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

Solitons and vortices in nonlinear two-dimensional photonic crystals of the Kronig-Penney type

Open Access Open Access

Abstract

Solitons in the model of nonlinear photonic crystals with the transverse structure based on two-dimensional (2D) quadratic- or rhombic-shaped Kronig-Penney (KP) lattices are studied by means of numerical methods. The model can also applies to a Bose-Einstein condensate (BEC) trapped in a superposition of linear and nonlinear 2D periodic potentials. The analysis is chiefly presented for the self-repulsive nonlinearity, which gives rise to several species of stable fundamental gap solitons, dipoles, four-peak complexes, and vortices in two finite bandgaps of the underlying spectrum. Stable solitons with complex shapes are found, in particular, in the second bandgap of the KP lattice with the rhombic structure. The stability of the localized modes is analyzed in terms of eigenvalues of small perturbations, and tested in direct simulations. Depending on the value of the KP’s duty cycle (DC, i.e., the ratio of the void’s width to the lattice period), an internal stability boundary for the solitons and vortices may exist inside of the first bandgap. Otherwise, the families of the localized modes are entirely stable or unstable in the bandgaps. With the self-attractive nonlinearity, only unstable solitons and vortices are found in the semi-infinite gap.

©2011 Optical Society of America

1. Introduction and the model

In this work, we consider the basic model of periodically patterned photonic media with the two-dimensional (2D) transverse structure, which are uniform in the longitudinal direction, z. These media represent quasi-2D photonic crystals [1, 2], which are tantamount, as a matter of fact, to photonic-crystal fibers (PCFs) [311]. The transverse structures that we deal with can be fabricated in the form of a host medium perforated by a system of parallel square-shaped voids, which are filled with a different material, featuring a nearly matched refractive index, n 0, simultaneously with a strong contrast in the Kerr coefficient, n 2. Two different types of such 2D structures, that we will call square-shaped and rhombic ones, respectively, are displayed in Figs. 1 and 2. Both types may be considered as 2D versions of the Kronig-Penney (KP) lattice, diverse forms of which were investigated in photonics in the 1D geometry, in terms of the nonlinear transmission [1417] and spatial solitons alike [1824]. 1D linear and nonlinear potentials of the KP type where also investigated in models of Bose-Einstein condensates (BECs) [25, 26]. The 2D KP structures may be realized in all-solid PCFs, cf. Refs. [12, 27], or as PCFs with the voids infiltrated by liquids with specially chosen optical properties [13, 2831]. The latter setting admits an additional possibility to control optical properties of the medium by means of electric fields [28, 29].

 figure: Fig. 1

Fig. 1 (Color online) (a) The transverse structure of the photonic crystal of the Kronig-Penney type, with the square-shaped modulation pattern corresponding to Eq. (2). Panels (b), (c) and (d) display the linear dispersion relation for this photonic crystal with depth U = 10 in Eq. (2), plotted along the edge Γ → XM → Γ of the irreducible Brillouin zone, for three values of the duty-cycle parameter: DC = 0.75, DC = 0.5 and DC = 0.25, respectively. Photonic bandgaps are shaded (the lowest one is the semi-infinite gap), while white areas represent the first two Bloch bands.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 (Color online) (a) The shape of the rhombic 2D Kronig-Penney structure with duty cycle DC = 0.5. (b) The linear dispersion relation for this structure, formed by the superposition of two modulation functions (2) with U = 10, is plotted along the edge Γ → XM → Γ of the irreducible Brillouin zone. Shaded and white areas correspond to the gaps (the semi-infinite and two finite ones) and the first two Bloch bands, respectively.

Download Full Size | PDF

Such composite optical media have drawn considerable attention as realizations of nonlinear lattices, i.e., media with periodically modulated effective nonlinear potentials, which support specific species of solitons, see recent review [32]. Other possibilities for the creation of non-linear lattices may be based on the use of doping techniques in solid-state waveguides [33], and structures induced in liquid crystals be means of a properly patterned external electric field [34].

On the other hand, in the more traditional situation, with only n 0 subject to the spatially periodic modulation, while n 2 is constant, the PCF is replaced by the nonlinearly uniform medium equipped with a linear lattice potential [35]; in the limit of a very deep lattice potential, this medium becomes effectively discrete [3639]. In particular, the 2D linear potential lattice of the KP type gives rise to stable nonlinear Bloch waves [40], as well as solitons and solitary vortices [41], including the case when the uniform nonlinearity includes a “supercritical” self-focusing quintic term [42]. It is also relevant to mention that 2D variants of the KP potential for the electron gas were experimentally created in solid-state physics, by means of superlattice heterostructures [43].

The model of the nonlinear photonic crystal outlined above, which includes the periodic modulation of both n 0 and n 2 in the transverse plane, is based on the 2D nonlinear Schrödinger (NLS) equation for the spatial evolution of the complex amplitude of the electromagnetic field, Ψ(x,z):

iΨz+Ψxx+Ψyy+W(x,y)(1+σ|Ψ|2)Ψ=0.
Here W (x,y) is the KP modulation function, which, in the case of the square pattern shown in Fig. 1, describes the periodic lattice of the nonlinear guiding channels, each of width D and depth U > 0, which are separated by linear buffer stripes of width LD, i.e., D/L ≡ DC may be defined as the “duty cycle” of the guiding set:
W(x,y)={U,Ln<x<D+LnandLn<y<D+Ln,0,D+Ln<x<L(1+n)and/orD+Ln<y<L(1+n),
n = 0, ±1, ±2... . Further, signs σ = +1 and −1 in Eq. (1) correspond, respectively, to the self-focusing and self-defocusing nonlinearity. In this study, we fix the normalizations in Eq. (2) by setting L = 2π, and consider a value of the modulation depth, U = 10, which makes it possible to present generic results, while the DC of the modulation pattern [Eq. (2)] will be given two characteristic values, 0.75 and 0.5.

