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

Light bullets in coupled nonlinear Schrödinger equations with variable coefficients and a trapping potential

Open Access Open Access

Abstract

We analyze three-dimensional (3D) vector solitary waves in a system of coupled nonlinear Schrödinger equations with spatially modulated diffraction and nonlinearity, under action of a composite self-consistent trapping potential. Exact vector solitary waves, or light bullets (LBs), are found using the self-similarity method. The stability of vortex 3D LB pairs is examined by direct numerical simulations; the results show that only low-order vortex soliton pairs with the mode parameter values n ≤ 1, l ≤ 1 and m = 0 can be supported by the spatially modulated interaction in the composite trap. Higher-order LBs are found unstable over prolonged distances.

© 2017 Optical Society of America

1. Introduction

Three-dimensional optical spatiotemporal solitary waves, also known as light bullets (LBs), are self-sustained wave packets that are localized in both spatial dimensions and time [1–4]. They may form in dielectric media in which diffraction and dispersion are balanced by medium’s nonlinearity. The generation of LBs is a nontrivial task from the analytical and numerical points of view, and even more complicated in real experimental settings [5,6].

It is well known that optical solitons appear in the nonlinear Schrödinger equation (NLSE) with cubic self-focusing optical nonlinearity, which models a wealth of phenomena in nonlinear science in general and physics in particular. Owing to the occurrence of optical beam collapse, optical solitons are unstable in two and three dimensions (2D, 3D) [7,8]. However, in the past decades, the stability of solitons in 2D and 3D has been improved, using different media and methods. One typically uses quadratic and cubic nonlinear optical media that support stable optical solitons regardless the physical dimension in which they form and propagate [8,9]. In 1999, a seminal experimental study was reported by the group of Wise [10], in which, using the achromatic phase matching technique and generating the necessary anomalous GVD, robust (2 + 1)D nonlinear LBs were formed in quadratic nonlinear crystals. We also mention several other physical settings which are adequate for getting robust LBs. These include (a) either saturable [11,12] or nonlocal [13,14] optical materials, (b) optical media with competing quadratic, cubic or cubic-quintic nonlinearities [15,16], (c) confining by two- or three-dimensional optical lattices [17–20], and (d) periodic waveguide structures with controlled diffraction and/or GVD [21–23].

Spatial solitons can be generally divided into scalar solitons (single-component) and vector solitons (multicomponent), according to the number of field components and the nature of the medium [24]. Here, we are interested in the vector type, arising in (3 + 1) dimensions in a Kerr medium with variable coefficients and a trapping potential. It is well known that vector solitons commonly form in the absence of interference between components. In general, there exist many ways to generate vector solitons, such as cross-phase modulation (XPM) [25] or considering beams of different wavelengths [26] and using mutually incoherent beams which consist of two-component azimuthons [27]. In nonlinear optics, the XPM-mediated interaction between mutually incoherent or orthogonally polarized waves leads to the formation of bound states that are also vector solitons.

A field in which spatiotemporal solitons appear regularly is the Bose-Einstein condensation [28]. Recently, a theoretical analysis and numerical study of quantized vortices in a rotating Bose-Einstein condensate (BEC) with spatiotemporally modulated interaction in a harmonic potential, was reported in [29]. In [30], 2D vector matter waves in the form of soliton-vortex and vortex-vortex pairs have been investigated for the case of attractive intra-component interaction in two-component BECs. Further, in two-component BECs, the super-flow of atomic spinor BECs optically trapped in a ring-shaped geometry were considered in [31,32]. In our previous studies, 3D approximate but still analytical spatiotemporal vector LBs were built with the help of spherical harmonics, including multipole solutions and necklace rings [33]. In [34], we constructed exact self-similar soliton solutions of 3D coupled Gross-Pitaevskii equations for two-species BECs in a combined time-dependent harmonic-lattice potential. We have investigated the control and manipulation of solitary waves for three kinds of BECs with changing diffraction and nonlinearity coefficients; the solutions include Ma breathers, and Peregrine and Akhmediev solitons [34].

However, thus far no exact 3D spatial vector LBs in two-component NLSE with spatially modulated coefficients and a trapping potential have been found, either theoretically or experimentally. Here, we extend the analysis performed in [33–36], to demonstrate such exact solutions of the coupled NLSE in (3 + 1)D with varying coefficients and a self-consistent trapping potential.

