This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Interfaces between Bose–Einstein and Tonks–Girardeau atomic gases

and

Published 3 February 2016 © 2016 IOP Publishing Ltd and Deutsche Physikalische Gesellschaft
, , Focus on Strongly interacting quantum gases in one dimension Citation Giovanni Filatrella and Boris A Malomed 2016 New J. Phys. 18 025005 DOI 10.1088/1367-2630/18/2/025005

1367-2630/18/2/025005

Abstract

We consider one-dimensional mixtures of an atomic Bose–Einstein condensate (BEC) and Tonks–Girardeau (TG) gas. The mixture is modeled by a coupled system of the Gross–Pitaevskii equation for the BEC and the quintic nonlinear Schrödinger equation for the TG component. An immiscibility condition for the binary system is derived in a general form. Under this condition, three types of BEC–TG interfaces are considered: domain walls (DWs) separating the two components; bubble-drops (BDs), in the form of a drop of one component immersed into the other (BDs may be considered as bound states of two DWs); and bound states of bright and dark solitons (BDSs). The same model applies to the copropagation of two optical waves in a colloidal medium. The results are obtained by means of systematic numerical analysis, in combination with analytical Thomas–Fermi approximations (TFAs). Using both methods, families of DW states are produced in a generic form. BD complexes exist solely in the form of a TG drop embedded into the BEC background. On the contrary, BDSs exist as bound states of TG bright and BEC dark components, and vice versa.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Binary systems, whose behavior crucially depends on the underlying condition of immiscibility or miscibility [1], play a fundamentally important role in many areas of physics. In the case of immiscibility, a major effect is the formation of domain walls (DWs) between regions occupied by immiscible components. Commonly known are DWs in media featuring a vectorial order parameter, such as ferromagnets [2], ferroelectrics [3], and liquid crystals [4]. In self-defocusing optical media, DWs separate regions occupied by electromagnetic waves with orthogonal circular polarizations of light [5, 6]. Similar interface patterns were predicted in arrays of nonlinear optical waveguides, modeled by discrete nonlinear Schrödinger equations (NLSEs) [8].

DWs are known in superfluids too, where they are formed by immiscible binary Bose–Einstein condensates (BECs), as predicted theoretically [9] and demonstrated in experiments [10]. In the mean-field approximation [11], such settings are modeled by systems of nonlinearly coupled Gross–Pitaevskii equations (GPEs) with the cubic self-repulsive nonlinearity, which are similar to coupled NLSEs describing the above-mentioned optical DWs [5, 6]. In their stationary form, these equations coincide with coupled cubic Ginzburg–Landau equations modeling DWs in dissipative patterns, such as interfaces between rolls with different orientations in large-area Rayleigh–Benard convection [7].

The analysis of the DWs in BEC was extended for broader settings, including linear interconversion between the immiscible components (this is possible when they represent two different hyperfine states of the same atom coupled by a resonant radiofrequency wave) [12], dipolar [13] and spinor (three-component) condensates [14], as well as the BEC discretized by trapping in a deep optical-lattice (OL) potentials [15]. Furthermore, the study of the DWs was recently extended for immiscible binary BECs with three-particle collisions [16], in the case when the related losses may be neglected, the respective coupled GPEs featuring the cubic–quintic repulsive nonlinearity [17].

In the effectively one-dimensional (1D) setting, ultracold bosonic gases with strong inter-atomic repulsion may be cast in the Tonks–Girardeau (TG) state, which emulates the gas of noninteracting fermions [18], provided that the energy of the repulsive interaction between bosons exceeds their kinetic energy, while the opposite situation corresponds to the BEC phase in the bosonic gas (a review of the TG model was given in [21]). The TG gas of hard-core bosons has been realized experimentally, using tight transverse confinement [19, 20]. In particular, a longitudinal OL potential was used to increase the effective mass in the trapped state, thus making the kinetic energy small enough [19].

