Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Manipulation of a stable dark soliton train in polariton condensate

Open Access Open Access

Abstract

With a potential trap to mimic the finite-size effect, we predict the generation of dark soliton train in a nonresonant polariton condensate excited with the spatially homogeneous continuous wave (cw) field by exploiting the quantum interference in spatial domain. The number of solitons in the train is demonstrated to be controllable on demand in the domain of material parameters. We show that, similar to dark solitons in conservative systems, these nonlinear excitations have infinite lifetime and remain spatially localized even for the periodically oscillating and colliding state of dark soliton train, which do not depend on the parameters of the condensate. Especially, a final state of anti-dark soliton is observed for all kinds of dark soliton trains with the increase of pump power.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Exciton polaritons [1] in semiconductor microcavity are a breed of bosonic optelectronic excitations mixed by cavity photons and intracavity excitons in the strong-coupling regime. These quasiparticles possess both a photonic and excitonic nature. While the fast escape of photons out of the microcavity makes polaritons intrinsically open-dissipative, their excitonic component provides strong repulsive interactions. With the optical pump to replenish the particles against the radiative loss, microcavity polaritons constitute an archetypal system to study the effect of quantum fluids, in which remarkable effects such as superfluidity [2], bistability [3], Bogoliubov excitation spectrum [4], and the hydrodynamic nucleation of quantized vortices [5–7] or dark solitons [8,9] have been observed.

Dark soliton, a fundamental nonlinear collective excitation, is a self-localized and shape-preserving stationary solution in a one-dimensional (1D) quantum degenerated fluids with positive mass and repulsive interactions. They can be thought of natural modes of nonlinear wave equations and have been observed in a wide variety of systems, such as atomic condensates [10,11], optical fibers [12], superconductors [13,14] and magnetic films [15]. In polariton system, due to the intrinsic dissipative nature, solitons are known to be qualitatively different from the ones present in Hamiltonian systems. They may exist with an unstable state and be evolved in a complicated and often chaotic manner [16, 17]. Several proposals are presented to improve the stability of the dark soliton in some particular cases: under the nonresonant excitation, the dark soliton trains with stability are firstly theoretically demonstrated in the spinor system with the help of potential step [18], and three years later, the controls of the stable dark soliton on demand [19] is reported in the scalar system. Under the coherent resonant excitation, scalar dark soliton trains [20] are also demonstrated with the pulsed excitation [21] based on the mechanism of quantum interference. Detailed studies about the dark soliton (trains) with stability are still required with the consideration of underlying application in photonics and quantum information.

The purpose of this paper is to, by exploiting the quantum interference in spatial domain, investigate the generation and stability of the sink-type dark soliton trains in a nonresonant polariton condensate excited with the cw homogeneous field. The nonresonant excitation with cw-homogeneous pump is crucial because of the fact in this context: the free evolution of the condensate phase in contrast to the resonant injection scheme that would imprint the phase. With the help of potential trap to mimic the finite-size effect, the polariton condensate is spatially separated into two parts to induce the quantum interference in spatial domain, which introduce the generation of a pair of sink-type dark solitons. The domain of the material parameters to observe the sink-type soliton train is also given, which is in accordance with the restriction of repulsive interaction for the generation of polariton dark soliton [22]. Further, the number of solitons in the train is controllable on demand in the domain of parameters such as potential width and amplitude. With the increase of pump power, the stationary sink-type dark soliton train could achieve enough energy to periodically oscillate and collide and finally enters into the final state of anti-dark soliton. Finally, with the consideration of disturbance from noise, these nonlinear excitations are approved to have infinite lifetime and remain spatially localized even for the periodically oscillating and colliding state, which don’t depend on the parameters of the condensate. The stability is further investigated with the analysis of Bogliubov method.