The paper is organized as follows. The coupled (3 + 1)D NLSE with the modulated coefficients and a composite trapping potential are analyzed in Sec. II, where the nonlinearity coefficients are allowed to vary in space. Also presented in Sec. II are the vortex-vortex LB pairs, obtained using the self-similarity method. The stability of these vortex LB pairs is discussed in Sec. III. Characteristic distributions of LBs are provided in Sec. IV. The concluding remarks with a simple summary are given in Sec. V.

2. The model and exact soliton solutions

In this paper we consider vector solitons consisting ofNmutually incoherent optical components in a nonlinear Kerr medium. The propagation of the slowly-varying field componentsuj(r,φ,θ,z)(j=1,2,3N) is described by a system of coupled dimensionless (3 + 1)D NLSEs with changing coefficients, under the influence of an optical trapping potential [37, 38]:

iujz+122uj+χ(r)Iuj+V(r)uj=0,
where2=1r2r(r2r)+1r2sinθθ(sinθθ)+1r2sin2θ2φ2is the 3D Laplacian in spherical coordinates, r=x2+y2+τ2 is the spatiotemporal radius, andφis the azimuthal coordinate in the transverse(x,y)plane. The evolution variable is the propagation distance z, not time. The retarded timeτis the third ‘transverse’ coordinate in this coordinate system. Since cos(θ)=τr, the colatitude angle θ is also spatiotemporal in nature. The self-consistent composite trapping potential V is specified below. The symbol I in coupled Eq. (1) represents the total intensity, I=|uj|2, thus the nonlinearity is of the Kerr type.

We assume the solutions of Eq. (1) in a form that separates the angular variables from the spatial onesuj=ψ(r,z)Yj(θ,φ). When the self-consistency condition j=1n|Yj(θ,φ)|2=1 is imposed, substituting uj into (1) leads to:

1Yj[1sinθθ(sinθθ)Yj+1sin2θ2Yjφ2]=l(l+1),
2r2ψ[iψz+12(2rψr+2ψr2)+χ(r)|ψ|2ψ+Vψ]=l(l+1),
where l is the separation constant, cast in a convenient form. As a consequence of the self-consistency condition, the total intensity I = I(r,z) should be radially symmetric at any z; hence, in our search for vector solutions of Eq. (1) we are restricted to the solutions that maintain this property.

Equation (2) allows solutions in the form of spherical harmonics Yn=KΦnm(φ)Plm(cosθ), whereK=(2l+1)(lm)!4π(l+m)!andPlm(cosθ) are the associated Legendre polynomials withlm0.The azimuthal part of Yn is of the form Φnm(φ)=ancos(mφ)+bnsin(mφ) where m is the topological charge (TC) and an and bn are the coefficients satisfying the conditions specified below [35,36]; index n here enumerates the number of components. In this manner, the angular part of the solutions to Eq. (1) carries two indices, l and m. For a fixedlthe indexmtakes the values0,1,,l1,l.We chose the simplest case of 3D vector solitons, with two components,N=2. The complex coefficients a1,a2andb1,b2are calculated from Eqs. (4) and (5) below, and satisfy the following relations [38,39]:

a1=(1+q2)1/2,b1=iqa1,
a2=qa1,b2=±ia1,
where the parameterq[0,1] determines the modulation depth of the beam [29]. Hence, the coefficients an and bn depend on q only.

Our aim in finding analytical solutions of Eq. (3) is to connect them with the stationary NLSE

12dU2dR2+EU+χ0U3=0
whereU=U(R)depends only onR=R(z,r), both of which are real functions. Here,Eis the eigenvalue of the NLS equation, andχ0the nonlinearity coefficient. This equation has many known solutions; for example, whenχ0=1and E<0, the bright soliton solution of Eq. (6) is found,
U(R)=2Esech(2ER).
To connect solutions of Eq. (3) with those of Eq. (6), one introducesψ=ρU[R(r)]eiϕ, where the amplitudeρ(z,r)and the phaseϕ(z,r)are real functions of z and r [27]. Combining with Eq. (6), a closed system of equations is obtained. By analyzing these equations, one finds:
χ0=χ2r4ρ6,Rz+Rrϕr=0
ρz+12(ρ2ϕr2+2ρrϕr+2ρrϕr)=0,
ρϕz+12[2ρr2ρ(ϕr)2+2rρrρm2r2]+Vρ=0,
2ρrRr+ρ2Rr2+2ρrRr=0,
To find exact solutions of Eqs. (8)-(11), we introduce a self-similarity transformation [33,35,40] ρ(z,r)=κF(θ1)/w3/2(z), ϕ(z,r)=a(z)r2+b(z). Here,κis the normalization constant, w(z)is the beam width, θ1(z,r) is the similarity variable to be determined, and a(z) and b(z) are the wave front curvature and the phase offset, respectively. These variables vary with the distance z. Substituting the presumed solutions for the amplitude and the phase into Eqs. (8)-(11),one obtainsθ1(z,r)=r2/w2(z),a(z)=dw2wdz,R(z,r)=0r2/w21τ2F2(τ)dτ, and χ(z,r)=2χ0ρ2(Rr)2. Thus, the nonlinearity coefficient is specified in terms of the beam characteristics.