It is commonly known that GPEs furnish very accurate description of the BEC in atomic gases. A similar macroscopic model of the TG gas is offered by the NLSE with the quintic self-repulsion term [22]. In a rigorous form, the relevance of the corresponding sextic term in the free-energy density of the three-dimensional bosonic gas in its ground state, which reduces to the quasi-1D TG phase, was demonstrated in [23], under condition ${ \mathcal G }{ \mathcal L }\gg { \mathcal N }$, where ${ \mathcal G }$, ${ \mathcal L }$, and ${ \mathcal N }$ are, respectively, the inter-atomic repulsion strength, system's length, and the total number o atoms. The quintic model was used in various contexts, including shock waves [24], dark [25] and gap-mode [26] solitons, as well as bright solitons supported by dipole–dipole interactions [27], and, recently, DWs in immiscible binary TG gases [16]. Further, oscillation frequencies derived from fermionic hydrodynamic equations, which apply to the hard-core TG gas, were found to be close to their counterparts predicted by the quintic NLSE [28]. Coupled quintic NLSEs also arise in works aimed at constructing the ground state of a binary TG mixture in the harmonic-oscillator potential by means of the density-functional method [29]. On the other hand, this approach may not apply to TG gases beyond the framework of static configurations and hydrodynamic regimes. In particular, it fails for strongly nonequilibrium problems, such as merger of distinct gas clouds [30].

As concerns the mixtures, it may be interesting to consider binary systems including the TG gas and another quantum-gas component. In particular, exact solutions were found for the ground state of TG–Fermi mixtures [31]. The binary gas of impenetrable bosons is solvable too [32]. The objective of the present work is to introduce basic nonlinear complexes, such as DWs, bubble-drop (BD) modes (bound pairs of two DWs), and dark–bright solitons (DBSs), in an immiscible system of TG and BEC gases. In the experiment, the system may be realized, in particular, as a bosonic gas composed of two atomic species under tight transverse confinement, with a longitudinal OL potential acing on (being relatively close to a resonance with) one species only. Then, as the experimental setting presented in [19] suggests, a large effective mass of the near-resonant component will bring it into the TG state, while the other component may stay in the BEC phase.

As a model for this system, in section 2 we adopt the cubic GPE for the self-repulsive BEC component coupled by the cubic (collisional) repulsive term to the quintic NLSE for the TG species. The use of the latter equation is appropriate, as we study only static configurations of the system. The same model may find a realization in optics as a model of colloidal waveguides. It has been recently demonstrated that, selecting the size of metallic nanoparticles in the colloid and their concentration, one can engineer desirable coefficients of the corresponding cubic and quintic nonlinearity [33]. In particular, it is possible to design a waveguide which features a nearly pure quintic nonlinearity at a particular wavelength, while the cubic response dominates at a different wavelength. The copropagation of optical signals carried by these wavelengths will then emulate the TG-BEC system.

DWs separating the BEC and TG phases are addressed in section 3. We derive the respective immiscibility condition (see equation (19) below), and then generate DW states in a systematic way, using both numerical solutions and the analytical Thomas–Fermi approximation (TFA). The latter method makes it possible to obtain some DWs in an explicit analytical form, as given below by equations (24)–(26). DB and DBS complexes are considered in sections 4 and 5, respectively, again using a combination of analytical and numerical methods. The analysis predicts that the DB states exist solely in the form of the TG drop embedded into the BEC background (bubble), but not in the opposite case; on the other hand, the DBS are predicted in either case of the bright TG soliton embedded into the BEC dark soliton, or vice versa. These predictions are fully corroborated by numerical results. The paper is concluded by section 6.

2. The model

The system of coupled NLSEs with the cubic and quintic nonlinear terms for the BEC and TG components, ${\psi }_{1}$ and ${\psi }_{2}$, with respective scaled masses m1 and

Equation (1)

(i.e., m is the relative effective mass of the TG component which, as said above, may be made larger than the actual atomic mass [19]) is

Equation (2)

where real parameters $\gamma \geqslant 0$ and ${\rm{\Gamma }}\gt 0$ are strengths of the self-repulsion of the BEC component, and repulsion between the BEC and TG ones, respectively, while ${\hslash }$ and the coefficient of the effective quintic self-repulsion of the TG component are scaled to be 1 (in the notation of [22], the natural value of the latter coefficient is ${\pi }^{2}$). Substitution

Equation (3)

makes it possible to further fix ${\rm{\Gamma }}\equiv 1$ and ${m}_{1}\equiv 1$, thus simplifying equation (2) to a system with two free coefficients, m and g:

Equation (4)

Equation (5)

For the above-mentioned spatial-domain optical model (with t replaced by the propagation distance, Z, and transverse coordinate X), the scaled propagation equations are derived, using the standard procedure [37], as

Equation (6)

Here ${{\rm{\Psi }}}_{1}$ and ${{\rm{\Psi }}}_{2}$ represent amplitudes of the copropagating electromagnetic waves with relative wavenumber and frequency $k={k}_{2}/{k}_{1}$ and $\omega ={\omega }_{2}/{\omega }_{1}$, see equation (1). Further, ${G}_{\mathrm{3,5}}$ in equation (6) are the cubic and quintic SPM coefficients for the two waves, while the cubic XPM coefficient is normalized to be 1. Additional rescaling,

transforms equation (6) into the system of equations (4) and (5), with $m\equiv k$.