2. Model

In the analysis below, we consider the system of a wire-shaped microcavity modeled in Fig. 1, where the polaritons are bounded to a quasi-1D channel as implemented in Ref. [23]. A metallic mesas is deposited on top of the sample to create the potential trap V(x) for polariton particles with a tunable amplitude by applying an electric field [24]. The polariton condensate is assumed to be nonresonantly populated with a far detuned cw-homogeneous pump P over the whole sample. The polariton wave function ψ(x) evolves along the open-dissipative Gross-Pitaevskii equation (ODGPE), coupled to the rate equation of exciton reservoir density nR(x):

dψ(x)=[2m*2ψ(x)x2(i+D0nR(x))+12(R1DnR(x)γC)Rψ(x)i(gC1D|ψ(x)|2ψ(x)+gR1DnR(x)ψ(x)+V(x)ψ(x))]dt+dWnR(x)t=D12nR(x)x2+P(γR+R1D|ψ(x)|2)nR(x)
where the real coefficients D0 and D1 represent respectively the energy relaxation in the condensate and the spatial diffusion rate of reservoir polaritons. m* is the effective polariton mass, R1D presents the incoming rate of polaritons from the exciton reservoir, γi(i = C, R) is the decay rate of polaritons and exciton reservoir, gi1D(i = C, R) depicts the interaction coefficient between polaritons and exciton reservoir. dW is the quantum fluctuation noise given by a complex stochastic term in the truncated Wigner approximation (TWA)
<dW(x)dW(x)>=0<dW(x)dW*(x)>=dt2dx(R1DnR+γC)δx,x
the potential trap V(x) described in this work is given by
V(x)={V0|x|a0|x|>a
here a is set to mimic the finite-size effect so that the condensate is spatially separated into two parts to induce the rich dynamics of counter-propagating polariton waves.

 figure: Fig. 1

Fig. 1 Sketch of a potential sample consisting in a quasi-1D semiconductor microcavity [distributed Bragg reflectors (DBR)]. A metallic mesas is embedded and deposited in the central area with a tunable potential amplitude by applying an electric field.

Download Full Size | PDF

3. Generation and modulation of dark soliton train

In the scalar system with potential trap V(x) described in previous section, polariton condensates are spatially divided into two parts with the nonresonant excitation of one cw homogeneous field above the threshold Pth = γCγR/R1D. The quantum interference [21] between two neighbouring polariton condensates plays a keys role for the dynamics of the scalar system. Fig. 2(a) shows the dynamical evolution of the scalar system with parameters V0 = 0.2mev, a = 3μm and P = 1.2Pth. Under the effect of homogeneous nonresonant pump, polariton particles are stimulated toward the lowest energy state and form the condensate in momentum space with a chemical potential μgC1D|ψ|4+gR1DnR|ψ|2. In spatial space, however, rich nonlinear phenomena such as a pair of sink-type dark solitons appear. Two parts of polariton condensates could be understood as two counterpropagating nonlinear waves with dissipation, which heal from the potential step and collide at the center of potential trap to form the quantum interference with dissipation. While the constructive quantum interference occurs accompanied with the balance between the kinetic energy 22m*ψ*2ψ and the nonlinear energy gC1D|ψ|4, the phenomena of a pair of sink-type dark solitons appear. It has been predicted numerically [17] and analytically [22] that dark solitons in nonresonant polariton condensate relax by blending with the background at a finite time. With the help of potential trap, however, the blending is avoided and the stationary case is observed in our model with the consideration of both quantum and classical noise. Here t = 104ps is the cutoff time of the simulation.

 figure: Fig. 2

Fig. 2 (a) dynamical evolution of the system towards the generation of a pair of sink-type dark solitons, (b) |ψ| and phase of a pair of sink-type dark solitons as a function of spatial coordinate at t = 104ps, (c) phase diagram depicting parameters for which sink-type dark soliton trains are achieved. Parameters are: (a) and (b) γR=2γC=115ps1, P = 1.2Pth; (c) γC=130ps1, α = γC/γR. other parameters are: V0=−0.2mev, a = 3μm, m* = 5 × 10−5me, R1D = 2.24 × 10−4μm ps−1, gR1D=2gC1D=0.95μeVμm

