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

Self-structuring of stable dissipative breathing vortex solitons in a colloidal nanosuspension

Open Access Open Access

Abstract

The self-structuring of laser light in an artificial optical medium composed of a colloidal suspension of nanoparticles is demonstrated using variational and numerical methods extended to dissipative systems. In such engineered materials, competing nonlinear susceptibilities are enhanced by the light induced migration of nanoparticles. The compensation of diffraction by competing focusing and defocusing nonlinearities, together with a balance between loss and gain, allow for self-organization of light and the formation of stable dissipative breathing vortex solitons. Due to their robustness, the breathers may be used for selective dynamic photonic tweezing of nanoparticles in colloidal nanosuspensions.

© 2017 Optical Society of America

1. Introduction

The structuring and controlling of laser light by its propagation through appropriate nonlinear media is the subject of great interest [1–7]. Such manipulation of laser beams and pulses requires engineering of appropriate artificial materials, like colloidal suspensions of nanoparticles exhibiting strong nonlinear response. The medium is modified by the laser propagation, causing in turn the self-organized localization and structuring of the light itself. The main issue in nonlinear propagation of localized optical structures is their stability. Solitons are localized structures self-generated far from thermodynamic equilibrium [8]. In conservative systems, one-dimensional (1D) optical solitons, described by the nonlinear Schrödinger equation (NSE), are self-trapped through the simultaneous balance of diffraction or dispersion by the self-focusing. In two dimensions, the catastrophic collapse of beams occurs that can be prevented by the presence of negative quintic nonlinearity [1,9]. The generation of 2D optical solitons in a cubic-quintic centrosymmetric medium has been recently demonstrated in an experiment [10].

However, in nature losses are always present, causing soliton annihilation in the absence of any gain. Complex Ginzburg-Landau equations that feature both gain and loss adequately describe the self-generation of dissipative solitons in nanophotonics, plasmonics, fluids, plasmas, and electromagnetism, as well as superconductivity, superfluidity, elementary particles, and biological systems [11,12]. In nonlinear media with gain and loss, multidimensional solitons self-stabilize due to the simultaneous balance of diffraction by competing cubic-quintic nonlinearity, as well as by the compensation of loss by gain [13]. Of particular interest for manipulation and tweezing of nanoparticles and nanorods in various colloidal suspensions are vortex solitons [14]. The intensity of vortices displays a volcano shape with zero electrical field in the center, corresponding to a singularity represented by the topological charge S [15–17]. The resulting nonzero angular momentum M of the vortex beam is proportional to the product of the charge and beam power P.

In this paper, we investigate the conditions for self-generation of dissipative vortex solitons in a colloidal suspension of nanoparticles. In particular, we focus on the stability of such structures when laser light redistributes nanoparticles, modifying the spatial distribution of an effective nonlinear susceptibility. The paper is organized as follows. In Sec. 2, theoretical analysis of the colloidal model is provided and an appropriate generalized complex cubic-quintic Ginzburg-Landau equation introduced. In Sec. 3, the variational approach, extended to dissipative systems, is applied to the equation and the stability analysis of obtained solutions carried through. A stability domain is established in the parameter space. In Sec. 4, numerical simulation of the input beams from this domain is carried out, to confirm their stability. Section 5 brings conclusions.

2. Theoretical analysis

We consider here the nonlinear response of a colloidal suspension of nanoparticles, originating from the gradient force exerted by light on the particles. This force causes particles to either agglomerate in high light intensity regions or disperse away from these regions, depending on whether the polarizability of particles is positive or negative. In either case, the effective refractive index seen by the light beam always increases with intensity, leading to self-focusing and ultimately to the formation of spatial solitons [14, 18–24]. The polarizability of particles with refractive index np immersed in a transparent liquid with refractive index nb is given by

α=3Vpε0nb2θ,
where Vp is the particle volume, ε0 is the free space permittivity, and θ=(np2/nb21)/(np2/nb2+2) is the relative refractive index. The gradient force exerted by light on particles determines the effective packing fraction of the particles η = ρVp, where ρ is the concentration of nanoparticles [18–20]. In particular, it has been shown that in the hard sphere approximation for particle-particle interaction, this packing fraction satisfies the following relation:
α(4kBT)1|U(η)|2=sgn(θ)[g(η)g(η0)],
where |U|2 represents the light intensity, g(η) = (3 − η)/(1 − η)3 + ln η, and sgn(θ) denotes the sign function [18–20]. Note that η0 is the packing fraction in the absence of laser field. In order to follow the evolution of electrical field with amplitude U and to demonstrate the self-generation of stable dissipative solitons, we consider the paraxial approximation of the Helmholtz equation
2U+k02neff2U=0,
where k0 is the free space wavevector and neff is the effective nonlinear refractive index change caused by nanoparticles moving under the influence of laser beam. Assuming a relatively small index contrast between nanoparticles and the background, neff2=nb2(1+3θη). Consequently, the slowly varying electric field envelope E(r) is defined by the expression [19]
U(r)(3Vpε0nb2|θ|/4kBT)1/2=E(r)exp[ik0nb(1+3θη0)1/2Z],
where Z stands for the propagation variable. Therefore, the intensity is expressed as
|E|2=sgn(θ)[g(η)g(η0)].