Further, we introduce a variable transformation F(θ1)=θ1m2eθ12f(θ1), so that Eq. (10) transforms into:

θ1d2fdθ12+(m+32θ1)dfdθ1+nf=0,
withw22dbdzm234=n, w2dw2dz2ηw2+12w2=0, V=E(dRdr)2+ηr2m22r2. Here,ηis constant and n is assumed to be a non-negative integer, which can be considered as the radial mode number. Finally, the trapping potential V is defined more closely; it is composed of a harmonic and an inverse harmonic term, in addition to a term depending on the transformed radial function R. Thus, V is determined self-consistently with the solution presumed. Differential Eq. (12) [41] is the well-known Kummer confluent hypergeometric function, whose solutions are known as Kummer’s functions, namely f(θ1)=k=0n(1)k(2m+2n+1)!22k22n(m+n)!(2k+2n+1)!θ1k. To find simple soliton solutions, we take w(z)|z=0=w0 and dw(z)dz|z=0=0, where the subscript '0' denotes the value of the corresponding quantity at z=0. Hence, the exact soliton solution is obtained withw=w0, when the optical diffraction is exactly balanced by the nonlinearity. Other parameters in this case are:a(z)=0, andb(z)=b0(2n+m+1)zw02. Plugging these results into Eq. (2), one obtains the following analytical solution for the beam components:

uj=Kκw03/2(rw0)m(ancosmφ+bnsinmφ)Plm(cosθ)f(θ1)U(R)er22w0+i(b02n+m+1w02z).

In Eq. (1), the nonlinearity coefficient and the external potential are given in terms of the radial solution. This requires a self-consistent solution procedure, in which the conditions put on the coefficients and the potential can be considered as integrability constraints. It is an inverted solution procedure, in which one first finds the solution of the equation and then defines the coefficients and the potential in the equation in terms of the solution itself. Thus, it is a self-consistent procedure which offers the convenience of localized solutions with desirable features of equations with changing coefficients and potentials that are related to the solutions self-consistently. The procedure might lead to interesting models with realistic space-varying coefficients and potentials. While one may argue about the availability of such coefficients and potentials, we still believe that there is enough freedom in the solution method, to provide interesting solutions to viable material models.

3. Stability of the localized nonlinear vector LBs

Next, we discuss the linear stability of solutions of Eq. (1) [35]. The perturbations in the solutions are assumed of the form [42]:

uj(x,y,τ,z)=e-iλz[u0j(x,y,τ)+μj(x,y,τ)eiδz+vj*(x,y,τ)eiδ*z],
where u0j(x,y,τ)=ψj(r,θ)is the steady-state component solution of Eq. (1), λ=(2n+m+1)/w02is the propagation constant,μj(x,y,τ)andvj(x,y,τ)are the real and imaginary parts of the perturbation, andδstands for the perturbation growth rate. The superscript * denotes the complex conjugate quantity. Substituting the perturbed solution uj(x,y,τ,z)into Eq. (1) and linearizing around the steady-state solution, one finds the following eigenvalue equations:
(LjBAABLjAAAAL3jCAACL3j)(μ1v1μ2v2)=δ(μ1v1μ2v2),
whereLj=12(xx+yy+zz)2χ(u0j2+u03j2)V+λ,A=χu01u02,B=χu012,C=χu022. The third variable z in the Laplacian, in this case is τ. This eigenvalue problem is computed by the Fourier Collocation Method for discrete eigenvalues, which allows the whole spectrum to be calculated at once [43]. Once parametersm,n,l,qandηare specified, one gets the imaginary and real parts ofδfrom Eqs. (13) and (14). If imaginary parts of all eigenvaluesδare equal to zero or are positive, the soliton solutions can be stable; otherwise, the perturbed solution would grow exponentially with z, and thus, the corresponding matter waves would become linearly unstable [35,43].