The Hamiltonian corresponding to equations (4) and (5) is

Equation (7)

In addition to H, the system preserves the norms (scaled numbers of atoms in the ultracold gas, or total powers of the two waves, in terms of the optical model)

Equation (8)

of the two components, and, for dynamical solutions, also the total momentum, $P={\rm{i}}{\displaystyle \int }_{-\infty }^{+\infty }{\sum }_{n=1}^{2}{\psi }_{n}(\partial {\psi }_{n}^{*}/\partial x){\rm{d}}x$, although, as mentioned above, the use of the quintic NLS equation for the description of dynamics of the TG gas may be impugnable.

Stationary solutions to equations (4) and (5) with positive chemical potentials ${\mu }_{\mathrm{1,2}}$ are looked for as

Equation (9)

with real functions ${\phi }_{\mathrm{1,2}}(x)$ satisfying equations

Equation (10)

Equation (11)

(in the optical model, $-{\mu }_{\mathrm{1,2}}$ represent the propagation constants of the two waves). Equations (10) and (11), if considered as equations of the evolution along x, conserve the formal Hamiltonian

Equation (12)

see equation (7).

Normalizing the stationary wave functions as

Equation (13)

one may fix ${\mu }_{2}=1$, which implies that the density of the uniform TG component is also fixed to be 1, see equation (14) below. Numerical results for DWs are presented in the following section chiefly for this case, while in sections 4 and 5 other normalizations are used for BD and DBS states.

3. Domain walls

3.1. Analytical considerations

Solutions of stationary equations (10) and (11) for the DW separating semi-infinite domains occupied by the BEC and TG components (or the spatial domains occupied by the copropagating waves in the above-mentioned optics model) are specified by the following boundary conditions (b.c.), which include the respective asymptotic densities, ${\phi }_{1}^{2}(x=-\infty )\equiv {({n}_{1})}_{\mathrm{asympt}}$ and ${\phi }_{2}^{2}(x\;=\;+\infty )\equiv {({n}_{2})}_{\mathrm{asympt}}$:

Equation (14)

The condition that formal Hamiltonian (12) must take the same values at $x\to -\infty $ and $x\to +\infty $ across the solution imposes a restriction on b.c. (14), which relates the two chemical potentials:

Equation (15)

In terms of the asymptotic densities, condition (14) takes the form of

Equation (16)

which actually implies the balance of the pressure applied to the DW from the two sides. Naturally, the mass ratio (m) does not appear in equations (15) and (16). Note that, if condition (16) between the densities does not hold initially, the DW in a finite system, which corresponds to real experimental settings, will move to a position at which the condition holds for the accordingly modified densities.

Further, the immiscibility condition for the BEC and TG, which is necessary for the existence of the DW separating the two quantum gases, is that the demixed configuration must provide a smaller energy density (defined as per equation (7)) than a uniformly mixed state with densities which are equal to half of asymptotic densities (14):

Equation (17)

The uniform state corresponds to values of the chemical potentials different from ${\mu }_{1}$ and ${\mu }_{2}$, namely, ${({\mu }_{1})}_{\mathrm{uni}}=g{({n}_{1})}_{\mathrm{uni}}+{({n}_{2})}_{\mathrm{uni}}$ , ${({\mu }_{2})}_{\mathrm{uni}}={({n}_{2})}_{\mathrm{uni}}^{2}+{({n}_{2})}_{\mathrm{uni}}$. Then, the comparison of the average energy densities of the demixed and uniform states gives rise to the immiscibility condition in the following form:

Equation (18)

Further, the substitution of relation (15) in equation (18) transforms it into an inequality for the self-repulsion strength of the BEC component, g:

Equation (19)

In particular, for ${\mu }_{2}=1$ (as said above, this value will be fixed by rescaling), equation (19) yields ${g}_{\mathrm{max}}\quad =\quad 48/49\approx 0.9796$.