The corresponding paraxial equation reads

iEz+γΔE+sgn(θ)(ηη0)E+iΓηE=0,
where z = 3 |θ| k0nb Z/2, γ=1/(3|θ|k02nb2) and Γ represents the linear scattering losses [19]. In Fig. 1, we illustrate the dependence of sgn (θ) (ηη0) on the light intensity |E|2 for a few values of η0, which corresponds to the packing fraction in the absence of light. After a careful fitting procedure, this curve is represented as the following polynomial relation:
sgn(θ)(ηη0)=(κ+σ|E|2ν|E|4)
with appropriate values of the parameters. For instance, when η0 = 0.01, the coefficients are κ = 0.024722, σ = 0.0035944, and ν = 0.00024943, and the fitting curve is basically indistinguishable from the original one. Recent experimental studies of centrosymmetric colloidal suspensions of metallic particles have confirmed the presence of nonlinear susceptibilities of odd orders, with alternating signs [25]. Indeed, it is well known that the polarization of centrosymmetric media corresponds to the series of only odd order nonlinear susceptibilities [1,11,12]. To allow for the formation of localized structures, the unavoidable presence of losses (having the negative sign) has to be counteracted by gain [1,11–13]. The gain could be introduced, for instance, by adding a dye such as rhodamine B into the colloidal suspension and then pump it with an external light source, to achieve amplification [26]. Consequently, it has become clear that the proper description of self-organization and soliton formation in colloidal suspensions must include a simultaneous balance of diffraction by competing nonlinearities, as well as the compensation of linear and nonlinear losses by gain. All of these conditions are met in the complex cubic-quintic Ginzburg-Landau equation (CQGLE).

 figure: Fig. 1

Fig. 1 Illustrating the dependence of packing fraction sgn (θ) (ηη0) on light intensity |E|2, for few values of the background packing fraction η0. Solid squares represent the fitting formula, Eq. (7), for the case η0 = 0.01

Download Full Size | PDF

Therefore, in what follows we investigate the nonlinear beam propagation in colloidal media using as a model the generalized CQGLE that takes into account various aspects of the particle-particle interaction:

iEz+γΔE+(κ+σ|E|2ν|E|4)E=iδE+iε|E|2Eiμ|E|4E+iβΔE.
Here, positively defined parameters γ, κ, σ, ν, δ, ε, μ, and β characterize respectively diffraction, linear term, cubic self-focusing, quintic self-defocusing, linear loss, cubic gain, quintic loss, as well as field diffusion. They depend of the chosen type of nanoparticles and the background medium. The variational method, extended to dissipative systems, is used to determine the domain of these dissipative parameters where the stable solitons can exist [13].

3. Stability criterion

The generalized complex cubic-quintic Ginzburg-Landau equation describes well the evolution of Gaussian laser beams in colloidal nanosuspensions. However, this equation does not admit exact solutions, except for a particular set of parameters [27,28]. Therefore, it is necessary to resort to numerical simulations. Nevertheless, an analytical approximation that provides clues for finding dissipative solitons in the numerical form has been developed using the variational approach (VA) adapted for dissipative systems [13,16,17,29].

