ) = L w ( x , t ) , w ( x , 0 ) = h ( x ) ,
where h is the initial value of w. Equation (10) has the solution
w ( x , t ) = E x e 0 t U ( Y ( s ) ) d s h ( Y ( t ) ) ,
where the new diffusion Y ( t ) has an infinitesimal generator, A = Δ 2 + ϕ ϕ with adjoint, A ( · ) = Δ 2 ϕ ϕ ( · ) . Here, ϕ 2 ( x ) is a stationary density of Y ( t ) , or equivalently, A ( ϕ 2 ) = 0 .
To see the connection between w ( x , t ) and ψ ( x , t ) , observe that for f = 1 and h = 1 ,
w ( x , t ) = ψ ( x , t ) ϕ ( x )
because w ( x , t ) satisfies Equation (10). The diffusion Y ( t ) solves the following stochastic differential equation, d Y ( t ) = ϕ ( Y ( t ) ) / ϕ ( Y ( t ) ) + d X ( t ) .
For numerical calculations, the Gaussian representation for the delta-function potential,
δ σ ˜ ( x i x j ) = 1 2 π σ ˜ exp ( x i x j ) 2 2 σ ˜ 2 ,
is used. The justification for choosing the negative Gaussian potential is given in Appendix B.
For the purpose of the path-integral Monte Carlo, the system is represented by a d-dimensional particle, with d = N , subject to the boundary conditions,
ψ x = ± = 0 , ψ x x = ± = 0 .
The raw FK equation [7] provides a Cauchy-type solution and a guided random walk is adopted here using a trial function that satisfies the required boundary conditions. Using Equation (10) and the GFK path integral representation [15], the solution for Equation (3) can be represented as [16]
ψ ( x , t ) = w ( x , t ) ϕ ( x ) = ϕ ( x ) E x e U ( Y ( s ) ) d s .
Equation (14) contains the two sums: the modified potential U ( Y ( s ) ) is summed over all the steps in a given trajectory, and then, exp [ U ( Y ( s ) ] d s is summed over all the trajectories. Here, in this computational problem, a soliton can be viewed as a bound state of 7 Li atoms interacting through a strong attractive potential, described in Equation (13). The nonnegative function ϕ ( x ) can be chosen as a trial function consistent with the symmetry of the problem. In the present case, the following function is used for ϕ ( x ) and is denoted as ϕ 0 ( x ) :
ϕ 0 ( x ) = C e b x 2 ,
where C is a normalization constant and b is a variational parameter. Let us now introduce a new perturbed potential,
V p ( x ) = e 0 U ( x ) = e 0 V ( x ) 1 2 Δ ϕ 0 ( x ) ϕ 0 ( x ) ,
in terms of U ( x ) and the energy e 0 , associated with the trial function ϕ 0 .
The Ornstein-Uhlenbeck process Y ( t ) is related to the Brownian process X ( t ) as follows:
d Y ( t ) = ϕ ( Y ( t ) ) / ϕ ( Y ( t ) ) + d X ( t ) .
In Equation (17), the first and second terms represent the drift and diffusion, respectively, and the presence of these terms enables the trajectory Y ( t ) to be highly localized. As a result, the important regions of the potential are frequently sampled and Equation (14) converges rapidly. Similarly, the expectation value of the operator A is given by [15]
Y | A | Y = lim t d Y ( t ) A ( Y ( t ) ) e V p ( Y ( s ) ) d s d Y ( t ) e V p ( Y ( s ) ) d s .
Equation (18) is the key formula used here to calculate the quantum fluctuations of the soliton-soliton velocity.
The goal is to compute the second moment of the two-body relative distance at t = T / 4 :
ψ ( t = T / 4 ) | ( x 1 x 2 ) 2 | ψ ( t = T / 4 ) .
The reason we are interested in is that this particular correlator is sensitive to the quantum fluctuations of the relative soliton-soliton velocity being itself a macroscopic variable. The large-N analytic predictions for the velocity fluctuations is [5]:
ψ ( t = 0 ) | V rel . 2 | ψ ( t = 0 ) = 0.0429 g ˜ 2 N ,
where the numerical prefactor comes from a numerically computed integral. This expression neglects the zero-point quantum fluctuations that are induced by the harmonic confinement. In turn, the variance of the relative distance between the centers of mass of the two solitons, after a quarter of a period, is given by
ψ ( t = T / 4 ) | X rel . 2 | ψ ( t = T / 4 ) = ψ ( t = 0 ) | V rel . 2 | ψ ( t = 0 ) .
The mean occupations of the two constituent solitons are N / 4 and 3 N / 4 . Therefore, the probability that two detected particles 1 and 2 of the two different solitons is 6 / 16 . Furthermore, assume that at T / 4 , the distance between the solitons exceeds their width, and therefore the 1 to 2 distance will be dominated by the distance between the centers of mass of the two solitons. Since ( 6 / 16 ) × 0.0429 = 0.0161 , one obtains:
ψ ( t = T / 4 ) | ( x 1 x 2 ) 2 | ψ ( t = T / 4 ) 0.0161 g ˜ 2 N .
As mentioned above, this estimate assumes that (a) the interaction-induced fluctuations in the relative velocity of the solitons exceed those generated by the zero-point fluctuations of the trap, and (b) the soliton-soliton separation at the quarter of the period exceeds the size of the initial density distribution. Let us check the validity of these assumptions.
For (a), the separation (19) is compared with the zero-point fluctuations of the relative distance, ( 8 / 3 ) / N . For (b), the worst-case scenario is considered and it is assumed that the detected atoms 1 and 2 were found on the opposite wings of their respective solitons. That would require that the soliton-soliton distance, given by (19), exceeds the resulting correction to the distance, N / 4 + 3 N / 4 , where N 2 / ( g ˜ N ) is the size of a soliton with N atoms. One gets:
1 ( g ˜ N ) 2 0.06 ,
1 ( g ˜ N ) 2 0.01 N ,
for the conditions (a) and (b), respectively.
The left-hand side of Equation (19) is numerically calculated here using Equation (18), where A = ( x 1 x 2 ) 2 is set. Here, T 2 π is the dimensionless form of the trapping period.