Download Full Size | PDF

Figure 2(c) presents a phase diagram of domain for the existence of sink-type dark soliton trains in the parameter space α = γC/γR and P. The shaded area, corresponding to parameters existing sink-type dark soliton trains, is limited by the maximum value of α given by the condition (gR1DγC)/(gC1DγR)<1. The limitation is originated from the nature for the creation of dark soliton: repulsion between particles. As demonstrated in Ref. [22], the character of nonlinearity is effectively switched from repulsive to attractive at (gR1DγC)/(gC1DγR)>1. The limitation of α is also investigated for the fixed parameter of γR = 100ps, not shown in the paper.

As discussed above, the generation of the sink-type dark soliton train is induced by the quantum interference originated from two counter propagating polariton waves. It has been pointed out that [21], under the excitation of two optical pulse, constructive and destructive quantum interference change periodically with the product of chemical potential of the condensate and distance between two pump fields. As shown in Fig. 3, the dependence of the quantum interference on the width of the potential trap is investigated with a fixed chemical potential by setting P = 1.2Pth. Here we set a < 10μm to be sure the finite-size effect so that polariton condensates are spatially separated into two parts. By modifying the width from a = 2μm to a = 9μm, the number of the solitons varies accordingly. The destructive quantum interference with odd number of sink-type solitons in the train is observed only with the condition of 1μm< a < 3μm, as shown in Fig. 3(a). With the further increase of a, the constructive quantum interference with even number of sink-type solitons in the train can always be observed and the only difference is the variation of the soliton number. The phenomena is a little different from what is expected for the periodical changes of quantum interference as discussed in Ref. [21]. The case is supposed to be relevant to the model of potential trap. Only the quantum interference with wavefunction in accordance with that of the bound states in the potential trap could be survived in the dynamical evolution of the system. Additionally, we would like to point out that, for sink-type soliton trains as shown in Fig. 3, all polariton particles condensate in an agglomeration region in momentum space with the same chemical potential. However, for the parameters where the sink-type soliton train is observed but easy to jump to another case with a small modification in parameters, a few polariton particles appear in the vicinity of the agglomeration region accompanied with a higher chemical potential, which is not shown in the paper.

 figure: Fig. 3

Fig. 3 the phenomena of sink-type dark soliton trains as a function of spatial coordinate for different potential width while t = 104ps. (a) a=2μm, (b) a=5μm, (c) a=8μm, (d) a=9μm. other parameters are the same as those in Fig. 2(a).

Download Full Size | PDF

Further, with a fixed width a of potential trap, the effect of pump power P on the sink-type dark soliton train is investigated as shown in Fig. 4. Similar as the case talked before, due to the action of bound states in potential trap, the type of quantum interference won’t be changed periodically with pump power P and depends on the depth of potential trap. The shallower the potential trap, the less number the bound states. The parity of wavefunction corresponding to the bound state represents the type of quantum interference possible to survive in the dynamical evolution. For the shallow trap such as V0 = −0.2mev and a = 3μm (corresponding to the case of Fig. 2(a)), there is only one bound state with wavefunction of even parity. This predicts that only a pair of sink-type dark solitons could be observed in this case. With the increase of pump power up to P = 2.0Pth, as shown in Fig. 4(a), the sink-type soliton train obtain enough energy to move and brings on the generation of periodically oscillating and colliding sink-type dark soliton pairs. With the further increase of pump power up to P = 2.2Pth, the polariton particles achieve much more energy to escape the bound of potential trap and condensate in momentum space with higher energy of k ≠ 0. It is conductive to the generation of the anti-dark state as shown in Fig. 4(b).

 figure: Fig. 4

Fig. 4 dynamical evolution of the system towards the generation of (a) a pair of periodically oscillating and colliding dark solitons as well as (b) anti dark soliton. Parameters are (a) P = 2.0Pth, (b) P = 2.2Pth. Other parameters are the same as those in Fig. 2(a).

