Open access peer-reviewed chapter

Dipole Solitons in a Nonlocal Nonlinear Medium with Self-Focusing and Self-Defocusing Quintic Nonlinear Responses

Written By

Mahmut Bağcı, Melis Turgut, Nalan Antar and İlkay Bakırtaş

Submitted: 17 May 2022 Reviewed: 30 June 2022 Published: 11 August 2022

DOI: 10.5772/intechopen.106207

From the Edited Volume

Vortex Dynamics - From Physical to Mathematical Aspects

Edited by İlkay Bakırtaş and Nalan Antar

Chapter metrics overview

114 Chapter Downloads

View Full Metrics

Abstract

Stability dynamics of dipole solitons have been numerically investigated in a nonlocal nonlinear medium with self-focusing and self-defocusing quintic nonlinearity by the squared-operator method. It has been demonstrated that solitons can stay nonlinearly stable for a wide range of each parameter, and two nonlinearly stable regions have been found for dipole solitons in the gap domain. Moreover, it has been observed that instability of dipole solitons can be improved or suppressed by modification of the potential depth and strong anisotropy coefficient.

Keywords

  • dipole solitons
  • nonlinear response
  • nonlocal nonlinear medium
  • quintic nonlinearity

1. Introduction

Many phenomena in nature are modeled mathematically using nonlinear differential equations. Traveling wave solutions of nonlinear partial differential equations play a significant role in nonlinear wave propagation problems that are observed in various fields such as nonlinear optics, fluid dynamics, plasma physics, elastic media, and biology [1]. Some of the solutions to such nonlinear wave propagation problems are called solitons, which are localized wave solutions.

Optical solitons are formed because of the balance between the medium’s diffraction and the self-phase modulation [2]. As a consequence of this, an optical field that does not change its shape occurs during propagation [3]. Recently, spatial solitons that can be used for optical switching and processing applications [4] have been extensively investigated in nonlinear optical systems with external optical lattices. There is a considerable amount of research about this subject in the literature. In 2003, Segev et al. experimentally observed spatial solitons in optically induced periodic potentials [5]. Fundamental and vortex solitons with real or complex lattices have been investigated in optical media with the cubic Kerr-type [6, 7, 8, 9, 10], the saturable [11], and competing nonlinearities [12]. Moreover, the existence of solitons has been observed in aperiodic or quasicrystal lattice structures [13, 14, 15, 16, 17, 18] and the lattices that possess defects [19, 20] and dislocations [21, 22].

The dynamics of solitons are governed by nonlinear Schrödinger (NLS) type equations in optical media with nonlinearities and/or external potentials as in the referred studies. Additionally, the cubic nonlinear NLS equation needs to be modified to describe nonlinear optical materials that have both cubic and quadratic nonlinear responses [23, 24, 25, 26, 27, 28, 29], such as potassium niobate (KNbO3) [30] or lithium niobate (LiNbO3) [31]. These dynamics in quadratically polarized media are governed by the NLS equation with coupling to a mean term (d.c. field), which are denoted as NLSM systems and sometimes referred to as Benney-Roskes or Davey-Stewartson systems [32, 33].

NLSM equations were first studied by Benney and Roskes for water of finite depth in the free surface conditions in 1969 [32]. Later, in 1974, Davey and Stewartson derived the limiting integrable case, which is a reduced case of the Benney–Roske’s system by studying the evolution of a 3D wave packet for water of finite depth [33]. In 1975, Ablowitz and Haberman [34] studied the integrability of NLSM systems in the shallow water limit. The effects of surface tension were included in the results of Benney and Roskes by Djordevic and Reddekopp [35] in 1977. From the first principles, Ablowitz et al. [23, 36, 37] discovered that NLSM-type equations describe the evolution of the electromagnetic field in a quadratic nonlinear media. The general NLSM system is given by [23, 36, 37]

iuz+Δu+u2uρuϕ=0,ϕxx+vϕyy=u2xx,E1

where u(x, y, z) corresponds to the normalized amplitude of the envelope of the static electric field propagating in the z direction, x and y are transverse spatial coordinates. Δuuxx+uyy corresponds to diffraction, the cubic term in u originates from the Kerr-type nonlinear change of the refractive index. The parameter ρ is a coupling constant that comes from the combined optical rectification and electro-optic effects modeled by the ϕxy field, and v is the coefficient that comes from the anisotropy of the material [37]. Such systems of equations arise due to the growth and depletion of the fundamental and second-harmonic fields at the moment that the phase velocity of the fundamental and the second-harmonic wave are not equal during propagation [38]. When the phase-matching condition is not satisfied, the equation of the second-harmonic field can be solved directly and generates an additional self-phase modulation contribution as a result of cascaded nonlinearity. Similarly, the NLSM systems describe the nonlocal–nonlinear coupling between the first harmonic with the cascading effect from the second harmonic and a static field that is related to the mean term [36, 37].