3. Results

The variance in the particle-particle distance that is of interest here is shown in Table 1. The agreement between the theoretical predictions and the numerical results is good, with the exception of the g ˜ = 0.78 case. Meantime, the condition () is barely satisfied, and it is likely that the solitons’ width still contributes to the variance in the interparticle distance.
In order to make a connection with experiments with 7 Li [3], it was assumed that the mass of an individual atom is m = 7.016 u (in unified atomic mass units), the post-quench scattering length, a sc = 16.2 a 0 ( a 0 is the Bohr radius), and the frequency of the radial trapping potential, ω r = 2 π × 297 Hz . The coupling constant is given by g = 2 ω r a sc . This set reproduces the conditions of the Rice experiment [3] verbatim, with the exception of the scattering length a sc . The different value of a sc accounts for the difference between the number of atoms in the experiment and that in the current study. In particular, a sc is adjusted so that the ratio between the number of atoms and the condensate collapse threshold, N c = 0.67 a r / | a sc | [20], is the same as it was in the experiment. The second line of Table 1, g ˜ = 0.55 , would correspond to a one-dimensional trapping frequency of ω = g 2 m / ( g ˜ 2 3 ) = 3.7 × 10 3 Hz with a propagation time of T / 4 = 4.3 × 10 2 s .
The propagation time obtained appears to be much larger than the estimate of 4.7 s in [5], for N = 3 × 10 3 . However, the former corresponds to a conservative estimate, based on a soliton-soliton separation exceeding six half-widths of the broader soliton. For a less conservative requirement, used in [5], the propagation time will be as short as T / 4 = 0.83 s , reached with ω = 1.9 Hz (accordingly, g ˜ = 0.024 ) and other experimental conditions kept intact.

4. Conclusions and Outlook

In the study presented here, it is justified that quantum fluctuations of a macroscopic observable—represented by the relative velocity of two solitons in a harmonically trapped Gross-Pitaevskii breather—can be observed in a predominantly mean-field environment. The scheme involves a harmonic quarter-period propagation of the breather: this time turns out to be sufficient for the breather to dissociate through purely quantum effects. As a computational method, the path integral Monte Carlo is used, the numerical results of which are consistent with the earlier predictions, based on the Bogoliubov approximation [5].
In further studies, stronger interactions to be considered both to move closer to the experimental conditions and to suppress the residual effects of the soliton width. We expect that then Table 1 will exhibit a closer correspondence between the theory predictions and the numerical calculations.