Download Full Size | PDF

It is important to note that, besides two factors of pump power and potential width, there is another factor of V0 noticeable in our system to modulate the dark soliton train. The larger value of V0, the more number of bound states to observe corresponding dark soliton trains. With a fixed pump power P = 1.6Pth and a given potential width a = 3μm, Fig. 5 shows the variation of the sink-type dark soliton trains for different value of V0. The number of solitons in the train appears from two to three and then to four with the increase of V0 from 0.2mev up to 0.7mev. With the consideration that the phase of condensate changes with the amplitude of potential trap, it is supposed that the choice of the bound state for the possible quantum interference in this model is relative to the flux flow velocity while the polariton particle heals from the potential step, which is determined by the variation of the phase of condensate.

 figure: Fig. 5

Fig. 5 dark soliton trains vs the amplitude of potential trap. (a) V0 = 0.2mev, (b) V0 = 0.7mev, (c) V0 = 1.8mev. P = 1.6Pth and other parameters are the same as those in Fig. 2(a).

Download Full Size | PDF

4. Stability of the dark soliton train

In the previous simulation, the random low-intensity white noise is seeded both in the initial condition and during the evolution. The fact of the solution staying for a long time indicates that this is a steady state in a nonequilibrium system. Next, we would like to further check the stability of the solution within the Bogoliubov-de Geneess approximation. The small fluctuations around the stationary state can be decomposed into their Fourier components:

ψ(x,t)=eiμtψ0(x)[1+k(ukeikxiλkt+vk*eikx+iλk*t)],nR(x,t)=nR0(x)(1+k(wkeikxiλkt+wk*eikx+iλk*t))
where λk is the mode frequency of the wavenumber k, and uk, vk, wk are small fluctuations. ψ0(x) and nR0(x) are stationary solutions of the system for the dark soliton train, with μ the corresponding chemical potential. The imaginary part of λk is equal to the exponential growth rate of an unstable mode. Substituting Eq. (4) into the Eq. (1) and keeping the linear terms only, the eigenvalue problem λk(uk, vk, wk)T = A(uk, vk, wk)T of the elementary excitations can be derived, with the operator matrix A dependent on the stationary solution ψ0(x).

Typical examples of the elementary excitation spectra are shown in Figs. 6(a)–(d), respectively corresponding to the sink-type dark soliton trains as shown in Figs. 3(a) and 3(b). These two cases represent the every possible scenario of dispersion of the sink-type dark soliton trains observed in our model. It is noted in Fig. 2(c) that the sink-type dark soliton trains in our model can be observed only within the parameter domain α = γC/γR ≤ 1/2 to maintain the repulsive property between polariton particles [22]. Here we would like to note that this condition is under the adiabatic condition of γCγR for polariton condensate. Therefore it is not surprised to observe the similar dispersion profiles (as shown in Figs. 6(a) and 6(b)) for the nonlinear excitation of dark soliton trains as those of the polariton condensate under the adiabatic condition [25, 26]: a Goldstone mode (indicated as + in the figure) whose real part is equal to zero and appears dispersionless at low k, accompanied with an imaginary part starting from zero in a quadratic way. Additionally, there is another kind of dispersion as shown in Figs. 6(b) and 6(d): although with the same real part as that in Fig. 3(a), there is a little difference in the imaginary part. It still starts from zero but not in a quadratic way and oscillation appears. Anyway, in all these cases, the imaginary part of λk no bigger than 0 depicts the stability of the sink-type dark soliton train created in our model, which is in accordance with the information provided by the numerical simulation with the consideration of noise.

 figure: Fig. 6

Fig. 6 real and imaginary part of the excitation spectrum of the dark soliton train. (a) and (b) corresponding to the case of Figs. 3(a), (c) and (d) corresponding to the case of Fig. 3(b).