4. Characteristic distributions of LBs

From Eq. (13), it is evident that the matter wave solitons are characterized by five parameters: the mode numbers n, m, l, the harmonic potential width η, and the modulation depth q. In this section, we present typical distributions of LBs, for some values of these parameters. Whenw=w0, the auxiliary function R, the nonlinearity coefficientχ, and the trapping potential V are the functions of the radial variable r only.

Typical distributions of the nonlinearity coefficientsχ(r)for different m and of the potential V(r) for different E are shown in Figs. 1 (a) and 1(b). When m=0, the nonlinearity coefficientχ(r)is localized in the form of asymmetric double annuluses; whenm=1, localized three amplitude peaks can be seen in Fig. 1(a). In form, the external potential V (r) is similar to the 2D pattern in [35]. Figures 1(c) and 1(d) illustrate the amplitude of the radial field distributionsψjfor different mode numbersm=0,1,2(c) and n = 1, 2, 3 (d). The optical field of LBs approaches 0 whenr=0andr=, forming different vortex structures. The ring size depends on the mode number n. The results show that the solutions in Eq. (13) indeed are localized.

 figure: Fig. 1

Fig. 1 (a), (b) Distributions of the nonlinearity coefficientχ(r)and the trapping potentialV(r)for n=1. (c) and (d) Amplitude of the radial field distributionsψjfor different mode numbers, m=0,1,2 (c) and n=2,3,4 (d). Other parameters: q = 0 and η=0.4.

Download Full Size | PDF

Figure 2 displays the intensity distributions of 3D LBs withn=m=1,η=0.4, andq=0.1, for different l (l=1,2,3) from top to bottom, atz=70. A variety of intensity distributions for solitons with different l are shown. For l=1 there are two layers (two pulses) along the vertical τ-axis, and the soliton is shaped as a pair of deformed ellipsoids. In general, by increasing l, the solution formsl+1layers in the vertical direction, forming multipole multi-pulse solitons. However, the total intensity distribution (the right column in Fig. 2) is radially symmetric. Obviously, when both parameters (l and m) are 0, the soliton is the fundamental spherical LB. The physical origin of this arrangement can qualitatively be understood from the nature of the self-focusing cubic nonlinearity as a third-order susceptibilityχ(3); it yields the nonlinearity polarization of the medium. Owing to the assumption of strong dispersion along the verticalτ axis, resulting in spherical harmonics, the associated Legendre polynomialP10will appear in the solution; it possesses a maximum value (plus one) and a minimum value (minus one). Therefore, the spherical harmonics will force a change of the sign near the zero point.

 figure: Fig. 2

Fig. 2 Intensity distributions of 3D LBs|u1|2,|u2|2, and the total intensity from left to right, and for different l (l=1,2,3) from top to bottom, at z=70. Parameters: m = 1, n=1, q=0.1. Other parameters are as in Fig. 1.

Download Full Size | PDF

Figure 3 shows some further examples of vector vortex LBs with l=3, q=0.1, and for differentm=0,2,3from top to bottom, also at z=70. The first row just depicts the same three figures, because m = 0. For given l, the larger the m, the larger the LB radius in the horizontal plane. Looking at the field components, the number of solitary beads in the same layer is 2m; it comes from the azimuthal angular dependence in Eq. (13). Becauseq1andm0, there are 2m maxima and minima in the range0mφ2π. In this sense, the ring-necklace components are produced steadily as m increases. The larger the m, the shorter the pulse in the vertical direction. The total vector intensity is still radially symmetric, as it must be.

 figure: Fig. 3

Fig. 3 Same as Fig. 2 but for different m (m=0,2,3), from top to bottom at z=70. The parameters are same as in Fig. 1, but for l=3, from left to right.

Download Full Size | PDF

Figure 4 displays the vortex-shaped distributions of the LBs form=l=2,q=0.5and n=1,2,3from left to right. It is seen that the profiles of LBs possess polycyclic structure with several amplitude peaks covering the same ring. LBs are self-similar and composed of four symmetrical petals in each ring (the bottom row in Fig. 4). The number of outer bright rings is n+1, the optical intensity in the center is zero, and the intensity of rings surrounding the center decreases with the increasing radial distance.

 figure: Fig. 4

Fig. 4 Evolution of the stable 3D LB distributionψ1 for m=l=2, q = 0.5, andn=1,2,3from left to right.