Wave collapses play a significant role in various branches of science. The peak amplitude of the wave solutions tends to infinity (blow-up) in finite time or finite propagation distance when a singularity occurs. This phenomenon is often called wave collapse [39]. In the NLS equation, it was first observed numerically by Kelley in 1965 [40]. In fact, this wave collapse phenomenon is similar for the NLSM systems. Wave collapse in the NLSM systems occurs with a modulated profile [41]. Merle and Raphael [42] analyzed the collapse behavior of the NLS equation and other related equations in detail. Moreover, Moll et al. investigated experimental observations of optical wave collapse in cubic nonlinearity and showed that the amplitude of the wave increases as the spatial extent decreases in a self-similar profile [43]. In Ref [39], Ablowitz and coworkers studied wave collapse that occurs with a quasi-self-similar profile in the NLSM system and found that collapse can be arrested by the small nonlinear saturation. Furthermore, in Ref [30], NLSM collapse was arrested by wave self-rectification. In this aforementioned study, they considered only the nonlinear evolution of beams with an initial Gaussian beam profile with several values of input power and/or beam ellipticity and found that the wave collapse can be arrested by increasing the coupling constant ρ or for an initially highly elliptic beam. Recently, the NLSM system collapse was arrested by adding a real periodic [24] and partially parity-time-symmetric [44] and azimuthal [45] external lattices (potential) to the governing system, and it was shown numerically that modification of potential depth provides great controllability on the stability of soliton.

More recently, Bağcı et al. [46] have numerically investigated stability dynamics of fundamental lattice solitons that are solutions of extended NLSM system in a nonlocal nonlinear medium with self-focusing and self-defocusing quintic nonlinearity. It has been shown that as the absolute value of γ increases for both self-focusing and self-defocusing cases, the obtained fundamental solitons become nonlinearly unstable. However, the stability of unstable fundamental solitons can be improved by modification of potential depth [46].

Dipole (two-phased) and higher-phase vortex solitons in the presence of an induced lattice have been studied analytically and experimentally in Bose-Einstein condensates (BECs) [47, 48] and in optical Kerr media [49, 50, 51, 52, 53, 54]. In recent years, these types of solitons have attracted considerable interest because of their unique features and potential applications [55].

In this chapter, we numerically study the existence and stability of dipole soliton solutions of the NLSM system in a nonlocal nonlinear medium with the self-defocusing quintic nonlinear response by adding an external lattice. In fact, this study is about the dynamics of dipole solitons instead of fundamental solitons in the problem that Bağcı and coworkers have addressed in recent book chapter [46]. The purpose of this study is to numerically investigate the effects of the strength of quintic nonlinearity that specify characteristics of the model and variation of potential depth on the existence and stability of dipole solitons. In several applications, many optical materials such as chalcogenide glasses are required quintic and seventh-order effects in addition to cubic nonlinear effects [56], and effective higher-order nonlinearities can reveal with pure Kerr materials in an inhomogeneous propagation media [57, 58, 59].

The chapter is outlined as follows: In Sec. 2, we present the model equations, and the squared-operator method is explained so that it is modified for the model. The dipole solitons are computed by this numerical method. Nonlinear evolution of the dipole solitons is examined to perform stability analysis, In Sec. 3. Finally in Sec. 4, results of this study are outlined.

Advertisement

2. The model

In this chapter, we modify the NLSM system (1) as follows to describe the dynamics of lattice solitons in a nonlocal nonlinear medium with cubic and quintic nonlinearity

iuz+12Δu+βu2uρuϕ+γu4uVxyu=0,ϕxx+vϕyy=u2xxE2

where γ is the coefficient of quintic nonlinearity and V(x, y) is the optical lattice. In this chapter, we consider lattices that can be written as the intensity of a sum of N phase-modulated plane waves [13]

Vxy=V0N2n=0N1eikxnx+kyny2,E3

where V0> 0 is the peak depth of the potential and the wave vector kxnkyn=Kcos2πn/NKsin2πn/N. The potential for N = 2, 3, 4, 6 yield crystal (periodic) lattices, while N = 5, 7 yield quasi-crystals (aperiodic) lattices. Contour image, contour plot, and diagonal cross-section of the lattice V(x, y) are plotted in Figure 1 for V0=12.5,N = 4 and kx=ky=2π. It can be seen that the lattice is periodic, and the center of lattice is a local maximum.

Figure 1.

(a) Contour image, (b) contour plot, and (c) diagonal cross-section of the lattice V(x, y) when V0=12.5, N = 4 and (x,y) ∈ [15,15].

2.1 Numerical solution for the dipole solitons

Yang and Lakoba developed an iterative numerical method called the squared-operator method (SOM) [60]. The idea of this method is to iterate a modified differential equation whose linearization operator is square of the original equation together with a preconditioning (or acceleration) operator. To obtain the soliton solution of the (2+1)D NLSM model, this method is modified as follows:

Soliton solutions are sought in the form uxyz=Uxyeiμz where U(x, y) is real-valued function and μ is the propagation constant (or eigenvalue). Substituting the ansatz u(x, y, z), we get the following expressions:

uz=Ueiμz,uxx=Uxxeiμz,uyy=Uyyeiμz,u2=UeiμzUeiμz=U2,u4=U2U2=U4E4

where U=U in our case. Substituting the set of the terms in Eq. (4) into the (2+1) NLSM model, the following nonlinear equations for U are obtained