Download Full Size | PDF

5. Conclusion

In conclusion, we found that, with the action of potential trap to prepare the spatially separated binary condensate, the nonlinear excitation of sink-type dark soliton train could be generated and stabilized in a 1D nonresonant pumped polariton condensate. The mechanism is relative to the quantum interference originated from two counter propagating polariton waves. We proceed to explore the on demand control of the soliton number in the train with the help of potential parameters. The parameter domain to generate the sink-type dark soliton train is also presented, which renders them available for experimental observations. We’d like to note that the sink-type solution with stability [27] has also been discused recently with the help of spatially localized pumping where the heating effect is the considered. The present results on the generation and control of the sink-type dark soliton train could also have implication in matter-wave interferometry.

Funding

National Natural Science Foundation of China (NSFC) (11374125); China Postdoctoral Science Foundation (2015M581391, 2015M570266).

References

1. C. Weisbuch, M. Nishioka, A. Ishikawa, and Y. Arakawa, “Observation of the coupled exciton-photon mode splitting in a semiconductor quantum microcavity,” Phys. Rev. Lett. 69, 3314–3317 (1992). [CrossRef]   [PubMed]  

2. A. Amo, J. Lefrére, S. Pigeon, C. Adrados, C. Ciuti, I. Carusotto, R. Houdré, E. Giacobino, and A. Bramati, “Superfluidity of polaritons in semiconductor microcavities,” Nat. Phys. 5, 805–810 (2009). [CrossRef]  

3. V. Goblot, H. S. Nguyen, I. Carusotto, E. Galopin, A. Lemaître, I. Sagnes, A. Amo, and J. Bloch, “Phase-controlled bistability of a dark soliton train in a polariton fluid,” Phys. Rev. Lett. 117, 217401 (2016). [CrossRef]   [PubMed]  

4. P. Stepanov, I. Amelio, J. Rousset, J. Bloch, A. Amo, A. Minguzzi, I. Carusotto, and M. Richard, “Two-components nature of the excitations in a polariton superfluid,” arXiv 1810.12570v1 (2018).

5. K. G. Lagoudakis, M. Wouters, M. Richard, A. Baas, I. Carusotto, R. André, L. S. Dang, and B. D. Plédran, “Quantized vortices in an exciton-polariton condensate,” Nat. Phys. 4, 706–710 (2008). [CrossRef]  

6. D. Sanvitto, F. M. Marchetti, M. H. Szymanska, G. Tosi, M. Baudisch, F. P. Laussy, D. N. Krizhanovskii, M. S. Skolnick, L. Marrucci, A. Lemaître, J. Bloch, C. Tejedor, and L. Viña, “Persistent currents and quantized vortices in a polariton superfluid,” Nat. Phys. 6, 527–533 (2010). [CrossRef]  

7. G. Nardin, G. Grosso, Y. Léger, B. Pietka, F. Morier-Genoud, and B. Deveaud Plédran, “Hydrodynamic nucleation of quantized vortex pairs in a polariton quantum fluid,” Nat. Phys. 7, 635–641 (2011). [CrossRef]  

8. A. Amo, S. Pigeon, D. Sanvitto, V. G. Sala, R. Hivet, I. Carusotto, F. Pisanello, G. Leménager, R. Houdré, E. Giacobino, C. Ciuti, and A. Bramati, “Polariton superfluids reveal quantum hydrodynamic solitons,” Science 332, 1167–1170 (2011). [CrossRef]   [PubMed]  

9. M. Sich, D. N. Krizhanovskii, M. S. Skolnick, A. V. Gorbach, R. Hartley, D. V. Skryabin, E. A. Cerda-Méndez, K. Biermann, R. Hey, and P. V. Santos, “Observation of bright polariton solitons in a semiconductor microcavity,” Nat. Photon. 6, 50–55 (2012). [CrossRef]  

10. R. Dum, J. I. Cirac, M. Lewenstein, and P. Zoller, “Creation of dark solitons and vortices in Bose-Einstein condensates,” Phys. Rev. Lett. 80, 2972–2975 (1998). [CrossRef]  