The Lagrangian corresponding to Eq. (8) reads as

L=i2(E*zEEzE*)+γ|E|2κ|E|2σ2|E|4+ν3|E|6.
Following an optimization procedure of the extended VA, a trial function in the form of a Gaussian beam is chosen,
E=BA(rGR)Sexp[r2G2R2+iCG2r2+iSφ+iΨ],
where the azimuthal coordinate is φ, and A, R, C, S, and Ψ represent the amplitude, radius, wavefront curvature, topological charge and the phase of the beam, respectivelly. To start with, we consider the topological charge S = 1, with the scaling factors G = 16/9 and B=(92/8)γ/σ Using the VA for dissipative systems introduced in Ref. [13], the four variational parameters, the amplitude, radius, wavefront curvature, and phase – all functions of the propagation variable z – are optimized following Hamilton’s principle. The extremum function E (see Eq. (10)) renders the Lagrangian integral stationary, under the condition that the four Euler-Lagrange equations corresponding to Eq. (8) are valid. The variation of Lagrangian in Eq. (9) with respect to the phase Ψ yields the first Euler-Lagrange equation:
G24γ(R2dAdz+ARdRdz)+A[R2(δ02εσA2+3μγ2σ2A4)+βγ(1+4C2R4)]=0.
The optimization of amplitude A gives the second relation:
G22γdψdz+R2(4C2+G2γdCdz)+1R2κ02A2+32νγσ2A4=0.
The variation with respect to the radius R yields the third equation:
G22γdψdz+2R2(4C2+G2γdCdz)κ0A2+νγ2σ2A4+4βγC=0.
The last Euler-Lagrange equation corresponds to the optimized curvature C:
G2γ(R2dAdz+2ARdRdz)+A[R2(4C+2δ03εσA2+2μγσ2A4)+βγ(1+12C2R4)]=0,
where δ0 = G2δ/(2γ) and κ0 = G2κ/(2γ). The laser beam power P = 2π ∫ |u|2 rdr = πS!B2G2 A2R2, angular momentum M=2πIm(u*uφ)dr, and vortex width L=(2π/P)|u|2r3dr are no more conserved in the dissipative systems, as opposed to the conservative ones. Following the variational method for dissipative systems, after some algebra a set of four ordinary differential equations (ODEs) is obtained from Eqs. (1114) [13,16,30]:
dAdz=γAG2[5εσA24(C+μγσ2A4+βγR2C2)3βγR22δ0],
dRdz=γRG2(4CεσA2+μγσ2A4+βγR24βγC2R2),
dCdz=γG2(1R44C2A2R2+νγσ2A4R24βγCR2),
dψdz=γG2(κ02R2+3A252νγσ2A4+4βγC).
To simplify the problem a bit, we consider the approximation β << 1. The stationary solutions of the set of ODEs, Eqs. (1517), are obtained by solving them for the vanishing derivatives of the amplitude, width, and curvature. The fixed points of these equations correspond to the steady state solutions of Eq. (8) in the variational approximation, with the wavefront curvature C = 0.25[(ε/σβ/γ)A2−(μγβν)A4/σ2](1+ β2/γ2)−1 and the radius R = (A2 − (νγ/σ2)A4)−1/2. Concerning the amplitude, the family of solutions reduces, in this dissipative case, to a fixed double solution for a given set of dissipative parameters. Indeed, the amplitude in a steady state solution of Eqs. (1517) has two discrete values A+ and A,
A±=(2ε/σβ/γ)±(2ε/σβ/γ)22δ0(3μγ/σ22βν/σ2)(3μγ/σ22βν/σ2).
According to the general principles of the analysis of dissipative systems, A+ solution may be stable, while A is always unstable [1, 12, 13]. The bifurcation point is given by ε=σ[β/γ+2δ0/σ2(3μγ2βν)]/2. In Fig. 2, the upper stable branch corresponds to the A+ solution, while the lower unstable one represents the A solution. The former solution satisfies the condition C < 0 that is necessary for the self-organization of dissipative vortex solitons, based on the simultaneous cross-compensation between loss and gain, as well as the competing cubic-quintic nonlinearity that exceeds diffraction [13,29,30]. A steady state solution of the above set of ODEs, Eqs. (1517) is stable in Lyapunov sense if and only if all the eigenvalues λ of the Jacobian
λ3+α1λ2+α2λ+α3=0
have negative real parts [13,29,30]. Therefore, the Hurwitz coefficients of Eq. (20) have to be all positive in order for a fixed point to be in the stable dynamic equilibrium,
α1=γG2[6βγR2+16μγσ2A410εσA2+8C(1+βγCR2)]>0,
α2=γ2G48R4[1+β2γ2A2R2(1+6βεγσ9βμσ2A22νγσ2A2)2CA2R4(βγ+5εσ8μγσ2A22βνσ2A2)+4C2R4(1+β2γ23βεγσA2R2+5βμσ2A4R2)]>0,
α3=γ3G632A2R4[βγ2εσ+β2γ2εσ+A2(3μγσ2+β2μγσ22βνσ2)4CR2βγ(εσβγ2μγσ2A2+2βνσ2A2)4C2R4(2εσ+βγ+3β2γ2εσ3μγσ2A25β2μγσ2A22βνσ2A2)]>0,
and
α123=α1α2α3>0.
Such a stability criterion determines the stable fixed points for a large domain of dissipative parameters. We established that their stability crucially depends on the dimensionless nonlinear gain parameter ε and the nonlinear loss parameter μ. Therefore, the domain of stable steady state solutions is charted as a function of those two parameters in Fig. 3. Following our exhaustive numerical simulations, the variation of the remaining two dissipative coefficients associated with the linear loss (δ) and field diffusion (β) does not entail essential changes. Consequently, main results are adequately represented by fixing these parameters to δ = 0.01, and β = 0.05. Also, the diffusion parameter is fixed at γ = 0.0034 in accordance with the parameters κ, σ and ν fixed earlier. The realistic numerical values of dissipative parameters can be determined experimentally by the choice of laser beam intensity and the active background medium, e.g. by varying the dye concentration. Indeed, in the domain of random lasers, dye-solutions containing TiO2 or silica nanoparticles were employed for the observation of laser-like emission [26,31]. Up-conversion random laser emission was realized in various colloidal dye solutions [32–34].