μU+12ΔU+βU2UρϕU+γU4UVU=0,ϕxx+vϕyy=U2xx.E5

Applying the Fourier transform to the eigenequations system (5) yields

μÛ12kx2+ky2Û+FβU2UρϕU+γU4UVU=0,kx2ϕ̂+vky2ϕ̂=kx2FU2,E6

where F denotes the Fourier transform, Û=FU,kx and ky are the Fourier transform variables. Isolating ϕ̂ from the second equation of Eq. (6) gives

ϕ̂=kx2FU2kx2+vky2.E7

Taking the inverse Fourier transform of Eq. (7), we get

ϕ=F1kx2FU2kx2+vky2,E8

where F1 denotes the inverse Fourier transform and during iteration, the first element of kx2+vky2 is set to 1 in order to avoid division by zero error. By applying the inverse Fourier transform to first equation of Eq. (6) and substituting Eq. (8) into the obtained equation, we get

μU+F112kx2+ky2Û+βU2UρF1kx2FU2kx2+vky2U+γU4UVU=0.E9

To obtain operator L0, Eq. (9) can be written as

L0U=F112kx2+ky2Û+T0U=0,E10

where

T0=μ+βU2ρF1kx2FU2kx2+vky2+γU4V.E11

Now, we should obtain operator L1, which denotes the linearized operator of L0U=0, with respect to the solution U, i.e., L0U+U=L1U+OU2, where U1. However, it should be noted that we have obtained the operator L0 by substituting the mean field term ϕxy into the governing equation. Therefore, at this point, we have to perturb ϕxy function as well. In accordance with this purpose, the soliton solution and the mean-field term should be perturbed as follows, respectively,

uxyz=Uxy+Uxyeiμzϕxy=ϕxy+ϕxy,E12

where U1 and ϕ1. Firstly, consider that Eq. (5) can be written as a general type of nonlinearities

μU+12ΔU+FU2UρϕUVU=0,ϕxx+vϕyy=U2xx,E13

where FU2=βU2+γU4 for cubic-quintic nonlinearity. Then, substituting perturbation U+U and using linear Taylor expansion yield

FU2=FU+U2=FU2+2UU+U2FU2+2UUFU2+2UUFU2U2,E14

where U=U,U=Uin our case and FU2=∂F/U2

FU2U=FU+U2U+UFU2+2UUFU2U2U+UU+UFU2+2U2UFU2U2+2UU2FU2U2U+UFU2+2U2UFU2U2+OU2.E15

Substituting perturbations in Eq. (12) and Eq. (15) into Eq. (13) and only terms of OU and Oϕ are retained, we get

μU+U+12ΔU+U+U+UFU2+2U2UFU2U2ρϕU+UρϕUVU+U=0,ϕxx+vϕyy+ϕxx+vϕyy=U2xx+2UUxx.E16

Substituting FU2=βU2+γU4FU2U2=β+2γU2 into Eq. (16) yields

μU+U+12ΔU+U+U+UβU2+γU4+2U2Uβ+2γU2ρϕU+UρϕUVU+U=0,ϕxx+vϕyy+ϕxx+vϕyy=U2xx+2UUxx.E17

Applying the Fourier transform to the first perturbed equation in Eq. (17) and the inverse Fourier transform to the obtained equation, we get

μU+U+F112kx2+ky2Û+Û+U+UβU2+γU4+2U2Uβ+2γU2ρϕU+UρϕUVU+U=0.E18

Using Eq. (5) and the second equation in Eq. (17), following equation is obtained

ϕxx+vϕyy=2UUxx.E19

Applying Fourier transform to Eq. (19) and isolating ϕ̂ from obtained equation, we get

kx2ϕ̂+vky2ϕ̂=kx2F2UUϕ̂=kx2F2UUkx2+vky2.E20

Taking the inverse Fourier transform of ϕ̂ yields

ϕ=F1kx2F2UUkx2+vky2.E21

Substituting Eq. (8) and Eq. (21) into Eq. (18) yields

μU+U+F112kx2+ky2Û+Û+U+UβU2+γU4+2U2Uβ+2γU2ρF1kx2FU2kx2+vky2U+UρF1kx2F2UUkx2+vky2UVU+U=0.E22

After grouping the terms, Eq. (22) can be written as

μU+F112kx2+ky2Û+βU2UρF1kx2FU2kx2+vky2U+γU4UVU+μU+F112kx2+ky2Û+βU2U+γU4U+2βU2U+4γU4UρF1kx2FU2kx2+vky2UρF1kx2F2UUkx2+vky2UVU=0.E23

Hence,

μU+F112kx2+ky2Û+βU2UρF1kx2FU2kx2+vky2U+γU4UVU+μU+F112kx2+ky2Û+3βU2UρF1kx2FU2kx2+vky2U+5γU4UρF1kx2F2UUkx2+vky2UVU=0.E24

From Eq. (9), we know that the first bracket is identically zero. Consequently, we obtain

μU+F112kx2+ky2Û+3βU2UρF1kx2FU2kx2+vky2U+5γU4UρF1kx2F2UUkx2+vky2UVU=0.E25

Moreover, Eq. (24) satisfied L0U+U=L1U+OU2. Therefore, to obtain a linearized operator L1L1, Eq. (25) can be written as