11. P. Engels and C. Atherton, “Stationary and nonstationary fluid flow of a Bose-Einstein condensate through a penetrable barrier,” Phys. Rev. Lett. 99, 160405 (2007). [CrossRef]   [PubMed]  

12. A. M. Weiner, J. P. Heritage, R. J. Hawkins, R. N. Thurston, E. M. Kirschner, D. E. Leaird, and W. J. Tomlinson, “Experimental observation of the fundamental dark soliton in optical fibers,” Phys. Rev. Lett. 61, 2445–2448 (1988). [CrossRef]   [PubMed]  

13. K. Harada, O. Kamimura, H. Kasai, and T. Matsuda, “Direct observation of vortex dynamics in superconducting films with regular arrays of defects,” Science 274, 1167–1170 (1996). [CrossRef]   [PubMed]  

14. D. Roditchev, C. Brun, L. Serrier-Garcia, J. C. Cuevas, V. H. L. Bessa, M. V. Milosevic, F. Debontridder, V. Stolyarov, and T. Cren, “Direct observation of Josephson vortex cores,” Nat. Phys 11, 332–337 (2015). [CrossRef]  

15. M. Chen, M. A. Tsankov, J. M. Nash, and C. E. Patton, “Microwave magnetic-envelope dark solitons in yttrium iron garnet thin films,” Phys. Rev. Lett. 70, 1707–1710 (1993). [CrossRef]   [PubMed]  

16. P. Cilibrizzi, H. Ohadi, T. Ostatnicky, A. Askitopoulos, W. Langbein, and P. Lagoudakis, “Linear wave dynamics explains observations attributed to dark solitons in a polariton quantum fluid,” Phys. Rev. Lett. 113, 103901 (2014). [CrossRef]   [PubMed]  

17. Y. Xue and M. Matuszewski, “Creation and abrupt decay of a quasistationary dark soliton in a polariton condensate,” Phys. Rev. Lett. 112, 216401 (2014). [CrossRef]  

18. F. Pinsker and H. Flayac, “On-demand dark soliton train manipulation in a spinor polariton condensate,” Phys. Rev. Lett. 112, 140405 (2014). [CrossRef]   [PubMed]  

19. X. Ma, O. A. Egorov, and S. Schumacher, “Creation and manipulation of stable dark solitons and vortices in microcavity polariton condensates,” Phys. Rev. Lett. 118, 157401 (2017). [CrossRef]   [PubMed]  

20. V. Goblot, H. S. Nguyen, I. Carusotto, E. Galopin, A. Lemaître, I. Sagnes, A. Amo, and J. Bloch, “Phase-controlled bistability of a dark soliton train in a polariton fluid,” Phys. Rev. Lett. 117, 217401 (2016). [CrossRef]   [PubMed]  

21. H. Ohadi, R. L. Gregory, T. Freegarde, Y. G. Rubo, A. V. Kavokin, N. G. Berloff, and P. G. Lagoudakis, “Nontrivial phase coupling in polariton multiplets,” Phys. Rev. X 6, 031032 (2016).

22. L. A. Smirnov, D. A. Smirnova, E. A. Ostrovskaya, and Y. S. Kivshar, “Dynamics and stability of dark solitons in exciton-polariton condensates,” Phys. Rev. B 89, 235310 (2014). [CrossRef]  

23. E. Wertz, L. Ferrier, D. D. Solnyshkov, R. Johne, D. Sanvitto, A. Lemaître, I. Sagnes, R. Grousson, A. V. Kavokin, P. Senellart, G. Malpuech, and J. Bloch, “Energy relaxation in a 1-D polariton condensate,” Nat. Phys. 6, 860–864 (2010). [CrossRef]  

24. H. Flayac, G. Pavlovic, M. A. Kaliteevski, and I. A. Shelykh, “Electric generation of vortices in polariton superfluids,” Phys. Rev. B 85, 075312 (2012). [CrossRef]  