 figure: Fig. 2

Fig. 2 Illustrating the stability of the beam amplitude A as a function of the nonlinear gain ε. The upper (solid line) branch corresponds to the stable A+ solution and the lower (dashed line) branch corresponds to the unstable A solution.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Stability domain (in yellow) produced by the VA-generated fixed points, in the plane of the nonlinear-gain strength ε and the nonlinear-loss strength μ (both dimensionless). The stability of vortex solitons is confirmed by direct simulations of Eq. (8) for parameters inside the region delimited by dashed lines.

Download Full Size | PDF

4. Numerical self-generation of stable dissipative vortex solitons

In section 3 we demonstrated that, following our variational method extended to dissipative systems, for each set of dissipative parameters, plotted in Fig. 3, only one stable steady state solution exists, as opposed to the family of NLS solitons [12, 13, 16, 29, 30]. Each such fixed point solution can be used as an input for numerical solution of Eq. (8), in order to confirm its solitonic nature [13,16,29,30]. The amplitude A, power P, and angular momentum M are recorded after each propagation step, from z = 0 until z = 21500 (see Fig. 4 for a typical choice of dissipative parameters, ε = 0.38 and μ = 2). Only after a long evolution (about 7000 steps in Figs. 4 and 5) the amplitude, power, and angular momentum start to perform steady state oscillations, while the width L remains practically constant. Hence, a breathing vortex soliton gets self-organized and remains stable during the subsequent evolution. Nanoparticles with positive polarizability are attracted by the laser field of volcano shape. During numerical evolution, nanoparticles are partially filling the crater, as can be seen from Fig. 5. Simultaneously, particles are climbing Gaussian hill from outside, enlarging the plateau of the “millstone” in Fig. 5. The particles concentration growth induces the nonlinear susceptibility increase. Consequently, the self-focusing induced by the nonlinear refractive index causes amplitude growth that would lead to catastrophic collapse in the absence of losses (see the amplitude behavior in Figs. 4 and 6). Linear and nonlinear losses, together with the quintic self-defocusing and diffraction, saturate this growth. This is a slow cyclic process with the breathing period of T = 240 propagation steps. Indeed, after this period the amplitude, power, and angular momentum suddenly start decreasing (see Fig. 6). This would annihilate the soliton without the refluxing effect of the gain that restores the amplitude and consequently, the power and angular momentum, initiating a new breathing cycle. This regular breathing behavior (see Figs. 46) is the signature of breathing vortex soliton self-trapping. The regular modulation of the breather over the period T = 240 is shown in Fig. 7 (see also [16,29]). At the amplitude maximum (z ≈ 20160), the vortex resembles a “millstone” with sharp edges. An abrupt fall of the amplitude and quasi-immediate recovery at z = 20170 transforms the “millstone” into a dome. During amplitude saturation (until z ≈ 20220), the edges become sharper. At z = 20400 the vortex resumes its pre-cycle shape. This robust cyclic behavior confirms that such a vortex with constant width is really a breathing soliton. Exhaustive numerical simulations allowed us to establish an area between dashed lines in Fig. 3 where the stability of dissipative vortex solitons is confirmed.

 figure: Fig. 4

Fig. 4 Amplitude A, power P, and angular momentum M monitored after each propagation step, from z = 0 to z = 21500 (for dissipative parameters ε = 0.38 and μ = 2). The ratio M/P corresponds to the constant value of S = 1.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Development of a stable breathing vortex with S = 1. Initial vortex (at z = 0) corresponds to the input Gaussian with parameters β = 0.05, δ = 0.01, μ = 2, and ε = 0.38 from the stability domain. Breathing vortex soliton is self-generated after z = 7000 steps. It remains stable after z = 21000 steps.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Regular breathing behavior of the beam characteristics. It corresponds to the vortex soliton self-organization. Different scales are used for amplitude A, power P, angular momentum M, and width L.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Robust cyclic behavior with period T = 240 from a “millstone” at z = 21160 through various domes, to the same “millstone” at z = 21400, confirming the stability of a breathing vortex soliton.