Download Full Size | PDF

Increasing the modulation depth q from 0 to 1, with fixed n, m, and l, will lead to less azimuthally modulated vortex rings (Fig. 5). One can observe that the intensity and the phase of the beam is modulated by the modulation depth q. Whenq=0, for each component we find a necklace LB, with the number of beads equal to 2m. By increasing q, the distance between petals decreases, and the multi-TC vortices change into vortex rings. It is apparent that the soliton has just one layer in the vertical direction.

 figure: Fig. 5

Fig. 5 Same as Fig. 4 but for m=n=3, and q=0,0.5,1 from left to right.

Download Full Size | PDF

The stability of the obtained exact solutions is a very important aspect of the present discussion, which has to be addressed numerically. In order to check the stability of Lbs, we perform direct numerical simulations, using the split-step Fourier method [44] and solve Eq. (1) by taking the analytical solution (13) atz=0as an initial condition. In Fig. 6, we present the comparison of analytical (the first column) and numerical (the second and the third columns) intensity contour plots. The linear-stability spectra are displayed in the fourth column of Fig. 6. It is found that only when the topological charge ism=0,1andn=l=1, the numerical solutions of LBs are stable against perturbation with an initial Gaussian noise level of up to 10%. But when the topological charge ism>1, and mode numbers n and l are greater than one, the soliton solution (13) is unstable in propagation and splits into spreading quasi square-shaped band structures. Further, a quadruple complex eigenvaluesδare visible in Figs. 6(h) and 6(i), thus these solitons might be linearly unstable.

 figure: Fig. 6

Fig. 6 Evolution of 3D LB distributions and linear stability spectra of the condensateψ1against the perturbation with an initial random noise of level 10%, at different propagation distances. The first columnz=0, the second and the third columnsz=600. The fourth column presents the linear stability spectra. Top:m=0,n=1; middle: m=n=1; bottom:m=1,n=2. Other parameters are the same as in Fig. 1.

Download Full Size | PDF

5. Conclusion

In summary, we have studied the propagation of LBs in coupled nonlinear Schrödinger equations with spatially modulated coefficients and a self-consistent trapping potential. Using the self-similarity method, exact vector solitary waves, or light bullets, are obtained. The stability of vortex 3D LB pairs is examined by direct numerical simulation. Our results show that the only stable low-order vortex LB pairs withn1,l1andm=0can be supported by the spatially modulated interaction in the composite trap. Higher-order LBs are found unstable over prolonged distances. Our approach may prove useful for other related problems, e.g., light propagation in Bose-Einstein condensates and plasmas.

Funding

This work is supported in China by the Joint Key Grant of National Natural Science Foundation of China and Zhejiang Province (U1509217), Hubei Province Science and Technology Support Program (2015BAA001, 2015CFC779), Hubei SMEs Innovation Fund Project (2015DAL069), Dr. Start-up fund (BK1427). In addition, work in Qatar is supported by the NPRP 6-021-1-005 project with the Qatar National Research Fund (a member of the Qatar Foundation) and in China by the Natural Science Foundation of Guangdong Province, under Grant No. 1015283001000000.

Acknowledgments

MRB acknowledges additional support from the Al Sraiya Holding Group.

References and links

1. B. Malomed, L. Torner, F. Wise, and D. Mihalache, “On multidimensional solitons and their legacy in contemporary Atomic, Molecular and Optical physics,” J. Phys. At. Mol. Opt. Phys. 49(17), 170502 (2016). [CrossRef]  

2. M. Blaauboer, B. A. Malomed, and G. Kurizki, “Spatiotemporally localized multidimensional solitons in self-induced transparency media,” Phys. Rev. Lett. 84(9), 1906–1909 (2000). [CrossRef]   [PubMed]  

3. L. Bergé, “Wave collapse in physics: principles and applications to light and plasma waves,” Phys. Rep. 303(5-6), 259–370 (1998). [CrossRef]  

4. D. Abdollahpour, S. Suntsov, D. G. Papazoglou, and S. Tzortzakis, “Spatiotemporal Airy light bullets in the linear and nonlinear regimes,” Phys. Rev. Lett. 105(25), 253901 (2010). [CrossRef]   [PubMed]  

5. D. J. Frantzeskakis, H. Leblond, and D. Mihalache, “Nonlinear optics of intense few-cycle pulses: An overview of recent theoretical and experimental developments,” Rom. J. Phys. 59, 767–784 (2014).