25. M. Wouters and I. Carusotto, “Excitations in a nonequilibrium Bose-Einstein condensate of exciton polaritons,” Phys. Rev. Lett. 99, 140402 (2007). [CrossRef]   [PubMed]  

26. M. H. Szymańska, J. Keeling, and P. B. Littlewood, “Nonequilibrium quantum condensation in an incoherently pumped dissipative system,” Phys. Rev. Lett. 96, 230602 (2006). [CrossRef]  

27. I. Yu. Chestnov, T. A. Khudaiberganov, A. P. Alodjants, and A. V. Kavokin, “Heat-assisted self-localization of exciton polaritons,” Phys. Rev. B 98, 115302 (2018). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Sketch of a potential sample consisting in a quasi-1D semiconductor microcavity [distributed Bragg reflectors (DBR)]. A metallic mesas is embedded and deposited in the central area with a tunable potential amplitude by applying an electric field.
Fig. 2
Fig. 2 (a) dynamical evolution of the system towards the generation of a pair of sink-type dark solitons, (b) |ψ| and phase of a pair of sink-type dark solitons as a function of spatial coordinate at t = 104ps, (c) phase diagram depicting parameters for which sink-type dark soliton trains are achieved. Parameters are: (a) and (b) γ R = 2 γ C = 1 15 ps 1, P = 1.2Pth; (c) γ C = 1 30 ps 1, α = γC/γR. other parameters are: V0=−0.2mev, a = 3μm, m* = 5 × 10−5me, R1D = 2.24 × 10−4μm ps−1, g R 1 D = 2 g C 1 D = 0.95 μ e V μ m
Fig. 3
Fig. 3 the phenomena of sink-type dark soliton trains as a function of spatial coordinate for different potential width while t = 104ps. (a) a=2μm, (b) a=5μm, (c) a=8μm, (d) a=9μm. other parameters are the same as those in Fig. 2(a).
Fig. 4
Fig. 4 dynamical evolution of the system towards the generation of (a) a pair of periodically oscillating and colliding dark solitons as well as (b) anti dark soliton. Parameters are (a) P = 2.0Pth, (b) P = 2.2Pth. Other parameters are the same as those in Fig. 2(a).
Fig. 5
Fig. 5 dark soliton trains vs the amplitude of potential trap. (a) V0 = 0.2mev, (b) V0 = 0.7mev, (c) V0 = 1.8mev. P = 1.6Pth and other parameters are the same as those in Fig. 2(a).
Fig. 6
Fig. 6 real and imaginary part of the excitation spectrum of the dark soliton train. (a) and (b) corresponding to the case of Figs. 3(a), (c) and (d) corresponding to the case of Fig. 3(b).

Equations (4)

Equations on this page are rendered with MathJax. Learn more.

d ψ ( x ) = [ 2 m * 2 ψ ( x ) x 2 ( i + D 0 n R ( x ) ) + 1 2 ( R 1 D n R ( x ) γ C ) R ψ ( x ) i ( g C 1 D | ψ ( x ) | 2 ψ ( x ) + g R 1 D n R ( x ) ψ ( x ) + V ( x ) ψ ( x ) ) ] d t + d W n R ( x ) t = D 1 2 n R ( x ) x 2 + P ( γ R + R 1 D | ψ ( x ) | 2 ) n R ( x )
< d W ( x ) d W ( x ) > = 0 < d W ( x ) d W * ( x ) > = d t 2 d x ( R 1 D n R + γ C ) δ x , x
V ( x ) = { V 0 | x | a 0 | x | > a
ψ ( x , t ) = e i μ t ψ 0 ( x ) [ 1 + k ( u k e i k x i λ k t + v k * e i k x + i λ k * t ) ] , n R ( x , t ) = n R 0 ( x ) ( 1 + k ( w k e i k x i λ k t + w k * e i k x + i λ k * t ) )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.