L1U=F112kx2+ky2Û+T1UρF1kx2F2UUkx2+vky2U=0,E26

where

T1=μ+3βU2ρF1kx2FU2kx2+vky2+5γU4V.E27

To obtain soliton solution U(x, y) in L0U=0, we numerically integrate the following distance-dependent squared-operator evolution equation

Uz=M1L1M1L0U,E28

where denotes the Hermitian of the operator and M is a real-valued positive definite Hermitian preconditioning operator that is introduced to accelerate the convergence. Since it is easily invertible to take the Fourier transform, we take the preconditioning operator M to be in the form of the following

M=cxx+yy,E29

where c > 0 is a parameter for parametrizing the numerical scheme. Applying the Fourier transform to Eq. (29) yields

FM=c+kx2+ky2.E30

Consequently, in Eq. (28)

M1L1M1L0U=F1FL1M1L0Uc+kx2+ky2.E31

Using the forward Euler method, steady-state solution U is computed by an iterative scheme as follows

Un+1=UnM1L1M1L0UU=UnΔz,E32

where Δz is an auxiliary distance-step parameter. It has been demonstrated that the SOM algorithm converges to a soliton solution for a wide range of nonlinear PDEs if the initial condition is sufficiently close to the exact solution and the distance-step Δz in the iteration scheme is less than a specific threshold value [60, 61]. To obtain a convergent soliton solution, c and Δz are chosen heuristically as positive real numbers. Moreover, our convergence criterion is that the obtained solution satisfies Eq. (10) with an absolute error less than 105.

In this chapter, to obtain dipole solitons, the initial condition of the SOM algorithm is chosen as a multi-humped Gaussian function which is given by

U0xy0=n=0H1eAx+xn2+y+yn2+iθn,E33

where xn and yn represent the location of the solitons on the lattice, H corresponds to the number of humps, A is a positive integer, and θn is the phase difference. Since we numerically investigate the dipole solitons, H is set to 2, thus Eq. (33) takes the following form:

U0xy0=eAx+x02+y+y02+iθ0+eAx+x12+y+y12+iθ1,E34

where (x0,y0) and (x1,y1) represent the locations of dipole solitons, θ0 and θ1 are the phase differences of dipole solitons. It was shown that the solitons located at the maximum of the lattices are unstable [13, 21, 24], due to this fact we will investigate the dipole solitons located on minima of the considered square lattice. A dipole (two-phased) localized soliton numerically found by

A=1,xn=rcosθn,yn=rsinθn,n=0,1.E35

Here r is set to be π and θn=, so that the humps of the initial condition are located at the local minima of the lattice where x0y0=π0 and x1y1=π0.

Unless otherwise stated, parameters in the NLSM model (2) are fixed to

μρvβγV0=0.10.51.520.112.5.E36

It is noted that ρ=0.5 and v=1.5 are especially chosen to simulate quadratic optical effects in potassium niobate (KNbO3) [30].

Dipole solitons of the NLSM model (2) are calculated by the SOM method. In Figure 2, 3D views (first column), phase structures (second column), and contour plots of dipole solitons on the underlying lattice (third column) are displayed for self-defocusing (γ<0) and self-focusing (γ>0) quintic nonlinearities. γ is set to be −0.3, −0.1 and 0.3 in Figure 2(a)(c), respectively, and all other parameters are fixed to the values given in Eq. (36). Figure 2 shows that the dipole solitons can be generated on the lattice minima (see the third column), and the amplitudes of dipole solitons are decreased as γ increased (from −0.3 to 0.3) (see the first column).

Figure 2.

3D dipole profiles centered at the lattice minima (first column), the phase structures of the dipole (second column), and the contour plot of the dipole solitons superimposed on the underlying lattice (third column), which are obtained for (a) γ=0.3,c=2.1,Δz=0.4 and error is order of 106, (b) γ=0.1,c=2.5,Δz=0.4 and error is order of 108, and (c) γ=0.3,c=2.5,Δz=0.4 and error is order of 108. All other parameters are fixed to the values in Eq. (36).

Advertisement

3. Stability analysis

The stability dynamics of dipole solitons obtained by the SOM method are studied by the power analysis and direct simulation of the nonlinear evolution.

The power of solitons plays an important role in the stability analysis and it is calculated by

Pμ=++U(xyμ)2dxdy.E37

Vakhitov and Kolokolov proved a necessary condition for the linear stability of solitons in Ref [62]. They demonstrated that a soliton is linearly stable only if its power increases as propagation constant (or eigenvalue) μ increases. In other words, a necessary condition for the stability of solitons is

dP>0.E38

Moreover, Weinstein and Rose [63, 64] proved that a necessary condition for the nonlinear stability of solitons is also the slope condition given in Eq. (38).

To analyze nonlinear stability of the NLSM model (2), we examine the direct simulation of dipole solitons obtained by the SOM method. A finite-difference discretization scheme is used in the spatial domain (x, y) and the dipole solitons are advanced in the z direction with a fourth-order Runge-Kutta method. The initial condition of the nonlinear evolution is taken to be a dipole soliton, and 1% random noise is inserted into the amplitude of the initial condition.