6. Y. Silberberg, “Collapse of optical pulses,” Opt. Lett. 15(22), 1282–1284 (1990). [CrossRef]   [PubMed]  

7. L. Bergé, “Wave collapse in physics: principles and applications to light and plasma waves,” Phys. Rep. 303(5-6), 259–370 (1998). [CrossRef]  

8. L. C. Crasovan, Y. V. Kartashov, D. Mihalache, L. Torner, Y. S. Kivshar, and V. M. Pérez-García, “Soliton “molecules”: robust clusters of spatiotemporal optical solitons,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 67(4), 046610 (2003). [CrossRef]   [PubMed]  

9. K. Gallo, A. Pasquazi, S. Stivala, and G. Assanto, “Parametric solitons in two-dimensional lattices of purely nonlinear origin,” Phys. Rev. Lett. 100(5), 053901 (2008). [CrossRef]   [PubMed]  

10. X. Liu, K. Beckwitt, and F. Wise, “Generation of optical spatiotemporal solitons,” Opt. Photonics News 82, 4631–4634 (1999).

11. D. E. Edmundson and R. H. Enns, “Particlelike nature of colliding three-dimensional optical solitons,” Phys. Rev. A 51(3), 2491–2498 (1995). [CrossRef]   [PubMed]  

12. Y. F. Chen, K. Beckwitt, F. W. Wise, and B. A. Malomed, “Criteria for the experimental observation of multidimensional optical solitons in saturable media,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 70(4), 046610 (2004). [CrossRef]   [PubMed]  

13. O. Bang, W. Krolikowski, J. Wyller, and J. J. Rasmussen, “Collapse arrest and soliton stabilization in nonlocal nonlinear media,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 66(4), 046619 (2002). [CrossRef]   [PubMed]  

14. W. Królikowski and O. Bang, “Solitons in nonlocal nonlinear media: exact solutions,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 63(1 Pt 2), 016610 (2001). [PubMed]  

15. D. Mihalache, D. Mazilu, L. C. Crasovan, I. Towers, A. V. Buryak, B. A. Malomed, L. Torner, J. P. Torres, and F. Lederer, “Stable spinning optical solitons in three dimensions,” Phys. Rev. Lett. 88(7), 073902 (2002). [CrossRef]   [PubMed]  

16. A. T. Avelar, D. Bazeia, and W. B. Cardoso, “Solitons with cubic and quintic nonlinearities modulated in space and time,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 79(2), 025602 (2009). [CrossRef]   [PubMed]  

17. N. K. Efremidis, J. Hudock, D. N. Christodoulides, J. W. Fleischer, O. Cohen, and M. Segev, “Two-dimensional optical lattice solitons,” Phys. Rev. Lett. 91(21), 213906 (2003). [CrossRef]   [PubMed]  

18. Y. V. Kartashov, V. A. Vysloukh, and L. Torner, “Rotary solitons in bessel optical lattices,” Phys. Rev. Lett. 93(9), 093904 (2004). [CrossRef]   [PubMed]  

19. S. Nixon, L. Ge, and J. Yang, “Stability analysis for solitons in PT-symmetric optical lattices,” Phys. Rev. A 85(2), 023822 (2012). [CrossRef]  

20. D. Mihalache, D. Mazilu, F. Lederer, B. A. Malomed, Y. V. Kartashov, L. C. Crasovan, and L. Torner, “Stable spatiotemporal solitons in bessel optical lattices,” Phys. Rev. Lett. 95(2), 023902 (2005). [CrossRef]   [PubMed]  

21. A. V. Gorbach and D. V. Skryabin, “Spatial solitons in periodic nanostructures,” Phys. Rev. A 79(5), 053812 (2009). [CrossRef]  

22. A. Szameit, I. L. Garanovich, M. Heinrich, A. Minovich, F. Dreisow, A. A. Sukhorukov, T. Pertsch, D. N. Neshev, S. Nolte, W. Krolikowski, A. Tünnermann, A. Mitchell, and Y. S. Kivshar, “Observation of diffraction-managed discrete solitons in curved waveguide arrays,” Phys. Rev. A 78(3), 031801 (2008). [CrossRef]  

23. T. X. Tran and F. Biancalana, “Diffractive resonant radiation emitted by spatial solitons in waveguide arrays,” Phys. Rev. Lett. 110(11), 113903 (2013). [CrossRef]   [PubMed]  

24. C. Weilnau, M. Ahles, J. Petter, D. Träger, J. Schröder, and C. Denz, “Spatial optical (2+1)-dimensional scalar- and vector solitons in saturable nonlinear media,” Ann. Phys. 11(8), 573–629 (2002). [CrossRef]  

25. S. V. Manakov, “Theory of two-dimensional stationary self-focusing of electromagnetic waves,” Sov. Phys. JETP 38, 248–253 (1974).