The DW may be characterized by an effective width of its core, which may be naturally defined by the following integral expression:

Equation (20)

The analysis should produce W as a function of parameters m and g, once the TG asymptotic density is fixed by setting ${\mu }_{2}=1$, see figure 3 below.

Figure 1.

Figure 1. DW density profiles produced by the imaginary-time solution of equations (4), (5), and subsequently verified by the real-time propagation, for ${\mu }_{2}=1$, ${\mu }_{1}=0.258$, g = 0.05, and two values of the relative mass, m = 0.1 and 10. In terms of the optics model, the BEC and TG densities correspond to power densities of the waves subject to the action of the cubic and quintic self-defocusing nonlinearity, respectively (in captions to the following figures, they are referred to as 'cubic' and 'quintic' components, respectively).

Standard image High-resolution image
Figure 2.

Figure 2. The DW density profile, produced by the numerical solution of equations (4) and (5), and its comparison with the analytical prediction (TFA), given by equations (22), (24) and (25), for large g (and/or small m), viz., g = 5, m = 0.152.

Standard image High-resolution image
Figure 3.

Figure 3. (a) The DW width, defined as per equation (20 ) versus the TG/BEC (quintic/cubic) mass ratio, m, at different fixed values of the interaction constant, g. The dashed black line shows the analytical prediction (26) produced by the TFA for $g=3/4$. (b) The DW width versus g, at several fixed values of m. In both panels, ${\mu }_{2}=1$ is fixed.

Standard image High-resolution image

Close to the existence boundary of the DW, i.e., at $0\lt {g}_{\mathrm{max}}-g\ll {g}_{\mathrm{max}}$, the DW becomes very wide. In this limit, the dependence between the width and proximity to the threshold may be estimated as follows: the density of the gradient energy in equation (7) scales as $1/{W}^{2}$, hence the full gradient energy scales as $1/W$, and the respective effective force may be estimated as $-\partial H/\partial W$ $\sim 1/{W}^{2}$. It must be balanced by the effective bulk force vanishing at $g={g}_{\mathrm{max}}$, which scales as ${g}_{\mathrm{max}}-g$. Thus, the equilibrium condition predicts that the DW's width diverges at ${g}_{\mathrm{max}}-g\to 0$ as

Equation (21)

Accurate analytical results for the DW can be obtained in the limit case corresponding to $g\gg 1$, or to $m\ll 1$, when the derivative term in equation (10) may be neglected (which actually implies the use of the TFA [11]), yielding

Equation (22)

Strictly speaking, the assumption of $m\ll 1$ may contradict the experimentally relevant way of the realization of the TG gas, based on making the respective effective mass large [19]. Nevertheless, the other option justifying the applicability of the TFA, $g\gg 1$, is quite relevant.

The substitution of expressions (22) and (15) in equation (11) leads to the single stationary equation for ${\phi }_{2}$, with the cubic–quintic nonlinearity:

Equation (23)

(Equation (15) was used here to eliminate ${\mu }_{1}$.) Then, using a known particular exact solution for the DW solution of equation (23) [38], we obtain a solution at the following values of the chemical potentials:

Equation (24)

i.e., ${({n}_{1})}_{\mathrm{asympt}}=3/(4{g}^{2})$, ${({n}_{2})}_{\mathrm{asympt}}=3/(4g)$, in the form of

Equation (25)

Note that ${({\mu }_{2})}_{\mathrm{DW}}$ satisfies condition (19), and rescaling (13) transforms values (24) into ${({\mu }_{2})}_{\mathrm{DW}}={({\mu }_{1})}_{\mathrm{DW}}=1$. The width of this solution, defined according to equation (20), is

Equation (26)

Comparison of analytical solution (25) with its numerical counterpart is displayed in figure 2.

An analytical approach can also be developed in the opposite limit of $m\to \infty $ (very heavy TG atoms), when the TFA may be applied to equation (11) (i.e., the derivative term may be dropped in this equation), reducing it to

Equation (27)

In this case, equation (10) takes the form of

Equation (28)

see equations (22) and (23) derived above in the opposite limit of $m\to 0$. In the particular case of g = 0 (no intrinsic interaction in the BEC component), equation (28) coincides with the equation considered in [39], in the context of a completely different physical model, for a resonantly absorbing Bragg reflector in optics. The formal Hamiltonian for equation (28) is