The juxtaposition of two square patterns [Eq. (2)] with DC ≤ 0.5, shifted relative to each other in the diagonal direction by L/ 2, gives rise to the rhombic structure. In particular, it takes the form displayed in Fig. 2 in the case of DC = 0.5. Only this case will be considered below as a representative of the 2D KP lattices of the rhombic type.

A straightforward estimate demonstrates that, with periodicity L = 10λ of the microstructure guiding the light with wavelength λ, the depth U = 10 corresponds to the refractive-index contrast δn 0 ≃ 0.02 between the channels and the host medium, in the all-solid setting or the one infiltrated with a liquid. The intensity of the input light beams, which create various solitons reported below, must be high enough, to generate the characteristic nonlinearity length on the order of few millimeters, assuming λ ∼ 1 μm.

The “checkerboard”-like pattern adopted in Eqs. (1) and (2) may also be created in a virtual form, rather than as the solid structure in the material medium, by means of electromagnetically-induced transparency, as recently proposed in Ref. [44]. Further, the same model applies, in a completely different physical setting, to the description of effectively 2D patterns in BECs (Bose-Einstein condensates). The linear potential of the KP type, which appears in Eqs. (1) and (2), can be created by dint of magnetic lattices [45] (the well-known technique based on optical lattices [46,47] usually gives rise to effective sinusoidal potentials in the BEC), while the nonlinearity pattern may be induced by a system of focused laser beams, which modify the local scattering length via the optically-controlled Feshbach resonance [48, 49]. In the latter case, Eq. (1), with evolutional variable z replaced by time t, plays the role of the corresponding Gross-Pitaevskii equation.

The objective of this work is to find fundamental and vortical 2D solitons in Eq. (1), supported by the combination of the linear and nonlinear potentials determined by the KP modulation function [Eq. (2)], for both signs of the nonlinearity, self-focusing and defocusing. In the former case, solitons reside in the semi-infinite gap of the respective linear spectrum (where they turn out to be unstable, for the parameter values considered below), while in the latter case gap solitons will be found in two finite bandgaps. In previous works, solitons in 2D PCF media were found in other settings. Gap solitons and soliton trains in periodic and quasi-periodic 2D arrays of cylindrical voids were reported in Ref. [5]. The symmetry classification of fundamental and vortical solitons in hexagonal and more general periodic transverse structures were developed in detail for fundamental solitons in Refs. [6, 7], and for vortices—in Ref. [8].

The rest of the paper is structured as follows. In Section II methods of the numerical analysis used in this work are outlined (we rely on numerical methods, as the present 2D model is too difficult for analytical considerations). In Section III, the results are collected for gap solitons, their bound states (dipoles and four-peak complexes), and vortices of two types (off- and on-site-centered ones), all found in two finite bandgaps of the lattice’s linear spectrum, in the case of the quadratic transverse structure (see Fig. 1) and self-defocusing nonlinearity. The stability of the various solitary modes is investigated through the calculation of eigenvalues for small perturbations, and is verified by dint of direct simulations. In Section IV, the results are reported for the localized modes in the model based on the rhombic KP lattice, see Fig. 2. In that case, fundamental solitons, including those with complex shapes (found in the second finite bandgap) may be stable, but vortices are not. In Section IV, we also briefly summarize results obtained for fundamental solitons and vortices in the semi-infinite gap of the model with the self-attractive nonlinearity, all of which turn out to be unstable. The paper is concluded by Section V.

2. The framework of the analysis

2.1. Stationary solutions and the bandgap spectrum

Stationary soliton solutions to Eq. (1) are sought for as Ψ = eiβzΦ(x,y), where β is the propagation constant, and real wave function Φ(x,y) obeys the equation

βΦ+Φxx+Φyy+W(x,y)(Φ+σΦ3)=0.
Localized solutions to Eq. (3) were constructed by means if the modified Newton’s method developed in Ref. [50] (see also book [2]). Families of soliton solutions are characterized by the total power (or norm, in terms of the BEC model),
P(β)++|Ψ(x,y)|2dxdy,
which is a dynamical invariant of Eq. (1), together with the Hamiltonian of this equation.

Before proceeding to the analysis of solitons and vortices in the photonic-crystal model, we aim to construct the linear bandgap spectrum of the 2D KP modulation pattern [Eq. (2)]. Pursuant to the Floquet-Bloch theory, solutions to the linearized version of Eq. (3) can be sought for as Φ(x,y) = e ikxx+ikyy ϕ (x,y,kx, ky), where β = β (kx, ky) is dispersion relation, wavenumbers kx, ky belong to the first Brillouin zone, while ϕ (x,y,kx, ky) is a periodic function of x and y with the periodicity of the underlying KP lattice [Eq. (2)]. The dispersion relation for the 2D Bloch wave functions is found by solving the ensuing linear eigenvalue problem:

[(x+ikx)2+(y+iky)2+W(x,y)]φ(x,y)=βφ(x,y)
For the parameters fixed here, L = 2π and U = 10, Figs. 1(b), 1(c), and 1(d) and 2(b) display the numerically generated dispersion relation and the corresponding bandgaps of Eq. (5) along the edge of the irreducible Brillouin zone (points Γ → XM → Γ [1]) for, respectively, the square lattice [Eq. (2)] with DC = 0.75, 0.5 and 0.25, and for the rhombic lattice with DC = 0.5. Note that finite bandgaps are absent in Fig. 1 when DC and/or U are too small—e.g., in the case shown in Fig. 1(d) for DC = 0.25. The bandgaps also disappear when the depth of the KP lattice falls to U ≈ 2.5 for DC = 0.75.