Download Full Size | PDF

For larger gain parameter ε, the breathing vortex soliton self-organizes more rapidly. For instance, for gain and loss parameters ε = 1.8 and μ = 3 in Fig. 8, the breathing with considerably shorter period TA = 4 starts already after 500 steps. The breather modulation from a “millstone” into a dome and back is much shorter. Figure 9 displays the perfect correspondence between domes and “millstones” in two cycles, distant for 20 steps. Such a rapid breathing does not alter the remarkable stability of dissipative vortex solitons. However, a new superimposed beat of period TB ≈ 500 appears at double the breathing period (half the frequency) for the smaller parameter ε (see M evolution in Fig. 8). The same results can be obtained for the negative polarizability nanoparticles that escape laser field and in such a way also increase the nonlinear susceptibility, contributing to the self-organization of dissipative vortex solitons. Such a self-structuring of the laser light in colloidal nanosuspensions can be applied for optical manipulation and selective dynamic tweezing of nanoparticles and nanorods [35,36], as well as for all-optical switching and signal processing [37].

 figure: Fig. 8

Fig. 8 Similar to Fig. 4 but for larger gain and loss parameters, ε = 1.8 and μ = 3. The amplitude, power, and angular momentum oscillate with a short period TA = 4 and a superimposed beat period, TB = 500, as a double breather. The breathing with shorter period TA = 4 starts already after 500 steps.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Perfect correspondence between domes and “millstones” appearing successively after TA + 1 = 5, in two cycles distant for 20 steps.

Download Full Size | PDF

5. Conclusions

We use analytical and numerical approaches to study the localization of light in a model of artificial media. Such models include metal-dielectric nanocomposites formed as colloidal suspensions of either gold or silver nanoparticles in liquids, polymers, and liquid crystals. These media can be suitably engineered, so as to enhance the competing nonlinear susceptibilities by the migration of nanoparticles toward the laser field or away from it. Independent of their polarizability, the nanoparticles move in such a way that the effective nonlinearity inside the beam always increases. We used the model of interacting hard-sphere nanoparticles, as well as experimental results from Ref. [25], to demonstrate higher-order nonlinear susceptibility in nanosuspensions and to corroborate the validity of the CQCGL equation, which describes the self-structuring of laser light into vortex solitons in such media. The corresponding generalized Ginzburg-Landau model exhibits overcompensation of diffraction by the saturating cubic-quintic nonlinearity, expressed analytically and numerically by the negative wave front curvature. This effective self-focusing would cause catastrophic collapse in the absence of losses. The amplification coming from the medium enhances laser field, preventing the soliton annihilation. Consequently, the self-focusing again dominates diffraction, restarting a new self-organization cycle. Therefore, the simultaneous cross-compensation between loss and gain, and the competing cubic-quintic nonlinearity exceeding diffraction, allows self-structuring of laser light into novel practically dissipationless breathing vortex solitons. The established breathers may act as robust dynamic tweezers of nanoparticles, controlling in real time the spatial distribution of effective nonlinear susceptibility in metal-dielectric nanocomposites and allowing for engineering of reconfigurable guides and smart grids.

Funding

National Priorities Research Program Grants No. 6-021-1-005 and No. 9-020-1-006 from the Qatar National Research Fund (a member of Qatar Foundation). Work at the Institute of Physics Belgrade is supported by the Ministry of Education and Science of the Republic of Serbia, under Project OI 171006.

Acknowledgments

This publication was made possible by the National Priorities Research Program Grants No. 6-021-1-005 and No. 9-020-1-006 from the Qatar National Research Fund (a member of Qatar Foundation). MRB acknowledges support from the Al Sraiya Holding Group.

References and links

1. Y. S. Kivshar and G.P. Agrawal, Optical Solitons: From Fibers to Photonic Crystals (Academic, 2003).

2. D. Pierangeli, F. Di Mei, C. Conti, A. J. Agranat, and E. DelRe, “Spatial rogue waves in photorefractive ferroelectrics,” Phys. Rev. Lett. 115, 093901 (2015). [CrossRef]   [PubMed]  

3. E. DelRe, E. Spinozzi, A. J. Agranat, and C. Conti, “Scale-free optics and diffractionless waves in nanodisordered ferroelectrics,” Nat. Photon. 5, 39–42 (2011). [CrossRef]  

4. A. Armaroli, S. Trillo, and A. Fratalocchi, “Suppression of transverse instabilities of dark solitons and their dispersive shock waves,” Phys. Rev. A 80, 053803 (2009). [CrossRef]  

5. S. Roy, A. Marini, and F. Biancalana, “Free-carrier-driven spatiotemporal dynamics in amplifying silicon waveguides,” Phys. Rev. A 89, 053827 (2014). [CrossRef]  

6. N. Ghofraniha, C. Conti, G. Ruocco, and S. Trillo, “Shocks in nonlocal media,” Phys. Rev. Lett. 99, 043903 (2007). [CrossRef]   [PubMed]  

7. C. Conti, A. Fratalocchi, M. Peccianti, G. Ruocco, and S. Trillo, “Observation of a gradient catastrophe generating solitons,” Phys. Rev. Lett. 102, 083902 (2009). [CrossRef]   [PubMed]  