Equation (29)

see equation (12). The DW solution corresponds to a solution of equation (28) with the b.c. produced by equation (14): ${\phi }_{1}^{2}(x=-\infty )={\mu }_{1}/g,\quad {\phi }_{1}(x\quad =\quad +\infty )=0$, and, simultaneously, ${\phi }_{1}^{2}(x=-\infty )={\mu }_{2}$, the latter relation following from equation (27). Obviously, this b.c. set is self-consistent solely for ${\mu }_{1}=g{\mu }_{2}$. In the combination with this relation, equating values of h corresponding to $x=-\infty $ and $x\;=\;+\infty $, as per equation (29), yields $g\sqrt{{\mu }_{2}}=-4/3$, which is impossible, as g cannot be negative. Thus, the DW solution does not exist in this approximation. Nevertheless, it readily produces analytical solutions for complexes of other types, BD and DBS, as shown below.

3.2. Numerical findings

Generic DW structures obtained from numerical solutions of equations (4) and (5) are displayed in figure 1. Stationary solutions have been obtained by means of the imaginary-time-propagation method. The validity and stability of the solutions was then checked by simulations in real time. The simulations in both imaginary and real time were run by means of the split-step Crank–Nicholson method [40].

Stable solutions have been found in a wide range of values of m at $g\lt 1$, and no solutions could be obtained at $g\gt 1$, which is readily explained by equation (19) (recall we set ${\mu }_{2}=1$). Actual numerical results for the DWs have been obtained for $g\leqslant 0.85$. In other words, a natural conclusion is that the DW exists if the BEC–TG repulsion is stronger than the intrinsic repulsion of the BEC component. In the remaining interval of $0.85\lt g\lt 48/49$ (see equation (19)), it is difficult to generate DWs numerically, as the initial guess for finding the solution by means of the imaginary-time propagation should be very close to the true one, in case the solution is sought for close to its existence boundary. Unlike g, the dependence of the DW solutions on m is very weak, as seen in figure 1 which displays examples for m = 0.1 and m = 10 (the comparison of the results obtained for large and small values of m, presented here and below, is appropriate even if very small values of m may be experimentally irrelevant, as mentioned above).

In addition, figure 2 displays the comparison of the analytical approximation (TFA) based on equations (22), (24), and (25) with the numerical solution obtained for the same values of parameters. It is observed that the analytically predicted density profiles are virtually identical to their numerical counterparts.

To characterize the entire family of the DW solutions, their width defined as per equation (20) has been numerically evaluated, as shown in figure 3(a), versus the mass ratio, m, in the interval of $0.05\leqslant m\leqslant 50$ . It is seen that the dependence of the width on m is rather weak, in agreement with examples displayed in figure 1. Further, the dependence of the DW's width on the BEC self-repulsion constant, g, is shown in figure 3(b). The range of values of g displayed in the figure is bounded by existence limits of the DW solutions, as given by equation (19). The increase of the width with g is a natural property of the system, as stronger self-repulsion stretches the transient layer (in the analytical form, this is clearly shown by equation (21) and solution (25).)

4. BD states

4.1. Analytical considerations

BD solutions are even ones, with ${\phi }_{\mathrm{1,2}}(-x)={\phi }_{\mathrm{1,2}}(x)$, singled out by b.c.

Equation (30)

(a BEC drop embedded into the TG background), or

Equation (31)

(a TG drop embedded into the BEC background), see equation (14). In fact, the BDs may be considered as bound pairs of the DWs, with the layer BEC or TG trapped between the two semi-infinite TG or BEC domains, respectively.

The existence and stability of these solutions has been numerically checked by fixing the chemical potential of the background to unity, and varying the norm of the drop, or more precisely:

  • setting ${\mu }_{2}=1$ for the TG background, see equation (30), and selecting several fixed values of N1 for the BEC drop, see equation (8);
  • setting ${\mu }_{1}=1$ for the BEC background, see equation (31), and selecting several fixed values of N2 for the TG drop, see equation (8).

It is relevant to stress that, unlike the DWs, the chemical potentials of the bubble and drop components are not related by any condition similar to equation (15). In this setting, the remaining free parameters are relative mass m and BEC self-repulsion coefficient g.