Author Contributions

S.D., V.D. and M.O. equally contributed to the work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the NSF (Grants No. PHY-1912542, and No. PHY-1607221) and the Binational (U.S.-Israel) Science Foundation (Grant No. 2015616). Partial support from the Department of Science and Technology(DST), India (under SERB core grant (award no. EMR/2016/005492) is gratefully acknowledged.

Data Availability Statement

Not acceptable.

Acknowledgments

One of the authors (SD) would like to thank OmDayal Group of Institutions, Engineering College, Uluberia, Howrah 711316, West Bengal, India for providing partial support in the initial phase of the project.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Details of Numerical Calculations

The formalism, described in Section 2, can include any generalized potential [21] and is valid for any arbitrary dimension d ( d = 3 N ). To implement Equation (3) numerically, the 3N-dimensional Brownian motion can be replaced by properly scaled one-dimensional random walks as follows [12,17,22]:
W ( l ) W ( t , n , l ) = w 1 1 ( t , n , l ) , w 2 1 ( t , n , l ) , w 3 1 ( t , n , l ) , , w 1 N ( t , n , l ) , w 2 N ( t , n , l ) , w 3 N ( t , n , l ) ,
where
w j i ( t , n , l ) = k = 1 l ϵ j k i n ,
with w j i ( 0 , n , l ) = 0 for i = 1 , 2 , , N , j = 1 , 2 , 3 , and l = 1 , 2 , , n t . Here, ϵ s denote the binomially distributed random variables, which are chosen independently and randomly with probability P for all i , j , k such that P ( ϵ j k i = 1 ) = P ( ϵ j k i = 1 ) = 1 / 2 , and t is the simulation time. It is known (from an invariance principle [23]) that for every ν and W ( l ) , defined in Equation (A2),
lim n P 1 n l = 1 n t V ( W ( l ) ) = P 0 t V [ X ( s ) ] d s ν .
Consequently, for large n,
P exp 0 t V ( X ( s ) ) d s P exp 1 n l = 1 n t V ( W ( l ) ) ν .
Finally, by generating N rep independent realization Z 1 , Z 2 , , Z N rep of
Z m = exp 1 n l = 1 n t V ( W ( l ) ,
and using the law of large numbers, with regard to Equation (A3), one concludes that
( Z 1 + Z 2 + + Z N rep ) / N rep = Z ( t )
is an approximation of Equation (8). Here, W m ( l ) , m = 1 , 2 , N rep , denotes the m th realization of W ( l ) out of N rep independently run simulations. In the limit of large t and N rep , this approximation approaches an equality and forms the basis of a computational scheme for the lowest energy of a many-particle system with a prescribed symmetry.

Appendix B. Validity of the Gaussian Approximation for the δ-Function

To ensure that the pairwise Gaussian potential correctly describes the intended delta potential, the following has to be checked: (i) the potential supports only one bound state, and (ii) the energy of the bound state is less than the potential depth. To prove that the Gaussian potential with the parameters chosen in the current study supports only one bound state the prescriptions in [24] was folowed. In one dimension, the Wentzel–Kramers–Brillouin (WKB) integral for the energy E is given by x 1 x 2 2 [ E V ( x ) ] d x = ( n 1 2 ) π [25], where x 1 and x 2 are the turning points of the classical motion. For the Gaussian potential, the quantum number N of the last bound state can be obtained by using the above WKB equation for E = 0 : 2 V 0 + e α x 2 / 2 d x = ( N 1 2 ) π , V = V 0 e α x 2 ( α is a parameter related to width of the Gaussian potential), which leads to N = ( 2 / π ) V 0 / α + 1 / 2 . For finite V 0 and α , N is also finite. Therefore, the number of bound states is finite for the Gaussian well.