26. W. Torruellas, Y. S. Kivshar, and G. I. Stegeman, Quadratic Solitons (Springer, 2001).

27. W.-P. Zhong and M. Belic, “Resonance solitons produced by azimuthal modulation in self-focusing and self-defocusing materials,” Nonlinear Dyn. 73(4), 2091–2102 (2013). [CrossRef]  

28. V. S. Bagnato, D. J. Frantzeskakis, P. G. Kevrekidis, B. A. Malomed, and D. Mihalache, “Bose-Einstein condensation: Twenty years after,” Rom. Rep. Phys. 67, 5–50 (2015).

29. D. S. Wang, S. W. Song, B. Xiong, and W. M. Liu, “Quantized vortices in a rotating Bose-Einstein condensate with spatiotemporally modulated interaction,” Phys. Rev. A 84(5), 053607 (2011). [CrossRef]  

30. A. I. Yakimenko, Y. A. Zaliznyak, and V. M. Lashkin, “Two-dimensional nonlinear vector states in Bose-Einstein condensates,” Phys. Rev. A 79(4), 043629 (2009). [CrossRef]  

31. A. I. Yakimenko, K. O. Isaieva, S. I. Vilchinskii, and M. Weyrauch, “Stability of persistent currents in spinor Bose-Einstein condensates,” Phys. Rev. A 88(5), 051602 (2013). [CrossRef]  

32. A. I. Yakimenko, S. I. Vilchinskii, Y. M. Bidasyuk, Y. I. Kuriatnikov, K. O. Isaieva, and M. Weyrauch, “Generation and decay of persistent currents in a toroidal Bose-Einstein condensate,” Rom. Rep. Phys. 67, 249–272 (2015).

33. S. L. Xu, M. R. Belić, and W. P. Zhong, “Three-dimensional spatio-temporal vector solitary waves in coupled nonlinear Schrödinger equations with variable coefficients,” J. Opt. Soc. Am. B 30(1), 113–122 (2013). [CrossRef]  

34. S. L. Xu, G. P. Zhou, N. Z. Petrović, and M. R. Belić, “Nonautonomous vector matter waves in two-component Bose-Einstein condensates with combined time-dependent harmonic-lattice potential,” J. Opt. 17(10), 105605 (2015). [CrossRef]  

35. S. L. Xu, J. X. Cheng, M. R. Belić, Z. L. Hu, and Y. Zhao, “Dynamics of nonlinear waves in two-dimensional cubic-quintic nonlinear Schrödinger equation with spatially modulated nonlinearities and potentials,” Opt. Express 24(9), 10066–10077 (2016). [CrossRef]   [PubMed]  

36. S. L. Xu, N. Z. Petrovic, M. R. Belić, and W. W. Deng, “Exact solutions for the quintic nonlinear Schrödinger equation with time and space,” Nonlinear Dyn. 84(1), 251–259 (2016). [CrossRef]  

37. A. S. Desyatnikov and Y. S. Kivshar, “Necklace-ring vector solitons,” Phys. Rev. Lett. 87(3), 033901 (2001). [CrossRef]   [PubMed]  

38. S. L. Xu and M. R. Belic, “Light bullets in coupled nonlinear Schrödinger equation with spatially modulated coefficients and Bessel trapping potential,” J. Mod. Opt. 62(9), 683–692 (2015). [CrossRef]  

39. W. P. Zhong, M. Belić, G. Assanto, B. A. Malomed, and T. Huang, “Self-trapping of scalar and vector dipole solitary waves in Kerr media,” Phys. Rev. A 83(4), 043833 (2011). [CrossRef]  

40. R. Hirota, “Exact Solution of the Korteweg-de Vries Equation for Multiple Collisions of Solitons,” Phys. Rev. Lett. 27(18), 1192–1194 (1971). [CrossRef]  

41. D. Zwillinger, Handbook of Differential Equations, 3rd ed. (Academic, 1997).

42. M. Dehghan and D. Mirzaei, “The dual reciprocity boundary element method (DRBEM) for two-dimensional sine-Gordon equation,” Comput. Methods Appl. Mech. Eng. 197(6-8), 476–486 (2008). [CrossRef]  

43. B. A. Malomed, D. Mihalache, F. Wise, and L. Torner, “Spatiotemporal optical solitons,” J. Opt. B Quantum Semiclassical Opt. 7(5), R53–R72 (2005). [CrossRef]  

44. T. I. Lakoba, “Stability analysis of the split-step Fourier method on the background of a soliton of the nonlinear Schrödinger equation,” in Methods for Partial Differential Equations, 4974, 641–649 (2010).