Analytical results for the BD structures are available in the limit case of $m\to \infty $ (heavy TG atoms), which corresponds to equations (27) and (28). Indeed, in this approximation the BD solutions, defined by b.c. (30) or (31) correspond, respectively, to bright solitons or bubbles produced by equation (28) ('bubbles' are solutions of NLSEs in the form of a local drop of the density supported by the flat background, without zero crossing, and without a phase shift at $x\to \pm \infty $, unlike dark solitons [34]). Further, the analysis of the corresponding Hamiltonian (29) demonstrates that equation (28) cannot generate bright solitons, but it readily gives rise to bubbles, i.e., the BD patterns in the form of the TG drop embedded into the BEC background. The fact that the BDs exist in the form of the TG drop embedded into the BEC bubble, but do not exist in the reverse form, is confirmed by numerical results presented below.

Analytical results for the BD structures are also available in the limit case of $g\to \infty $ or $m\to 0$, when the derivative term may be neglected in equation (10), leading to equation (22), as shown above. The substitution of this approximation into equation (11) leads to the single equation with the cubic–quintic nonlinearity. For ${\mu }_{1}$ not linked to ${\mu }_{2}$ by relation (15), which was specific for the DW, but is not relevant in the present case, the cubic–quintic equation takes a form slightly more general than equation (23) derived above:

Equation (32)

A straightforward consideration demonstrates that equation (32) gives rise to bubbles, i.e., effectively, the BDs of type (30), in the case of

Equation (33)

and to bright solitons, i.e., the BDs of type (31), in the range of

Equation (34)

In fact, the present limit of $g\to \infty $ or $m\to 0$ implies that interval (33) does not exist, while condition (34) may hold. Thus, this consideration again predicts that the BD exist solely in the form of a TG drop (bright soliton) embedded into the BEC bubble, which is exactly confirmed by numerical results following below.

In terms of the above-mentioned optics model, this conclusion means that a bright soliton driven by the quintic self-defocusing may be embedded into the background filled by the optical wave subject to the cubic self-defocusing, but not vice versa.

4.2. Numerical findings

For the state defined as per equation (30) (a BEC/cubic drop embedded into the TG/quintic background), it was not possible to find solutions in the entire parametric region investigated, which amounts to intervals $0.2\leqslant g\leqslant 5$ and $0.1\leqslant m\leqslant 10$, and $0.1\leqslant {N}_{1}\leqslant 10$. In imaginary-time simulations, this configuration, if taken as an input, transforms into a flat one. The nonexistence of this BD species is readily explained by the analytical results presented above.

On the other hand, also in agreement with the above analysis, in the same parameter region (with N1 replaced by N2) stable solutions have been readily found for the BD state defined by equation (31) (a TG/quintic drop embedded into the BEC/cubic background), see a typical example in figure 4. The BD may be characterized by the missing mass of the bubble void, i.e.

Equation (35)

Computations demonstrate that, for the family of the BD states, ${M}_{\mathrm{void}}$ very weakly depends on relative mass m. Figure 5 displays the missing mass of the BEC component as a function of g. The dependence observed in figure 5 may be approximated by relation ${M}_{\mathrm{void}}\simeq {N}_{2}/g$, which is easy to understand: according to equation (10), the TG component with density ${\phi }_{2}^{2}$ replaces the missing BEC field with effective density ${({\phi }_{1}^{2})}_{\mathrm{void}}={\phi }_{2}^{2}/g$. This argument also explains an approximately linear increase of ${M}_{\mathrm{void}}$ with the growth of the norm of the TG/quintic drop, N2, as clearly seen in figure 5.

Figure 4.

Figure 4. A typical example of a BD state, in the form of the TG (quintic) drop with ${N}_{2}=4$, immersed into the BEC (cubic) background, as produced by the numerical solution of equations (4) and (5) with b.c. given by equation (31). Other parameters are ${\mu }_{1}=1$, g = 1, m = 1, ${\mu }_{2}=0.87$, ${M}_{\mathrm{void}}=4$.

Standard image High-resolution image
Figure 5.

Figure 5. The missing mass in the bubble of the BEC gas (cubic component) ousted by the TG (quintic) drop of norm N2, as defined by equation (35), in the BD state. The BEC chemical potential is ${\mu }_{1}=1$, and the TG/BEC mass ratio is m = 1.

Standard image High-resolution image

5. Dark–bright solitons

5.1. The analytical formulation

Bright–dark soliton (DBS) solutions represent complexes with one even localized (bright) component, and the other delocalized spatially odd zero-crossing one (the dark soliton), the difference from the BD structure being that in the latter case both components were patterned as even ones, without zero crossing in the bubble structure. The corresponding b.c. for the DBS complex are given by

Equation (36)

(a bright BEC component embedded into the TG background), or

Equation (37)

(a bright TG component embedded into the BEC background), see equations (30) and (31). Note that, similar to the case of the BD, the DBS existence condition does not impose any relation on the chemical potentials of its components, unlike equation (15) for the DW.