References

  1. Satsuma, J.; Yajima, N.B. Initial value problems of one-dimensionals self-modulation of nonlinear waves in dispersive media. Prog. Theor. Phys. Supp. 1974, 55, 284–306. [Google Scholar] [CrossRef]
  2. Carli, A.D.; Colquhoun, C.D.; Henderson, G.; Flannigan, S.; Oppo, G.; Daley, A.J.; Kuhr, S.; Haller, E. Excitation modes of bright matter-wave solitons. Phys. Rev. Lett. 2019, 123, 123602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Luo, D.; Jin, Y.; Nguyen, J.H.; Malomed, B.A.; Marchukov, O.V.; Yurovsky, V.A.; Hulet, R.G. Creation and characterization of matter-wave breathers. Phys. Rev. Lett. 2020, 125, 183902. [Google Scholar] [CrossRef] [PubMed]
  4. Yurovsky, V.A.; Malomed, B.A.; Hulet, R.G.; Olshanii, M. Dissociation of one-dimensional matter-wave breathers due to quantum many-body effects. Phys. Rev. Lett. 2017, 119, 220401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Marchukov, O.V.; Malomed, B.A.; Dunjko, V.; Ruhl, J.; Olshanii, M.; Hulet, R.G.; Yurovsky, V.A. Quantum fluctuations of the center of mass and relative parameters of nonlinear schrödinger breathers. Phys. Rev. Lett. 2020, 125, 050405. [Google Scholar] [CrossRef] [PubMed]
  6. Opanchuk, B.; Drummond, P.D. One-dimensional Bose gas dynamics: Breather relaxation. Phys. Rev. A. 2017, 96, 053628. [Google Scholar] [CrossRef] [Green Version]
  7. Feynman, R.P.; Hibbs, A.R. Quantum Mechanics and Path Integrals; McGraw-Hill: New York, USA, 1965. [Google Scholar]
  8. Metropolis, N.; Ulam, S. The Monte Carlo method. J. Am. Stat. Assoc. 1949, 44, 335–341. [Google Scholar] [CrossRef] [PubMed]
  9. Wiener, N. Differential-space. J. Math. Phys. 1923, 2, 131–174. [Google Scholar] [CrossRef]
  10. Donsker, M.D.; Kac, M. A sampling method for determining the lowest eigenvalue and the principal eigenfunction of Schrödinger’s equation. J. Res. Nat. Bur. Stand. 1950, 44, 551–557. [Google Scholar] [CrossRef]
  11. Kac, M. On some connections between probability theory and differential and integral equations. In Proceedings of the Second Berkeley Symposium on Mathematical Statistics and Probability, Berkeley/Los Angeles, CA, USA, 31 July–12 August 1950; Neyman, J., Ed.; University of California Press: Berkeley/Los Angeles, CA, USA, 1951; pp. 189–215. Available online: https://projecteuclid.org/proceedings/berkeley-symposium-on-mathematical-statistics-and-probability/proceedings-of-the-second-berkeley-symposium-on-mathematical-statistics-and-probability/toc/bsmsp/1200500213 (accessed on 1 October 2021).
  12. Korzeniowski, A.; Fry, J.L.; Orr, D.E.; Fazleev, N.G. Feynman-Kac path-integral calculation of the ground-state energies of atoms. Phys. Rev. Lett. 1992, 69, 893–896. [Google Scholar] [CrossRef] [PubMed]
  13. Karatzas, I.; Shreve, S.E. Brownian Motion and Stochastic Calculus; Springer: New York, USA, 1991. [Google Scholar] [CrossRef]
  14. Soto-Equibar, F.; Claverie, P. Stochastic methods in quantum theory: Generalization of the Feynman-Kac formula. In Proceedings of the Latin American School of Physics: Stochastic Processes Applied to Physics and other Related Fields, Cali, Colombia, 21 June–10 July 1982; Gomez, B., Moore, S.M., Rodriguez-Vargas, A.M., Rueda, A., Eds.; World Scientific: Singapore, 1983; pp. 637–639. [Google Scholar]
  15. Caffarel, M.; Claverie, P. Development of a pure diffusion quantum Monte Carlo method using a full generalized Feynman-Kac formula. I. Formalism. J. Chem. Phys. 1988, 88, 1088–1099. [Google Scholar] [CrossRef]