2.2. The linear-stability analysis of solitons

To develop the stability analysis for 2D solitons, the solution is perturbed as follows:

Ψ(x,y,z)=eiβz{Φ0(x,y)+[v(x,y)w(x,y)]eλz+[v*(x,y)w*(x,y)]eλ*z}
where v(x,y) and w(x,y) represent small perturbations, and λ is the respective eigenvalue. Inserting this perturbed solution into Eq. (1) and linearizing, we arrive at the following linear-stability eigenvalue problem:
Lψ=λψ,
with the matrix operator
L=i(G02+G12+G2G0),ψ=(vw)G0=12(Φ02Φ0*2)σW(x,y)G1=β+W(x,y)+2σW(x,y)|Φ0|212σW(x,y)(Φ02+Φ0*2)G2=β+W(x,y)+2σW(x,y)|Φ0|2+12σW(x,y)(Φ02+Φ0*2)
The soliton is unstable if there are eigenvalues with nonzero real parts. The entire spectrum of the linear-stability operator L was constructed by means of the Fourier collocation method [51].

3. Photonic crystals with the square transverse structure

3.1. Fundamental solitons and their bound states

In the model with the square-lattice modulation function defined as per Eq. (2) and self-repulsive (defocusing) nonlinearity, which corresponds to σ = −1 in Eq. (1), 2D fundamental gap solitons have been found in the two finite bandgaps. Also found were diagonally oriented dipoles, i.e., bound states of two solitons with opposite signs, and in-phase four-soliton complexes of the square and rhombus types (alias off-site- and on-site-centered ones, respectively, i.e., rings formed by four peaks, without or with an empty site at the center—see, e.g., Ref. [52]). Here, we consider the complexes that carry no phase structure (i.e., they are not quadrupoles). For U = 10 and DC = 0.75, these families of the gap solitons are characterized by dependences of the integral power on the propagation constant, P(β) [see Eq. (4)], which are shown in Fig. 3(a) for the two finite bandgaps. Actually, the second bandgap is very narrow in this case,

8.315<β<8.398,
cf. Fig 1(b). Other panels in Fig. 3 display typical examples of stable and unstable fundamental and compound solitons found in the first bandgap. Examples of stable solitons of the same types, but found in the second bandgap, are displayed in Fig. 4. As concerns other types of gap-soliton complexes, such as straight dipoles (ones oriented along bonds of the KP lattice, rather than diagonally) and quadrupoles, they were found too, but these species of the localized patterns turn out to be completely unstable at DC = 0.75.

 figure: Fig. 3

Fig. 3 (Color online) (a) The power-vs.-propagation constant diagram for fundamental solitons (the red curve), diagonal dipoles (the blue curve), and bound complexes of four fundamental solitons (the magenta and black curves pertain to the square-shaped and rhombic complexes, respectively), in the two finite bandgaps of the 2D square-shaped Kronig-Penney structure with duty cycle DC = 0.75 and the self-defocusing nonlinearity. In this and other figures, solid and dashed segments of the curves represent stable an unstable modes, respectively. (b) Typical examples of stable and unstable fundamental solitons for β = 4.4, P = 70.71 and β = 4.0, P = 13.77, respectively. Here and in similar figures below, the left and right panels display, respectively, the distribution of the local amplitude in the stationary solutions, and the respective spectral plane of the stability eigenvalues. (c) Typical examples of stable and unstable square-shaped complexes for β = 4.4, P = 42.31 and β = 4.0, P = 50.79, respectively. (d) A typical example of a stable rhombic complex for β = 4.4, P = 41.98 (an example of unstable rhombuses is not displayed separately, as it does not show anything essentially different from the other modes). (e) Typical examples of stable and unstable diagonal dipoles for β = 4.4 and 4.1, respectively.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (Color online) Examples of stable fundamental solitons, diagonal dipoles, squares, and rhombuses, found in the narrow second bandgap of Fig. 3 at β = 8.35, with the total powers, respectively, P = 1.5013, P = 3.0022, P = 6.0184, and P = 5.9956.

Download Full Size | PDF

The stability of the solitons, for which the corresponding analysis reveals no eigenvalues with nonzero real parts, has been corroborated by direct simulations (not shown here). The simulations of the evolution of the solitons with small real parts, which typically account for the instability in the present case (see Fig. 3), performed by means of the standard split-step Fourier-transform method, with absorbers installed at edges of the integration domain, demonstrate a fairly weak instability, in the form of a very slow decay of perturbed solitons, as shown in Fig. 5.

 figure: Fig. 5

Fig. 5 (Color online) Typical examples of the evolution of weakly unstable solitons is displayed by means of the cross section of |Ψ(x,y)|2 drawn through point y = 0: (a) and (b) —a fundamental gap soliton and square-shaped four-soliton complex, both with β = 4.0; (c) —a diagonal dipole with β = 4.1.

Download Full Size | PDF

The situation for the gap solitons and their bound complexes is different at DC = 0.5. In this case, the second finite bandgap does not exist, while the families of the fundamental solitons and all their bound states, viz., straight and diagonal dipoles, square- and rhombus-shaped complexes, as well as square-shaped quadrupoles, are all entirely stable in the first bandgap (recall that the straight dipoles and quadrupoles were completely unstable at DC = 0.75). The corresponding P(β) dependences, as well as generic examples of the stable solitons and their bound states, are displayed in Fig. 6.

 figure: Fig. 6

Fig. 6 (Color online) (a) The red, blue, green, black, magenta and yellow curves show the P(β) dependences for fundamental solitons, straight dipoles, diagonal dipoles, rhombus complexes, square complexes and quadrupoles, respectively, in the single finite bandgap existing at DC = 0.5 in the model with the square-shaped transverse structure and self-defocusing nonlinearity. (b) Generic examples of fundamental solitons and quadrupoles (top and bottom, respectively). (c) Examples of squares and rhombuses. (d) Examples of straight and diagonal dipoles. All the solutions are displayed for β = 2.4. All the solitons and their bound states are stable, in this case.