The power diagrams of dipole solitons are displayed for varied μ, γ, β and ρ values in Figure 3(a)(d), respectively. It is noted that the domain of existence for the varied parameter is shown on the x-axis of each panel when other parameters are fixed to the values in Eq. (36). Figure 3 shows that the power of dipole solitons increases as μ and ρ increase, whereas the power of dipole solitons decreases as γ and β increase. Moreover, the stability (solid blue) and instability (red dotted) regions of parameters are determined by the nonlinear evolution of dipole solitons for each point on the power curves.

Figure 3.

Power of dipole solitons (a) for varying eigenvalue μ, (b) for varying quintic nonlinearity coefficient γ, (c) for varying cubic nonlinearity coefficient β, and (d) for varying quadratic nonlinear response ρ. The nonlinear stability and instability regions are shown by solid blue and red dotted lines, respectively.

The dipole solitons are found to be nonlinearly stable for self-defocusing quintic nonlinearity (γ=0.1) when the power P ∈ [0.99, 1.87] and propagation constant μ0.750.6. Also, the dipole solitons are nonlinearly stable when P ∈ [3.12, 4.26] and μ0.350.06, which is the second nonlinearly stable gap (see Figure 3(a)). These results are consistent with key analytical results on nonlinear stability, which Weinstein and Rose proved in Ref [63, 64], since slope of the power-eigenvalue (P − μ) diagram is positive. As can be seen from Figure 3(b), the dipole solitons are obtained for γ0.725, when other parameters are fixed, and dipole solitons are nonlinearly stable for γ0.210.25. Zoom-in view of this stability domain is depicted in Figure 3(b). Furthermore, it is observed that dipole solitons are nonlinearly stable for β1.618.9 (see Figure 3(c)), and dipole solitons are stable for ρ00.8 (see Figure 3(d)) in their existence domains when other parameters are fixed.

In Figure 4, nonlinear evolution of peak amplitudes, 3D views of the evolved dipole solitons, and the phase structures of evolved dipole solitons are plotted for the dipole solitons that are shown in Figure 2. The effect of quintic nonlinearity (γ) on nonlinear stability is investigated by fixing other parameters as in Eq. (36).

Figure 4.

Nonlinear evolution of maximum amplitudes as a function of propagation distance z (first column), 3D views of the dipole solitons after evolution (second column), and the phase structures of dipole solitons after evolution (third column) for (a) γ=0.3, (b) γ=0.1, and (c) γ=0.3. All other parameters are taken as in Eq. (36).

Figure 4(b) shows that peak amplitudes of dipole solitons oscillate mildly (first column), and the 3D profile (second column) and phase structure (third column) of dipole solitons are preserved for γ=0.1. Thus, the dipole solitons are nonlinearly stable for the defocusing quintic nonlinearity for the considered parameter regime. On the other hand, as shown in Figure 4(a) and (c), when the quintic nonlinearity is strong (γ=0.3 and γ=+0.3), peak amplitudes of dipole solitons increase significantly in a short propagation distance z (first column), dipole profiles (second column) cannot be preserved, and phase structures of dipole solitons (third column) break up after evolution. Comparing Figure 4(a) and (c), it is observed that the propagation distance of dipole solitons in a medium with strong self-focusing nonlinearity (γ=0.3) is longer than that of a medium with strong self-defocusing nonlinearity (γ=0.3). Considering these evolution results in Figure 4 and the existing domain for γ in Figure 3(b), it is demonstrated that both strong self-focusing and self-defocusing quintic nonlinearities have a negative effect on the nonlinear stability of dipole solitons.

In previous studies [6, 14, 24], it is found that modification of the depth of potential can suppress nonlinear instabilities. More recently, Bağcı and coworkers [46] have demonstrated that nonlinear stability of fundamental solitons in an NLSM system (2) with quintic nonlinearity can be improved by the modification of lattice depth V0. They showed that increased lattice depth supports the stability of fundamental solitons in a medium with strong self-focusing (γ=0.3) quintic nonlinearity, and the stability of solitons in a medium with strong self-defocusing (γ=0.3) quintic nonlinearity can be improved by decreasing lattice depth. For the dipole solitons, evolution of peak amplitudes is depicted for varying potential depths, when γ=0.3 and γ=0.3 in Figure 5(a) and (b), respectively. Figure 5(a) shows that the stability of dipole solitons is improved by decreasing lattice depth (from 25 to 5) for strong self-defocusing nonlinearity (γ=0.3), and collapse can be arrested when V0=5. In contrast, as shown in Figure 5(b), the propagation distance of dipole solitons in a medium with strong self-focusing nonlinearity (γ=0.3) is extended by increasing lattice depth (from 5 to 50). It should be noted that these results are in agreement with the findings of the aforementioned studies. Thus modification of the lattice depth can be utilized to improve the nonlinear stability of dipole solitons.

Figure 5.

Maximum amplitudes of the evolved dipole solitons for varying depth of potential V0, when the dipole soliton is obtained for (a) γ=0.3 and (b) γ=0.3.