  16. Korzeniowski, A. On computer simulation of Feynman-Kac path-integrals. J. Comp. App. Math. 1996, 66, 333–336. [Google Scholar] [CrossRef] [Green Version]
  17. Datta, S.; Fry, J.L.; Fazleev, N.G.; Alexander, S.A.; Coldwell, R.L. Feynman-Kac path-integral calculations with high-quality trial wave functions. Phys. Rev. A 2000, 61, 030502(R). [Google Scholar] [CrossRef]
  18. Datta, S. Importance Sampling in the Feynman-Kac Path Integral Approach to Quantum Mechanics. Ph.D. Thesis, The University of Texas at Arlington, Arlington, TX, USA, 1996. [Google Scholar]
  19. Kato, T. On the eigenfunctions of many-particle systems in quantum mechanics. Commun. Pure Appl. Math. 1957, 10, 151–157. [Google Scholar] [CrossRef]
  20. Gammal, A.; Frederico, T.; Tomio, L. Critical number of atoms for attractive Bose-Einstein condensates with cylindrically symmetrical traps. Phys. Rev. A. 2001, 64, 055602. [Google Scholar] [CrossRef] [Green Version]
  21. Simon, B. (Ed.) Functional Integration and Quantum Physics; Academic Press: New York, USA, 1979; Available online: https://www.sciencedirect.com/bookseries/pure-and-applied-mathematics/vol/86/suppl/C (accessed on 1 October 2021).
  22. Griego, R.; Hersh, R. Theory of random evolutions with applications to partial differential equations, Transact. Am. Math. Soc. 1971, 156, 405–418. [Google Scholar] [CrossRef]
  23. Billingsley, P. Convergence of Probability Measures; Wiley: New York, USA, 1968. [Google Scholar]
  24. Nandi, S. The quantum Gaussian well. Am. J. Phys. 2010, 78, 1341–1345. [Google Scholar] [CrossRef] [Green Version]
  25. Merzbacher, E. Quantum Mechanics; John Wiley & Sons: New York, NY, USA, 2005. [Google Scholar]
Table 1. Quantum fluctuations of the relative velocity of constituent solitons obtained from numerical calculations and from theory. N is the number of atoms, “scale” represents ( number of steps ) / t with t being the simulation time, “npi” stays for the number of trajectories, g ˜ and σ ˜ denote the dimensionless forms of the interaction strength and the width of the Gaussian potential, respectively. See text for details.
Table 1. Quantum fluctuations of the relative velocity of constituent solitons obtained from numerical calculations and from theory. N is the number of atoms, “scale” represents ( number of steps ) / t with t being the simulation time, “npi” stays for the number of trajectories, g ˜ and σ ˜ denote the dimensionless forms of the interaction strength and the width of the Gaussian potential, respectively. See text for details.
NScalenpi g ˜ σ ˜ ( x i x j ) 2 (Numerics) ( x i x j ) 2 (Theory)
10030500.50.0160.5884 ± 0.1647 0.4025
0.550.0150.5309 ± 0.1486 0.487
0.610.0150.6209 ± 0.1738 0.599
0.780.0120.4389 ± 0.1229 0.9795
0.830.011.6773 ± 0.4696 1.1091
0.850.011.8155 ± 0.5083 1.632
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Datta, S.; Dunjko, V.; Olshanii, M. Path Integral Estimates of the Quantum Fluctuations of the Relative Soliton-Soliton Velocity in a Gross-Pitaevskii Breather. Physics 2022, 4, 12-20. https://doi.org/10.3390/physics4010002

AMA Style

Datta S, Dunjko V, Olshanii M. Path Integral Estimates of the Quantum Fluctuations of the Relative Soliton-Soliton Velocity in a Gross-Pitaevskii Breather. Physics. 2022; 4(1):12-20. https://doi.org/10.3390/physics4010002

Chicago/Turabian Style

Datta, Sumita, Vanja Dunjko, and Maxim Olshanii. 2022. "Path Integral Estimates of the Quantum Fluctuations of the Relative Soliton-Soliton Velocity in a Gross-Pitaevskii Breather" Physics 4, no. 1: 12-20. https://doi.org/10.3390/physics4010002

Article Metrics

Back to TopTop