Download Full Size | PDF

3.2. Vortices

Similarly to the four-soliton complexes without the phase structure, that were considered above, solitary vortices with topological charge 1 can be built, in the first and second finite bandgaps, as squares and rhombuses, alias off-site- and on-site-centered four-peak rings, which carry the phase circulation of 2π (see Refs. [53] and [35] for a review of vortex solitons in linear lattices). For DC = 0.75, the P(β) curves for the vortices of both types and their typical examples are shown in Fig. 7(a) (rhombic vortices were not found at β < 4.4). Simulations of perturbed unstable vortices (not shown here in detail) demonstrate the evolution which is quite similar to that shown for the weakly unstable gap solitons and complexes in Fig. 5, i.e., a slow decay (with the loss of the vorticity).

 figure: Fig. 7

Fig. 7 (Color online) (a) The dotted and solid curves show the integral power, P, as a function of the propagation constant, β, for the square-shaped and rhombic vortices (with topological charge 1), respectively, at DC = 0.75, in the two finite bandgaps of the photonic crystal with the square transverse structure and self-defocusing nonlinearity. (b) A typical example of the amplitude and phase distributions, and the set of perturbation eigenvalues, for the stable square-shaped vortex with β = 4.4 and P = 41.845. (c) The same for a typical unstable one with β = 4.05 and P = 48.949. (d) A typical example of the stable rhombic vortex with β = 4.4 and P = 42.562.

Download Full Size | PDF

In the narrow second bandgap [see Eq. (9)] the vortices of both types are stable. Examples of those are displayed in Fig. 8.

 figure: Fig. 8

Fig. 8 (Color online) (Upper row) and (bottom row): Examples of stable square-shaped and rhombic vortices (the amplitude and phase distributions) found in the narrow second bandgap (9) for DC = 0.75, with β = 8.35 and P = 5.9879 or P = 6.0108, respectively.

Download Full Size | PDF

Finally, for DC = 0.5 (recall that there is the single finite bandgap in this case) solutions for square-shaped vortices were not found, while rhombic vortices exist and are stable in the entire bandgap in this case, as shown in Fig. 9. The fact that vortical rhombuses are more robust, in the sense of the existence and stability, than their square-shaped counterparts, is a general feature which was also found in other models [52].

 figure: Fig. 9

Fig. 9 (Color online) (a) The P(β) dependence for rhombus-shaped vortices in the finite bandgap at DC = 0.5 in the model with the square-shaped transverse structure and self-defocusing nonlinearity. (b) An example of the amplitude and phase distribution in the stable vortex for β = 2.4 and P = 26.1702.

Download Full Size | PDF

4. Photonic crystals with the rhombic transverse structure

4.1. Simple and complex gap solitons

In the model based on the rhombic structure with DC = 0.5, which is shown in Fig. 2, only fundamental gap solitons were found in the first finite bandgap, but not their bound states in the form of dipoles, squares, or rhombuses. The P(β) curve for the solitons in two finite bandgaps of this photonic crystal are shown in Fig. 10(a). The (in)stability of the gap solitons is illustrated in Fig. 10(b) by plotting the largest instability growth rate versus propagation constant β, the zero value of the latter implying the stability. In fact, stability intervals in the first bandgap are quite narrow: there are two of them on the top branch of the respective curve P(β),

4.15<β<4.00,24.14<P<25.69,
3.70<β<3.50,30.52<P<75.20,
and a single one on the bottom one,
4.30<β<4.27,10.35<P<10.63.
Generic examples of stable and unstable gap solitons found in the first finite bandgap are displayed in Figs. 10(c) and (d), respectively.

 figure: Fig. 10

Fig. 10 (Color online) The power-vs.-propagation constant curves for the fundamental solitons in two finite bandgaps of the photonic crystal with the rhombic transverse structure (see Fig. 2) and self-defocusing nonlinearity. In the second bandgap, the top branch is unstable, while the two lower ones are stable. (b) The largest instability growth rate of the gap solitons as functions of the propagation constant, β. The solid and dotted curves in the first bandgap in (b) correspond, respectively, to the bottom and top portions of the P(β) curve in (a). The curve in the second bandgap in (b) corresponds to the top P(β) branch in the same bandgap in (a). A typical example of stable solitons in the first bandgap is shown in panel (c) for β = 3.52 and P = 60.00. (d) Examples of unstable solitons with β = 3.9 in the first bandgap for P = 8.5687 and P = 26.6719 (top and bottom rows, respectively).

Download Full Size | PDF

The comparison of Fig. 10 to Fig. 3 demonstrates that the instability of the gap solitons (of those of them which are unstable) is much stronger in the present case than in the model of the photonic crystal with the square transverse structure. Accordingly, direct simulations, displayed in Figs. 11(a) and 11(b), demonstrate a much faster decay of the unstable solitons, cf. Fig. 5.

 figure: Fig. 11

Fig. 11 (Color online) (a) and (b): The evolution of unstable solitons corresponding to the top and bottom rows in Fig. 10 is shown in the cross section of |Ψ(x,y)|2 drawn through y = 0. (c): The same for the unstable pattern from Fig. 12(b).