It is also known that when the quadratic [15, 24, 44] and quintic [46] electro-optic effects are strong, the instability of fundamental solitons can be improved by increasing the anisotropy parameter. To examine the effect of anisotropy coefficient v on the nonlinear stability of dipole solitons in a medium with strong quintic nonlinearity (γ=0.3 and γ=0.3), evolution of the peak amplitudes is displayed for varied v values in Figure 6. Figure 6 shows that increasing the anisotropy coefficient v from 0.001 to 10 stabilizes the dipole solitons in a medium with strong self-defocusing nonlinearity (γ=0.3), and increasing v from 0.001 to 1000 extends the propagation distance of dipole solitons in a medium with strong self-focusing nonlinearity (γ=0.3). Thus, larger anisotropy coefficient supports the nonlinear stability of the dipole solitons, and this result complies with the results of the previous studies [46]. It is important to note that the parameters ρ and v are predetermined coefficients that depend on the type of optical materials; larger values of v cannot be applied to real optical systems. In this chapter, the effect of extremely large v values on the stability of dipole solitons is explored numerically.

Figure 6.

Maximum amplitudes of the evolved dipole solitons for varying anisotropy coefficients v, when the dipole soliton is obtained for (a) γ=0.3, and (b) γ=0.3.

Advertisement

4. Conclusions

In this chapter, the existence and nonlinear stability dynamics of dipole solitons have been investigated for a nonlocal nonlinear medium with quintic nonlinear response. This medium was characterized by the (2+1)D NLSM system with a periodic external lattice. Dipole solitons were obtained for self-defocusing (γ<0) and self-focusing (γ>0) quintic nonlinearities by the SOM method, and the nonlinear stability of these dipole structures has been investigated by the direct simulation of the model equations. Power of dipole solitons was determined for varying μ, γ, β, and ρ parameters and it has shown that the power of dipole solitons increases as the eigenvalue μ and quadratic nonlinear response ρ increase, whereas the power of dipole solitons decreases as quintic nonlinearity coefficient γ and cubic nonlinearity coefficient β increase.

Nonlinear evolution of the dipole solitons showed that the dipole solitons are stable for the weak self-focusing and self-defocusing quintic nonlinearity. In other words, as an absolute value of γ increases, the obtained dipole solitons become nonlinearly unstable in both self-focusing and self-defocusing media. It has been demonstrated that the collapse of dipole solitons can be arrested by decreased potential depth in a medium with strong self-defocusing quintic nonlinearity (γ=0.3), while the deeper lattice extends the propagation distance of dipole solitons in a medium with strong self-focusing quintic nonlinearity (γ=0.3). Furthermore, it has been observed that increasing the anisotropy coefficient (v) extends the propagation distance of the dipole solitons for strong self-focusing quintic nonlinearity, and it stabilizes the dipole solitons for strong self-defocusing quintic nonlinearity.