8. G. Nicolis and I. Prigogine, Self-organization in Nonequilibrium Systems (Wiley, 1997).

9. V. Skarka, V. I. Berezhiani, and R. Miklaszewski, “Spatiotemporal soliton propagation in saturating nonlinear optical media,” Phys. Rev. E 56, 1080 (1997). [CrossRef]  

10. E. L. Falcao-Filho, C. B. de Araújo, G. Boudebs, H. Leblond, and V. Skarka, “Robust two-dimensional spatial soitons in liqid carbon disulfide,” Phys. Rev. Lett. 110, 013901 (2013). [CrossRef]  

11. I. S. Aranson and L. Kramer, “The world of the complex Ginzburg-Landau equation,” Rev. Mod. Phys. 74, 99 (2002). [CrossRef]  

12. N. N. Akhmediev and A. Ankiewicz, Dissipative Solitons: From Optics to Biology and Medicine, Lect. Notes Phys. 751 (Springer, 2008).

13. V. Skarka and N.B. Aleksić, “Stability criterion for dissipative soliton solutions of the one-, two-, and three-dimesional complex cubic-quintic Ginzburg-Landau equations,” Phys. Rev. Lett. 96, 013903 (2006). [CrossRef]  

14. A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu, “Observation of a single-beam gradient force optical trap for dielectric particles,” Opt. Lett. 11, 288 (1986). [CrossRef]   [PubMed]  

15. V. I. Berezhiani, V. Skarka, and N. B. Aleksić, “Dynamics of localized and nonlocalized optical vortex solitons in cubic-quintic nonlinear media,” Phys. Rev. E 64, 057601 (2001). [CrossRef]  

16. V. Skarka, N. B. Aleksić, H. Leblond, B. A. Malomed, and D. Mihalache, “Varieties of stable vortical solitons in Ginzburg-Landau media with radially inhomogeneous losses,” Phys. Rev. Lett. 105, 213901 (2010). [CrossRef]  

17. B. N. Aleksic, N. B. Aleksic, V. Skarka, and M. Belic, “Stability and nesting of dissipative vortex solitons with high vorticity,” Phys. Rev. A 91, 043832 (2015). [CrossRef]  

18. M. Matuszewski, W. Krolikowski, and Y.S. Kivshar, “Spatial solitons and light-induced instabilities in colloidal media,” Opt. Express 16, 1371–1376 (2008). [CrossRef]   [PubMed]  

19. M. Matuszewski, W. Krolikowski, and Y. Kivshar, “Soliton interactions and transformations in colloidal media,” Phys. Rev. A 79, 023814 (2009). [CrossRef]  

20. M. Matuszewski, “Engineering optical soliton bistability in colloidal media,” Phys. Rev. A 81, 013820 (2010). [CrossRef]  

21. R. El-Ganainy, D. N. Christodoulides, C. Rotschild, and M. Segev, “Soliton dynamics and self-induced transparency in nonlinear nanosuspensions,” Opt. Express 15, 10207–10218 (2007). [CrossRef]   [PubMed]  

22. R. El-Ganainy, D. N. Christodoulides, E. M. Wright, W. M. Lee, and K. Dholakia, “Nonlinear optical dynamics in nonideal gases of interacting colloidal nanoparticles,” Phys. Rev. A 80, 053805 (2009). [CrossRef]  

23. W. Man, S. Fardad, Z. Zhang, J. Prakash, M. Lau, P. Zhang, M. Heinrich, D. N. Christodoulides, and Z. Chen, “Optical nonlinearities and enhanced light transmission in soft-matter systems with tunable polarizabilities,” Phys. Rev. Lett. 111, 218302 (2013). [CrossRef]   [PubMed]  

24. S. Fardad, A. Salandrino, M. Heinrich, P. Zhang, Z. Chen, and D. N. Christodoulides, “Plasmonic resonant solitons in metallic nanosuspensions,” Nano Lett. , 14, 2498–2504 (2014). [CrossRef]   [PubMed]  

25. Albert S. Reyna and Cid B. de Araújo, “Spatial phase modulation due to quintic and septic nonlinearities in metallic colloids,” Opt. Express 22(19) 22456–22469 (2014). [CrossRef]   [PubMed]  

26. A. M. Brito-Silva, A. Galembeck, A. S. L. Gomes, A. J. Jesus-Silva, and C. B. de Araújo, “Random laser action in dye solutions containing Stöber silica nanoparticles,” J. Appl. Phys. 108, 033508 (2010). [CrossRef]  

27. N. N. Akhmediev, V. V. Afanasjev, and J. M. Soto-Crespo, “Singularities and special soliton solutions of the cubic-quintic complex Ginzburg-Landau equation,” Phys. Rev. E 53, 1190–1201 (1996). [CrossRef]  