Download Full Size | PDF

The second bandgap hosts complex localized modes, both stable and unstable ones, which seem as bound states of fundamental solitons, see generic examples in Fig. 12. However, the simplest among these species, at least—the rhombus which is shown in the top row of Fig. 12(a)—is actually a fundamental gap soliton by itself, in the sense that there are no single-peak solitons that would build this four-peak pattern. As for the stable eight-peak ring shown in the bottom row of Fig. 12(a), and the complex unstable pattern displayed in Fig. 12(b), the relation between the shapes and powers of the three modes presented in Fig. 12, which pertain to the common value of the propagation constant, suggests that, in principle, the eight-peak ring may be realized as a superposition of two four-peak fundamental rhombuses, and the unstable pattern shown in Fig. 12(b) may be built as a complex of five eight-peak rings [the evolution of the latter unstable pattern is displayed in Fig. 11(c)]. We stress that all these three species of the complex modes carry no phase structure.

 figure: Fig. 12

Fig. 12 (Color online) Typical examples of solitons, found in the second finite bandgap of the photonic crystal with the rhombic structure and self-defocusing nonlinearity, for β = 2.1. In (a), the top and bottom rows display examples of stable solitons, with P = 5.9460 and P = 13.4476, which correspond, respectively, to the lower and middle P(β) curves in the second bandgap in Fig. 10. (b) A typical example of an unstable soliton with P = 53.0894, which corresponds to the upper curve in the second bandgap in Fig. 10.

Download Full Size | PDF

4.2. Vortices

Localized vortex patterns, with rather complex shapes, were also found in the two finite bandgaps of this model, but they all turn out to be unstable. Typical example of the shapes and instability spectra are displayed in Fig. 13. Direct simulations of the evolution of the unstable vortices demonstrate a quick transition to chaotic dynamics (not shown here).

 figure: Fig. 13

Fig. 13 (Color online) Examples of unstable vortices found, respectively, in the first and second finite bandgaps of the model with the rhombic structure, with β = −4.25, P = 30.4432 (a), and β = 2.1, P = 64.2146 (b).

Download Full Size | PDF

4.3. The self-focusing nonlinearity

In the same model with the self-attractive nonlinearity, which corresponds to σ = +1 in Eq. (1), ordinary fundamental solitons supported by the lattice structure [53] can be found in the semi-infinite gap, but they are unstable. The P(β) dependence for this family, along with a typical example of the shape and instability of the fundamental soliton, are displayed in Fig. 14. Rhombic vortices have also been found in the semi-infinite gap, but they are completely unstable too, suffering a fast decay (not shown here). Square-shaped vortices could not be constructed in the model with the self-focusing nonlinearity.

 figure: Fig. 14

Fig. 14 (Color online) (a) The power-vs.-propagation-constant dependence for solitons supported by the self-focusing nonlinearity in the model with the rhombic transverse structure. (b) A typical example of an unstable soliton, with β = 8.0 and P = 48.0398. (c) Decay of this soliton in direct simulations.

Download Full Size | PDF

5. Conclusion

This work reports results of the systematic analysis of 2D solitons, their bound states of various types, and solitary vortices in two fundamental models of nonlinear optical crystals and/or optical-crystal fibers, with the KP (Kronig-Penney) square or rhombic lattice of voids filled by a linear material. The same model can be realized in terms of the BEC (Bose-Einstein condensate) loaded into a combination of linear and nonlinear 2D lattice potentials. The solitons and solitary vortices can be created experimentally in the physical settings of both types.

The analysis was mainly focused on the models with the self-repulsive nonlinearity, which gives rise to many species of stable solitons and vortices in the two lowest finite bandgaps of the respective linear spectrum. In particular, stable fundamental gap solitons feature intricate shapes in the second finite bandgap of the photonic crystal with the rhombic transverse structure. The stability of the various modes was investigated through the calculation of the corresponding eigenvalues, and verified via direct simulations of the perturbed evolution. Depending on the value of the lattice’s DC (duty cycle, i.e., the ratio of the void’s width to the pitch of the KP lattice), an intrinsic stability border for the gap solitons, complexes built of them, and solitary vortices can be found inside the first finite bandgap. In other cases, the families of solitons and vortices are completely stable or unstable inside the respective finite bandgaps. The case of the self-attraction was briefly considered too, but only unstable solitary modes were found in that case in the semi-infinite gap.

The work may be extended in other directions, to consider additional physically relevant issues. In particular, a challenging problem is mobility of the solitons in the present setting (it is known that 2D gap solitons may be mobile in linear lattices [54, 55]). Another natural extension may aim to cover a broader range of parameters of the square-shaped and rhombic KP lattices, such as DC and the depth of the lattice, U, and to generalize the model for an anisotropic version of the lattice, cf. Refs. [55] and [52], where experimental and theoretical results were reported for solitons in anisotropic 2D linear-lattice potentials.

Acknowledgment

The work of T.M. was supported by the Thailand Research Fund through grant RMU5380005.

References and links

1. J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, Photonic Crystals: Molding the Flow of Light (Princeton University Press, 2008).