In conclusion, the existence and stability properties of dipole solitons have been numerically explored in a nonlocal nonlinear medium with quintic nonlinear response, and it has been demonstrated that the instability of dipole solitons can be suppressed by modification of the lattice depth and increased anisotropy coefficient.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Ablowitz MJ. Nonlinear Dispersive Waves: Asymptotic Analysis and Solutions. Cambridge University Press; 2011
  2. 2. Stegeman GIA, Christodoulides DN, Segev M. Optical spatial solitons: historical perspectives. IEEE Journal of Selected Topics in Quantum Electronics. 2000;6(6):1419-1427. DOI: 10.1109/2944.902197
  3. 3. Yuri S, Agrawal GP. Optical Solitons: From Fibers to Photonic Crystals. Academic Press; 2003
  4. 4. Andersen DR, Allan GR, Skinner SR, Smirl AL. Observation of fundamental dark spatial solitons in semiconductors using picosecond pulses. Optics Letters. 1991;16(3):156-158. DOI: 10.1364/OL.16.000156
  5. 5. Fleischer JW, Segev M, Efremidis NK, Christodoulides DN. Observation of two-dimensional discrete solitons in optically induced nonlinear photonic lattices. Nature. 2003;422:147
  6. 6. Ablowitz MJ, Antar N, Bakırtaş İ, Ilan B. Band-gap boundaries and fundamental solitons in complex two-dimensional nonlinear lattices. Physical Review A. 2010;81:033834
  7. 7. Bağcı M. Impact of the lattice period on the stability dynamics of defect solitons in periodic lattices. Physical Review A. 2022;105:043524
  8. 8. Christodoulides DN, Yang J. Parity-Time Symmetry and Its Applications. Singapore: Springer; 2018
  9. 9. Yang J, Musslimani ZH. Fundamental and vortex solitons in a two-dimensional optical lattice. Optics Letters. 2003;28(21):2094-2096. DOI: 10.1364/OL.28.002094
  10. 10. Kartashov LTYV, Egorov AA, Christodoulides DN. Stable soliton complexes in two-dimensional photonic lattices. Optics Letters. 2004;29(16):1918-1920. DOI: 10.1364/OL.29.001918
  11. 11. Göksel İ, Bakırtaş İ, Antar N. Nonlinear lattice solitons in saturable media. Applied Mathematics and Information Sciences. 2014;9:377-385
  12. 12. Göksel İ, Antar N, Bakırtaş İ. Two-dimensional solitons in PT-symmetric optical media with competing nonlinearity. Optik. 2018;156:470-478
  13. 13. Mark J, Ilan B, Schonbrun E, Piestun R. Solitons in two-dimensional lattices possessing defects, dislocations, and quasicrystal structures. Physical Review E. 2006;74:035601
  14. 14. Ablowitz MJ, Antar N, Bakırtaş İ, Ilan B. Vortex and dipole solitons in complex two-dimensional nonlinear lattices. Physical Review A. 2012;86:033804
  15. 15. Bağcı M. Soliton dynamics in quadratic nonlinear media with two-dimensional pythagorean aperiodic lattices. Optics Letters. 2021;38:1276
  16. 16. Qidong F, Wang P, Huang C, Kartashov YV, Torner L, Konotop VV, et al. Optical soliton formation controlled by angle twisting in photonic moiré lattices. Nature Photonics. 2020;14(11):663-668
  17. 17. Huang C, Ye F, Chen X, Kartashov YV, Konotop VV, Torner L. Localization-delocalization wavepacket transition in pythagorean aperiodic potentials. Scientific Reports. 2016;6(1):32546
  18. 18. Wang P, Zheng Y, Chen X, Huang C, Kartashov YV, Torner L, et al. Localization and delocalization of light in photonic moiré lattices. Nature. 2020;577(7788):42-46
  19. 19. Bağcı M, Bakırtaş İ, Antar N. Fundamental solitons in parity-time symmetric lattice with a vacancy defect. Optical Communication. 2015;356:472-481
  20. 20. Bağcı M. Effects of lattice frequency on vacancy defect solitons in a medium with quadratic nonlinear response. Bitlis Eren Üniversitesi Fen Bilimleri Dergisi. 2022;11:344-351
  21. 21. Bağcı M, Bakırtaş İ, Antar N. Vortex and dipole solitons in lattices possessing defects and dislocations. Optical Communication. 2014;331:204-218
  22. 22. Martin H, Eugenieva ED, Chen Z, Christodoulides DN. Discrete solitons and soliton-induced dislocations in partially coherent photonic lattices. Physical Review Letters. 2004;92:123902
  23. 23. Mark J, Biondini G, Blair S. Localized multi-dimensional optical pulses in non-resonant quadratic materials. Mathematical Computational Simulation. 2001;56(6):511-519
  24. 24. Bağcı M, Bakırtaş İ, Antar N. Lattice solitons in nonlinear Schrödinger equation with coupling-to-a-mean-term. Optical Communication. 2017;383:330-340
  25. 25. Bağcı M, Kutz JN. Spatiotemporal mode locking in quadratic nonlinear media. Physical Review A. 2020;102:022205
  26. 26. Buryak AV, Di Trapani P, Skryabin DV, Trillo S. Optical solitons due to quadratic nonlinearities: from basic physics to futuristic applications. Physical Reports. 2002;370(2):63-235
  27. 27. Hayata K, Koshiba M. Multidimensional solitons in quadratic nonlinear media. Physical Review Letters. 1993;71(20):3275-3278. DOI: 10.1103/PhysRevLett.71.3275
  28. 28. Torner L, Sukhorukov AP. Quadratic solitons. Optical Photonic News. 2002;13(2):42-47
  29. 29. Torruellas WE, Wang Z, Hagan DJ, VanStryland EW, Stegeman GI, Torner L, et al. Observation of two-dimensional spatial solitary waves in a quadratic medium. Physical Review Letters. 1995;74:5036
  30. 30. Crasovan L-C, Torres JP, Mihalache D, Torner L. Arresting wave collapse by wave self-rectification. Physical Review Letters. 2003;91:063904
  31. 31. Schiek R, Pertsch T. Absolute measurement of the quadratic nonlinear susceptibility of lithium niobate in waveguides. Optical Material Express. 2012;2(2):126-139