28. J. Atai and B. A. Malomed, “Exact stable pulses in asymmetric linearly coupled Ginzburg–Landau equations,” Phys. Lett. A 246, 412–422 (1998). [CrossRef]  

29. V. Skarka, N. B. Aleksić, M. Lekić, B. N. Aleksić, B. A. Malomed, D. Mihalache, and H. Leblond, “Formation of complex two-dimensional dissipative solitons via spontaneous symmetry breaking,” Phys. Rev. A 90 (2), 023845 (2014). [CrossRef]  

30. V. Skarka, D. V. Timotijević, and N. B. Aleksić, “Extension of the stability criterion for dissipative optical soliton solutions of a two-dimensional Ginzburg–Landau system generated from asymmetric inputs,” J. Opt. A: Pure Appl. Opt. 10, 075102 (2008). [CrossRef]  

31. N. M. Lawandy, R. M. Balachandran, A. S. L. Gomes, and E. Sauvain, “Laser action in strongly scattering media,” Nature 368(6470), 436–438 (1994). [CrossRef]  

32. G. Zacharakis, N. A. Papadogiannis, and T. G. Papazoglou, “Random lasing following two-photon excitation of highly scattering media,” Appl. Phys. Lett. 81, 2511–2513 (2002). [CrossRef]  

33. J. Chen, H. Mizuno, H. Kawano, A. Miyawaki, and K. Midorikawa, “Two-photon pumping of random lasers by picosecond and nanosecond lasers,” Appl. Phys. B 85, 45–48 (2006). [CrossRef]  

34. A. S. L. Gomes, M. T. Carvalho, C. T. Dominguez, C. B. de Araújo, and P. N. Prasad, “Direct three-photon excitation of upconversion random laser emission in a weakly scattering organic colloidal system,” Opt. Express 22, 14305–14310 (2014). [CrossRef]   [PubMed]  

35. D. Grier, “A revolution in optical manipulation,” Nature 424, 810–816 (2003). [CrossRef]   [PubMed]  

36. V. Shvedov, A. R. Davoyan, C. Hnatovsky, N. Engheta, and W. Krolikowski, “A long-range polarization-controlled optical tractor beam,” Nat. Photon. 8, 846–850 (2014). [CrossRef]  

37. Albert S. Reyna and Cid B. de Araújo, “An optimization procedure for the design of all-optical switches based on metal-dielectric nanocomposites,” Opt. Express 237659–7666 (2015). [CrossRef]   [PubMed]  

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 (9)

Fig. 1
Fig. 1 Illustrating the dependence of packing fraction sgn (θ) (ηη0) on light intensity |E|2, for few values of the background packing fraction η0. Solid squares represent the fitting formula, Eq. (7), for the case η0 = 0.01
Fig. 2
Fig. 2 Illustrating the stability of the beam amplitude A as a function of the nonlinear gain ε. The upper (solid line) branch corresponds to the stable A+ solution and the lower (dashed line) branch corresponds to the unstable A solution.
Fig. 3
Fig. 3 Stability domain (in yellow) produced by the VA-generated fixed points, in the plane of the nonlinear-gain strength ε and the nonlinear-loss strength μ (both dimensionless). The stability of vortex solitons is confirmed by direct simulations of Eq. (8) for parameters inside the region delimited by dashed lines.
Fig. 4
Fig. 4 Amplitude A, power P, and angular momentum M monitored after each propagation step, from z = 0 to z = 21500 (for dissipative parameters ε = 0.38 and μ = 2). The ratio M/P corresponds to the constant value of S = 1.
Fig. 5
Fig. 5 Development of a stable breathing vortex with S = 1. Initial vortex (at z = 0) corresponds to the input Gaussian with parameters β = 0.05, δ = 0.01, μ = 2, and ε = 0.38 from the stability domain. Breathing vortex soliton is self-generated after z = 7000 steps. It remains stable after z = 21000 steps.
Fig. 6
Fig. 6 Regular breathing behavior of the beam characteristics. It corresponds to the vortex soliton self-organization. Different scales are used for amplitude A, power P, angular momentum M, and width L.
Fig. 7
Fig. 7 Robust cyclic behavior with period T = 240 from a “millstone” at z = 21160 through various domes, to the same “millstone” at z = 21400, confirming the stability of a breathing vortex soliton.
Fig. 8
Fig. 8 Similar to Fig. 4 but for larger gain and loss parameters, ε = 1.8 and μ = 3. The amplitude, power, and angular momentum oscillate with a short period TA = 4 and a superimposed beat period, TB = 500, as a double breather. The breathing with shorter period TA = 4 starts already after 500 steps.
Fig. 9
Fig. 9 Perfect correspondence between domes and “millstones” appearing successively after TA + 1 = 5, in two cycles distant for 20 steps.