2. M. Skorobogatiy and J. Yang, Fundamentals of Photonic Crystals Guiding (Cambridge University Press, 2009).

3. J. C. Knight, J. Broeng, T. A. Birks, and P. St. J. Russel, “Photonic band cap guidance in optical fibers,” Science 282, 1476–1478 (1998). [CrossRef]   [PubMed]  

4. B. J. Eggleton, B. J., C. Kerbage, P. S. Westbrook, R. S. Windeler, and A. Hale, “Microstructured optical fiber devices,” Opt. Express 9, 698–713 (2001). [CrossRef]   [PubMed]  

5. P. Xie, Z.-Q. Zhang, and X. Zhang, “Gap solitons and soliton trains in finite-sized two-dimensional periodic and quasiperiodic photonic crystals,” Phys. Rev. E 67, 026607 (2003). [CrossRef]  

6. A. Ferrando, M. Zacarés, P. F. de Córdoba, D. Binosi, and J. A. Monsoriu, “Spatial soliton formation in photonic crystal fibers,” Opt. Express 11, 452–459 (2003). [CrossRef]   [PubMed]  

7. Y. V. Kartashov, A. Ferrando, A. A. Egorov, and L. Torner, “Soliton topology versus discrete symmetry in optical lattices,” Phys. Rev. Lett. 95, 123902 (2005). [CrossRef]   [PubMed]  

8. A. Ferrando, M. Zacarés, P. F. de Córdoba, D. Binosi, and J. A. Monsoriu, “Vortex solitons in photonic crystal fibers,” Opt. Express 12, 817–822 (2004). [CrossRef]   [PubMed]  

9. T. M. Monro and D. J. Richardson, “Holey optical fibres: Fundamental properties and device applications,” C. R. Physique 4, 175–186 (2003). [CrossRef]  

10. P. St. J. Russell, “Photonic-crystal fibers,” J. Lightwave Technol. 24, 4729–4749 (2006). [CrossRef]  

11. S. Arismar Cerqueira Jr., “Recent progress and novel applications of photonic crystal fibers,” Rep. Prog. Phys. 73, 024401 (2010). [CrossRef]  

12. F. Luan, A. K. George, T. D. Hedley, G. J. Pearce, D. M. Bird, J. C. Knight, and P. St. J. Russell, “All-solid photonic bandgap fiber,” Opt. Lett. 29, 2369–2371 (2004). [CrossRef]   [PubMed]  

13. T. T. Larsen, A. Bjarklev, D. S. Hermann, and J. Broeng, “Optical devices based on liquid crystal photonic bandgap fibres,” Opt. Express 11, 2589–2596 (2003). [CrossRef]   [PubMed]  

14. J. P. Dowling and C. M. Bowden, “Anomalous index of refraction in photonic bandgap materials,” J. Mod. Opt. 41, 345–351 (1994). [CrossRef]  

15. Q. Li, C. T. Chan, K. M. Ho, and C. M. Soukoulis, “Wave propagation in nonlinear photonic band-gap materials,” Phys. Rev. B 53, 15577–15585 (1996). [CrossRef]  

16. E. Lidorikis, Q. Li, and C. M. Soukoulis, “Wave propagation in nonlinear multilayer structures,” Phys. Rev. B 54, 10249–10252 (1996). [CrossRef]  

17. D. Hennig and G. P. Tsironis, “Wave transmission in nonlinear lattices,” Phys. Rep. 307, 333–432 (1999). [CrossRef]  

18. A. A. Sukhorukov and Y. S. Kivshar, “Nonlinear localized waves in a periodic medium,” Phys. Rev. Lett. 87, 083901 (2001). [CrossRef]   [PubMed]  

19. W. Li and A. Smerzi, “Nonlinear Krönig-Penney model,” Phys. Rev. E 70, 016605 (2004). [CrossRef]  

20. I. M. Merhasin, B. V. Gisin, R. Driben, and B. A. Malomed, “Finite-band solitons in the Kronig-Penney model with the cubic-quintic nonlinearity,” Phys. Rev. E 71, 016613 (2005). [CrossRef]  

21. Y. Kominis, “Analytical solitary wave solutions of the nonlinear Kronig-Penney model in photonic structures,” Phys. Rev. E 73, 066619 (2006). [CrossRef]  

22. Y. Kominis and K. Hizanidis, “Lattice solitons in self-defocusing optical media: analytical solutions of the nonlinear Kronig-Penney model,” Opt. Lett. 31, 2888–2890 (2006). [CrossRef]   [PubMed]  

23. Y. Kominis, A. Papadopoulos, and K. Hizanidis, “Surface solitons in waveguide arrays: Analytical solutions,” Opt. Express 15, 10041–10051 (2007). [CrossRef]   [PubMed]  

24. T. Mayteevarunyoo and B. A. Malomed, “Solitons in one-dimensional photonic crystals,” J. Opt. Soc. Am. B 25, 1854–1863 (2008). [CrossRef]  

25. B. T. Seaman, L. D. Carr, and M. J. Holland, “Nonlinear band structure in Bose-Einstein condensates: nonlinear Schrödinger equation with a Kronig-Penney potential,” Phys. Rev. A 71, 033622 (2005). [CrossRef]  

26. A. S. Rodrigues, P. G. Kevrekidis, M. A. Porter, D. J. Frantzeskakis, P. Schmelcher, and A. R. Bishop, “Matter-wave solitons with a periodic, piecewise-constant scattering length,” Phys. Rev. A 78, 013611 (2008). [CrossRef]  

27. G. Bouwmans, L. Bigot, Y. Quiquempois, F. Lopez, L. Provino, and M. Douay, “Fabrication and characterization of an all-solid 2D photonic bandgap fiber with a low-loss region ¡ 20 dB/km) around 1550 nm,” Opt. Express 13, 8452–8459 (2005). [CrossRef]   [PubMed]  

28. F. Du, Y. Q. Lu, and S. T. Wu, “Electrically tunable liquid-crystal photonic crystal fiber,” Appl. Phys. Lett. 85, 2181–2183 (2004). [CrossRef]  

29. M. W. Haakestad, T. T. Alkeskjold, M. D. Nielsen, L. Scolari, J. Riishede, H. E. Engan, and A. Bjarklev, “Electrically tunable photonic bandgap guidance in a liquid-crystal-filled photonic crystal fiber,” IEEE Photon. Technol. Lett. 17, 819–821 (2005). [CrossRef]  