  32. 32. Benney DJ, Roskes GJ. Wave instabilities. Studies in Application Mathematics. 1969;48:377-385
  33. 33. Davey A, Stewartson K. On three-dimensional packets of surface waves. Proceedings of the Royal Society A. 1974;338:101-110
  34. 34. Ablowitz MJ, Haberman R. Nonlinear evolution equations—two and three dimensions. Physical Review Letters. 1975;35:1185-1188
  35. 35. Djordjevic VD, Redekopp LG. On two-dimensional packets of capillary-gravity waves. Journal of Fluid Mechanics. 1977;79(4):703-714. DOI: 10.1017/S0022112077000408
  36. 36. Mark J, Biondini G, Blair S. Multi-dimensional pulse propagation in non-resonant materials. Physical Review Letters. 1997;236(5):520-524
  37. 37. Mark J, Biondini G, Blair S. Nonlinear Schrödinger equations with mean terms in nonresonant multidimensional quadratic materials. Physical Review. E. 2001;63:046605
  38. 38. Michael L. Sundheimer. Cascaded second-order nonlinearities in waveguides [PHD thesis]. 1994
  39. 39. Ablowitz MJ. Wave collapse in a class of nonlocal nonlinear Schrödinger equations. Physica D: Nonlinear Phenomena. 2005;207(3):230-253
  40. 40. Kelley PL. Self-focusing of optical beams. Physical Review Letters. 1965;15:1005-1008
  41. 41. Papanicolaou G, McLaughlin D, Weinstein M. Focusing singularity for the nonlinear Schrödinger equation. In: Fujita H, Lax PD, Strang G, editors. Nonlinear Partial Differential Equations in Applied Science; Proceedings of The U.S.-Japan Seminar, Tokyo, 1982, volume 81 of North-Holland Mathematics Studies. North-Holland; 1983
  42. 42. Merle FH, Raphael P. On universality of blow-up profile for l2 critical nonlinear schrödinger equation. Inventiones Mathematicae. 2004;156:565
  43. 43. Moll KD, Gaeta AL, Fibich G. Self-similar optical wave collapse: Observation of the townes profile. Physical Review Letters. 2003;90:203902
  44. 44. Bağcı M. Partially PT -symmetric lattice solitons in quadratic nonlinear media. Physical Review A. 2021;103:023530
  45. 45. Bağcı M. Vortex solitons on partially PT -symmetric azimuthal lattices in a medium with quadratic nonlinear response. Journal of Mathematical Sciences and Modelling. 2021;4:117-125
  46. 46. Bağcı M, Horikis TP, Bakırtaş İ, Antar N. Lattice solitons in a nonlocal nonlinear medium with self-focusing and self-defocusing quintic nonlinearity. In: Antar N, Bakırtaş İ, editors. The Nonlinear Schrödinger Equation. Rijeka: IntechOpen; 2022
  47. 47. Abo-Shaeer JR, Raman C, Vogels JM, Ketterle W. Observation of vortex lattices in bose-einstein condensates. Science. 2001;292:476-479
  48. 48. Matthews MR, Anderson BP, Haljan PC, Hall DS, Wieman CE, Cornell EA. Vortices in a bose-einstein condensate. Physical Review Letters. 1999;83:2498
  49. 49. Bartal G, Fleischer JW, Segev M, Manela O, Cohen O. Two-dimensional higher-band vortex lattice solitons. Optics Letters. 2004;29:2049
  50. 50. Fleischer JW, Bartal G, Cohen O, Manela O, Segev M, Hudock J, et al. Observation of vortex-ring “discrete” solitons in 2d photonic lattices. Physical Review Letters. 2004;92:3
  51. 51. Freedman B, Bartal G, Segev M, Lifshitz R, Christodoulides DN, Fleischer JW. Wave and defect dynamics in nonlinear photonic quasicrystals. Nature. 2006;440:7088
  52. 52. Kartashov YV, Malomed BA, Torner L. Solitons in nonlinear lattices. Reviews of Modern Physics. 2011;83:247-305
  53. 53. Yuri S. Dark optical solitons: physics and applications. Physics Reports. 1998;298:81-197
  54. 54. Leblond H, Malomed BA, Mihalache D. Spatiotemporal vortex solitons in hexagonal arrays of waveguides. Physical Review E. 2011;83:063825
  55. 55. Izdebskaya YV, Shvedov VG, Jung PS, Krolikowski W. Stable vortex soliton in nonlocal media with orientational nonlinearity. Optics Letters. 2018;43:66
  56. 56. Chen Y-F, Beckwitt K, Wise FW, Aitken BG, Sanghera JS, Aggarwal ID. Measurement of fifth- and seventh-order nonlinearities of glasses. Journal of Optical Society America B. 2006;23(2):347-352
  57. 57. Azzouzi F, Triki H, Grelu P. Dipole soliton solution for the homogeneous high-order nonlinear schrödinger equation with cubic-quintic-septic non-kerr terms. Applied Mathematical Modelling. 2015;39:3-1300
  58. 58. Komarov A, Leblond H, Sanchez F. Quintic complex ginzburg-landau model for ring fiber lasers. Physical Review E. 2005;72:025604
  59. 59. Alidou Mohamadou CG, Tiofack L, Kofané TC. Wave train generation of solitons in systems with higher-order nonlinearities. Physical Review E. 2010;82:016601
  60. 60. Yang J, Lakoba TI. Universally-convergent squared-operator iteration methods for solitary waves in general nonlinear wave equations. Studies in Applied Mathematics. 2007;118(2):153-197
  61. 61. Yang J. Nonlinear Waves in Integrable and Nonintegrable Systems. Philadelphia: SIAM; 2010
  62. 62. Vakhitov NG, Kolokolov AA. Stationary solutions of the wave equation in a medium with nonlinearity saturation. Radiophysics and Quantum Electronics. 1973;16(7):783-789
  63. 63. Harvey A, Weinstein MI. On the bound states of the nonlinear schrödinger equation with a linear potential. Physica D: Nonlinear Phenomena. 1988;30(1):207-218
  64. 64. Michael I. Weinstein, Modulational stability of ground states of nonlinear schrödinger equations. SIAM Journal on Mathematical Analysis. 1985;16(3):472-491

Written By

Mahmut Bağcı, Melis Turgut, Nalan Antar and İlkay Bakırtaş

Submitted: 17 May 2022 Reviewed: 30 June 2022 Published: 11 August 2022