Equations (24)

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

α = 3 V p ε 0 n b 2 θ ,
α ( 4 k B T ) 1 | U ( η ) | 2 = sgn ( θ ) [ g ( η ) g ( η 0 ) ] ,
2 U + k 0 2 n eff 2 U = 0 ,
U ( r ) ( 3 V p ε 0 n b 2 | θ | / 4 k B T ) 1 / 2 = E ( r ) exp [ i k 0 n b ( 1 + 3 θ η 0 ) 1 / 2 Z ] ,
| E | 2 = sgn ( θ ) [ g ( η ) g ( η 0 ) ] .
i E z + γ Δ E + sgn ( θ ) ( η η 0 ) E + i Γ η E = 0 ,
sgn ( θ ) ( η η 0 ) = ( κ + σ | E | 2 ν | E | 4 )
i E z + γ Δ E + ( κ + σ | E | 2 ν | E | 4 ) E = i δ E + i ε | E | 2 E i μ | E | 4 E + i β Δ E .
L = i 2 ( E * z E E z E * ) + γ | E | 2 κ | E | 2 σ 2 | E | 4 + ν 3 | E | 6 .
E = B A ( r G R ) S exp [ r 2 G 2 R 2 + i C G 2 r 2 + i S φ + i Ψ ] ,
G 2 4 γ ( R 2 d A d z + A R d R d z ) + A [ R 2 ( δ 0 2 ε σ A 2 + 3 μ γ 2 σ 2 A 4 ) + β γ ( 1 + 4 C 2 R 4 ) ] = 0 .
G 2 2 γ d ψ d z + R 2 ( 4 C 2 + G 2 γ d C d z ) + 1 R 2 κ 0 2 A 2 + 3 2 ν γ σ 2 A 4 = 0 .
G 2 2 γ d ψ d z + 2 R 2 ( 4 C 2 + G 2 γ d C d z ) κ 0 A 2 + ν γ 2 σ 2 A 4 + 4 β γ C = 0 .
G 2 γ ( R 2 d A d z + 2 A R d R d z ) + A [ R 2 ( 4 C + 2 δ 0 3 ε σ A 2 + 2 μ γ σ 2 A 4 ) + β γ ( 1 + 12 C 2 R 4 ) ] = 0 ,
d A d z = γ A G 2 [ 5 ε σ A 2 4 ( C + μ γ σ 2 A 4 + β γ R 2 C 2 ) 3 β γ R 2 2 δ 0 ] ,
d R d z = γ R G 2 ( 4 C ε σ A 2 + μ γ σ 2 A 4 + β γ R 2 4 β γ C 2 R 2 ) ,
d C d z = γ G 2 ( 1 R 4 4 C 2 A 2 R 2 + ν γ σ 2 A 4 R 2 4 β γ C R 2 ) ,
d ψ d z = γ G 2 ( κ 0 2 R 2 + 3 A 2 5 2 ν γ σ 2 A 4 + 4 β γ C ) .
A ± = ( 2 ε / σ β / γ ) ± ( 2 ε / σ β / γ ) 2 2 δ 0 ( 3 μ γ / σ 2 2 β ν / σ 2 ) ( 3 μ γ / σ 2 2 β ν / σ 2 ) .
λ 3 + α 1 λ 2 + α 2 λ + α 3 = 0
α 1 = γ G 2 [ 6 β γ R 2 + 16 μ γ σ 2 A 4 10 ε σ A 2 + 8 C ( 1 + β γ CR 2 ) ] > 0 ,
α 2 = γ 2 G 4 8 R 4 [ 1 + β 2 γ 2 A 2 R 2 ( 1 + 6 β ε γ σ 9 β μ σ 2 A 2 2 ν γ σ 2 A 2 ) 2 CA 2 R 4 ( β γ + 5 ε σ 8 μ γ σ 2 A 2 2 β ν σ 2 A 2 ) + 4 C 2 R 4 ( 1 + β 2 γ 2 3 β ε γ σ A 2 R 2 + 5 β μ σ 2 A 4 R 2 ) ] > 0 ,
α 3 = γ 3 G 6 32 A 2 R 4 [ β γ 2 ε σ + β 2 γ 2 ε σ + A 2 ( 3 μ γ σ 2 + β 2 μ γ σ 2 2 β ν σ 2 ) 4 CR 2 β γ ( ε σ β γ 2 μ γ σ 2 A 2 + 2 β ν σ 2 A 2 ) 4 C 2 R 4 ( 2 ε σ + β γ + 3 β 2 γ 2 ε σ 3 μ γ σ 2 A 2 5 β 2 μ γ σ 2 A 2 2 β ν σ 2 A 2 ) ] > 0 ,
α 123 = α 1 α 2 α 3 > 0 .
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.