30. A. Fuerbach, P. Steinvurzel, J. A. Bolger, A. Nulsen, and B. J. Eggleton, “Nonlinear propagation effects in antiresonant high-index inclusion photonic crystal fibers,” Opt. Lett. 30, 830 (2005). [CrossRef]   [PubMed]  

31. C. R. Rosberg, F. H. Bennet, D. N. Neshev, P. D. Rasmussen, O. Bang, W. Królikowski, A. Bjarklev, and Y. S. Kivshar, “Tunable diffraction and self-defocusing in liquid-filled photonic crystal fibers,” Opt. Express 15, 12145 (2007). [CrossRef]   [PubMed]  

32. Y. V. Kartashov, B. A. Malomed, and L. Torner, “Solitons in nonlinear lattices,” Rev. Mod. Phys. 83, 247–306 (2011). [CrossRef]  

33. J. Hukriede, D. Runde, and D. Kip, “Fabrication and application of holographic Bragg gratings in lithium niobate channel waveguides,” J. Phys. D 36, R1 (2003). [CrossRef]  

34. A. Fratalocchi, G. Assanto, K. A. Brzdakiewicz, and M. A. Karpierz, “Discrete propagation and spatial solitons in nematic liquid crystals,” Opt. Lett. 29, 1530–1532 (2004). [CrossRef]   [PubMed]  

35. Y. V. Kartashov, V. A. Vysloukh, and L. Torner, “Soliton shape and mobility control in optical lattices,” Prog. Opt. 52, 63–148 (2009). [CrossRef]  

36. D. N. Christodoulides and R. I. Joseph, “Discrete self-focusing in nonlinear arrays of coupled waveguides,” Opt. Lett. 13, 794–796 (1988). [CrossRef]   [PubMed]  

37. J. W. Fleischer, M. Segev, N. K. Efremidis, and D. N. Christodoulides, “Observation of two-dimensional discrete solitons in optically induced nonlinear photonic lattices,” Nature 422, 147–150 (2003). [CrossRef]   [PubMed]  

38. D. N. Christodoulides, F. Lederer, and Y. Silberberg, “Discretizing light behaviour in linear and nonlinear waveguide lattices,” Nature 424, 817–823 (2003). [CrossRef]   [PubMed]  

39. F. Lederer, G. I. Stegeman, D. N. Christodoulides, G. Assanto, M. Segev, and Y. Silberberg, “Discrete solitons in optics,” Phys. Rep. 463, 1–126 (2008). [CrossRef]  

40. B. Maes, P. Bienstman, and R. Baets, “Bloch modes and self-localized waveguides in nonlinear photonic crystals,” J. Opt. Soc. Am. B 22, 613–619 (2005). [CrossRef]  

41. R. Driben, B. A. Malomed, A. Gubeskys, and J. Zyss, “Cubic-quintic solitons in the checkerboard potential,” Phys. Rev. E 76, 066604 (2007). [CrossRef]  

42. R. Driben and B. A. Malomed, “Stabilization of two-dimensional solitons and vortices against supercritical collapse by lattice potentials,” Eur. Phys. J. D 50, 317–323 (2008). [CrossRef]  

43. H. L. Stormer, L. N. Pfeiffer, K. W. Baldwin, K. W. West, and J.Spector, atomically precise superlattice potential imposed on a 2-dimensional electron gas,” Appl. Phys. Lett. 58, 726–728 (1991). [CrossRef]  

44. Y. Li, B. A. Malomed, M. Feng, and J. Zhou, “Arrayed and checkerboard optical waveguides controlled by the electromagnetically induced transparency,” Phys. Rev. A 82, 633813 (2010). [CrossRef]  

45. S. Ghanbari, T. D. Kieu, A. Sidorov, and P. Hannaford, “Permanent magnetic lattices for ultracold atoms and quantum degenerate gases,” J. Phys. B 39, 847 (2006). [CrossRef]  

46. D. Jaksch, C. Bruder, J. I. Cirac, C. W. Gardiner, and P. Zoller, “Cold bosonic atoms in optical lattices,” Phys. Rev. Lett. 81, 3108 (1998). [CrossRef]  

47. M. Greiner, O. Mandel, T. Esslinger, T. W. Hansch, and I. Bloch, “Quantum phase transition from a superfluid to a Mott insulator in a gas of ultracold atoms,” Nature 415, 39 (2002). [CrossRef]   [PubMed]  

48. P. O. Fedichev, Y. Kagan, G. V. Shlyapnikov, and J. T. M. Walraven, “Influence of nearly resonant light on the scattering length in low-temperature atomic gases,” Phys. Rev. Lett. 77, 2913–2916 (1996). [CrossRef]   [PubMed]  

49. M. Theis, M., G. Thalhammer, K. Winkler, M. Hellwig, G. Ruff, R. Grimm, and J. H. Denschlag, “Tuning the scattering length with an optically induced Feshbach resonance,” Phys. Rev. Lett. 93, 123001 (2004). [CrossRef]   [PubMed]  

50. J. Yang, “Newton-conjugate gradient methods for solitary wave computations,” J. Comput. Phys. 228, 7007–7024 (2009). [CrossRef]  

51. J. Yang, Nonlinear Waves in Integrable and Nonintegrable Systems (SIAM, 2010). [CrossRef]  

52. T. Mayteevarunyoo, B. A. Malomed, B. B. Baizakov, and M. Salerno, “Matter-wave vortices and solitons in anisotropic optical lattices,” Physica D 238, 1439–1448 (2009). [CrossRef]  