The limit case of $m\to \infty $ (heavy TG atoms), which is represented by equations (27) and (28), makes it possible to produce analytical results for the DBS defined by b.c. (37), which corresponds to a dark soliton of equation (28). The analysis of the Hamiltonian (29) demonstrates that equation (28) indeed gives rise to dark solitons, i.e., the DBS patterns in the form of the bright TG soliton embedded into the dark BEC soliton.

The consideration of the DBS is also possible in the opposite limit of $g\to \infty $ or $m\to 0$, which amounts to the consideration of equation (32). Dark solitons of that equation correspond to the DBS defined by b.c. (36), and it is easy to see that equation (32) gives rise to the dark solitons in the range of $-\infty \lt m({\mu }_{1}/g-{\mu }_{2})\lt 3/8$. Obviously, this condition holds in the present limit of $g\to \infty $ or $m\to 0$.

Thus, the analysis of the limit cases readily predicts the existence of both species of the DBSs, which correspond to b.c. (36) and (37). This prediction is confirmed by the following numerical results.

5.2. Numerical results

Both types of the DBS have been produced by the numerical computations, as shown in figure 6.

Figure 6.

Figure 6. Generic profiles of BDS solutions. (a) The one with the bright BEC (cubic) and dark TG (quintic) components, as per equation (36). (b) The one with the bright TG and dark BEC components, as per equation (37). Parameters are (a): $m=g={\mu }_{2}=1$, ${N}_{2}=1$, (b): $m=g={\mu }_{1}=1$, ${N}_{1}=1$.

Standard image High-resolution image

The DBS complexes have their existence boundaries. They cease to exist when the self-repulsion BEC coefficient, g, becomes too large. This is shown in figure 7, that displays the boundaries in the plane of $(g,m)$. It is seen that a lower chemical potential of the dark component favors the existence of the DBS, as the boundaries move towards higher values of g with the decrease of the chemical potential. Relative mass m plays a different role: the existence area for the DBS complex with the bright BEC (cubic) component shrinks with the increase of m, while for the case of the bright TG (quintic) component the increase of m favors the DBS existence.

Figure 7.

Figure 7. The BDSs, with the bright BEC (cubic) (a) or TG (quintic) (b) component, exist on the left-hand side of the curves. Panels (a) and (b) are drawn for ${N}_{1}=1$ and ${N}_{2}=1$, respectively.

Standard image High-resolution image

6. Conclusion

We have introduced a binary system in the form of an immiscible pair of BEC and TG quantum gases, in the effectively 1D setting. Using the system of the GPE for the BEC, nonlinearly coupled to the quintic NLSE for the TG component, we have analyzed the possibility of the existence of three types of interfaces in this binary system: DWs, BDs, and bright–dark solitons. The same model applies to a bimodal light propagation in a specially designed colloidal medium in optics. The immiscibility condition for the binary system of the present type was obtained in the general form. Analytical results were produced by means of the TFAs, in the combination with the systematic numerical analysis. The DWs have been found in an explicit analytical form, by means of the TFA, and in the generic form numerically. The analysis and numerical results demonstrate that the BDs exist solely in the form of the TG drop immersed into the BEC background, while the BDS complexes may be built equally well of the TG bright and BEC dark components or vice versa. Thus, the predicted results suggest new experiments with tightly confined two-species mixtures of ultracold bosonic gases. The analysis can be extended for a chain of BDs and/or DBSs in a long system, including a circular one, which corresponds to the binary gas loaded into a tight toroidal trap [35, 36].

Acknowledgments

We appreciate valuable discussions with M Salerno. GF acknowledges financial support from Italian research programs PON Ricerca e Competitivit à 2007–2013, under grant agreement PON NAFASSY, PONa3_00007, and Programma regionale per lo sviluppo innovativo delle filiere Manifatturiere strategiche della Campania — Filiera WISCH Progetto 2: Ricerca di tecnologie innovative digitali per lo sviluppo sistemistico di computer, circuiti elettronici e piattaforme inerziali ad elevate prestazioni ad uso avionico.

Please wait… references are loading.
10.1088/1367-2630/18/2/025005