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 (a), (b) Distributions of the nonlinearity coefficient χ ( r ) and the trapping potential V ( r ) for n = 1 . (c) and (d) Amplitude of the radial field distributions ψ j for different mode numbers, m = 0 , 1 , 2 (c) and n = 2 , 3 , 4 (d). Other parameters: q = 0 and η = 0.4 .
Fig. 2
Fig. 2 Intensity distributions of 3D LBs | u 1 | 2 , | u 2 | 2 , and the total intensity from left to right, and for different l ( l = 1 , 2 , 3 ) from top to bottom, at z = 70 . Parameters: m = 1, n = 1 , q = 0.1 . Other parameters are as in Fig. 1.
Fig. 3
Fig. 3 Same as Fig. 2 but for different m ( m = 0 , 2 , 3 ), from top to bottom at z = 70 . The parameters are same as in Fig. 1, but for l = 3 , from left to right.
Fig. 4
Fig. 4 Evolution of the stable 3D LB distribution ψ 1 for m = l = 2 , q = 0.5, and n = 1 , 2 , 3 from left to right.
Fig. 5
Fig. 5 Same as Fig. 4 but for m = n = 3 , and q = 0 , 0.5 , 1 from left to right.
Fig. 6
Fig. 6 Evolution of 3D LB distributions and linear stability spectra of the condensate ψ 1 against the perturbation with an initial random noise of level 10%, at different propagation distances. The first column z = 0 , the second and the third columns z = 600 . The fourth column presents the linear stability spectra. Top: m = 0 , n = 1 ; middle: m = n = 1 ; bottom: m = 1 , n = 2 . Other parameters are the same as in Fig. 1.

Equations (15)

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

i u j z + 1 2 2 u j + χ ( r ) I u j + V ( r ) u j = 0 ,
1 Y j [ 1 sin θ θ ( sin θ θ ) Y j + 1 sin 2 θ 2 Y j φ 2 ] = l ( l + 1 ) ,
2 r 2 ψ [ i ψ z + 1 2 ( 2 r ψ r + 2 ψ r 2 ) + χ ( r ) | ψ | 2 ψ + V ψ ] = l ( l + 1 ) ,
a 1 = ( 1 + q 2 ) 1 / 2 , b 1 = i q a 1 ,
a 2 = q a 1 , b 2 = ± i a 1 ,
1 2 d U 2 d R 2 + E U + χ 0 U 3 = 0
U ( R ) = 2 E sec h ( 2 E R ) .
χ 0 = χ 2 r 4 ρ 6 , R z + R r ϕ r = 0
ρ z + 1 2 ( ρ 2 ϕ r 2 + 2 ρ r ϕ r + 2 ρ r ϕ r ) = 0 ,
ρ ϕ z + 1 2 [ 2 ρ r 2 ρ ( ϕ r ) 2 + 2 r ρ r ρ m 2 r 2 ] + V ρ = 0 ,
2 ρ r R r + ρ 2 R r 2 + 2 ρ r R r = 0 ,
θ 1 d 2 f d θ 1 2 + ( m + 3 2 θ 1 ) d f d θ 1 + n f = 0 ,
u j = K κ w 0 3 / 2 ( r w 0 ) m ( a n cos m φ + b n sin m φ ) P l m ( cos θ ) f ( θ 1 ) U ( R ) e r 2 2 w 0 + i ( b 0 2 n + m + 1 w 0 2 z ) .
u j ( x , y , τ , z ) = e -i λ z [ u 0 j ( x , y , τ ) + μ j ( x , y , τ ) e i δ z + v j * ( x , y , τ ) e i δ * z ] ,
( L j B A A B L j A A A A L 3 j C A A C L 3 j ) ( μ 1 v 1 μ 2 v 2 ) = δ ( μ 1 v 1 μ 2 v 2 ) ,
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.