53. B. A. Malomed, D. Mihalache, F. Wise, and L. Torner, “Spatiotemporal optical solitons,” J. Opt. B: Quant. Semiclass. Opt. 7, R53–R72 (2005). [CrossRef]  

54. H. Sakaguchi and B. A. Malomed, “Two-dimensional loosely and tightly bound solitons in optical lattices and inverted traps,” J. Phys. B 37, 2225–2239 (2004). [CrossRef]  

55. R. Fischer, D. Trager, D. N. Neshev, A. A. Sukhorukov, W. Królikowski, C. Denz, and Y. S. Kivshar, “Reduced-symmetry two-dimensional solitons in photonic lattices,” Phys. Rev. Lett.96, 023905 (2006). [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 (14)

Fig. 1
Fig. 1 (Color online) (a) The transverse structure of the photonic crystal of the Kronig-Penney type, with the square-shaped modulation pattern corresponding to Eq. (2). Panels (b), (c) and (d) display the linear dispersion relation for this photonic crystal with depth U = 10 in Eq. (2), plotted along the edge Γ → XM → Γ of the irreducible Brillouin zone, for three values of the duty-cycle parameter: DC = 0.75, DC = 0.5 and DC = 0.25, respectively. Photonic bandgaps are shaded (the lowest one is the semi-infinite gap), while white areas represent the first two Bloch bands.
Fig. 2
Fig. 2 (Color online) (a) The shape of the rhombic 2D Kronig-Penney structure with duty cycle DC = 0.5. (b) The linear dispersion relation for this structure, formed by the superposition of two modulation functions (2) with U = 10, is plotted along the edge Γ → XM → Γ of the irreducible Brillouin zone. Shaded and white areas correspond to the gaps (the semi-infinite and two finite ones) and the first two Bloch bands, respectively.
Fig. 3
Fig. 3 (Color online) (a) The power-vs.-propagation constant diagram for fundamental solitons (the red curve), diagonal dipoles (the blue curve), and bound complexes of four fundamental solitons (the magenta and black curves pertain to the square-shaped and rhombic complexes, respectively), in the two finite bandgaps of the 2D square-shaped Kronig-Penney structure with duty cycle DC = 0.75 and the self-defocusing nonlinearity. In this and other figures, solid and dashed segments of the curves represent stable an unstable modes, respectively. (b) Typical examples of stable and unstable fundamental solitons for β = 4.4, P = 70.71 and β = 4.0, P = 13.77, respectively. Here and in similar figures below, the left and right panels display, respectively, the distribution of the local amplitude in the stationary solutions, and the respective spectral plane of the stability eigenvalues. (c) Typical examples of stable and unstable square-shaped complexes for β = 4.4, P = 42.31 and β = 4.0, P = 50.79, respectively. (d) A typical example of a stable rhombic complex for β = 4.4, P = 41.98 (an example of unstable rhombuses is not displayed separately, as it does not show anything essentially different from the other modes). (e) Typical examples of stable and unstable diagonal dipoles for β = 4.4 and 4.1, respectively.
Fig. 4
Fig. 4 (Color online) Examples of stable fundamental solitons, diagonal dipoles, squares, and rhombuses, found in the narrow second bandgap of Fig. 3 at β = 8.35, with the total powers, respectively, P = 1.5013, P = 3.0022, P = 6.0184, and P = 5.9956.
Fig. 5
Fig. 5 (Color online) Typical examples of the evolution of weakly unstable solitons is displayed by means of the cross section of |Ψ(x,y)|2 drawn through point y = 0: (a) and (b) —a fundamental gap soliton and square-shaped four-soliton complex, both with β = 4.0; (c) —a diagonal dipole with β = 4.1.
Fig. 6
Fig. 6 (Color online) (a) The red, blue, green, black, magenta and yellow curves show the P(β) dependences for fundamental solitons, straight dipoles, diagonal dipoles, rhombus complexes, square complexes and quadrupoles, respectively, in the single finite bandgap existing at DC = 0.5 in the model with the square-shaped transverse structure and self-defocusing nonlinearity. (b) Generic examples of fundamental solitons and quadrupoles (top and bottom, respectively). (c) Examples of squares and rhombuses. (d) Examples of straight and diagonal dipoles. All the solutions are displayed for β = 2.4. All the solitons and their bound states are stable, in this case.
Fig. 7
Fig. 7 (Color online) (a) The dotted and solid curves show the integral power, P, as a function of the propagation constant, β, for the square-shaped and rhombic vortices (with topological charge 1), respectively, at DC = 0.75, in the two finite bandgaps of the photonic crystal with the square transverse structure and self-defocusing nonlinearity. (b) A typical example of the amplitude and phase distributions, and the set of perturbation eigenvalues, for the stable square-shaped vortex with β = 4.4 and P = 41.845. (c) The same for a typical unstable one with β = 4.05 and P = 48.949. (d) A typical example of the stable rhombic vortex with β = 4.4 and P = 42.562.
Fig. 8
Fig. 8 (Color online) (Upper row) and (bottom row): Examples of stable square-shaped and rhombic vortices (the amplitude and phase distributions) found in the narrow second bandgap (9) for DC = 0.75, with β = 8.35 and P = 5.9879 or P = 6.0108, respectively.
Fig. 9
Fig. 9 (Color online) (a) The P(β) dependence for rhombus-shaped vortices in the finite bandgap at DC = 0.5 in the model with the square-shaped transverse structure and self-defocusing nonlinearity. (b) An example of the amplitude and phase distribution in the stable vortex for β = 2.4 and P = 26.1702.
Fig. 10
Fig. 10 (Color online) The power-vs.-propagation constant curves for the fundamental solitons in two finite bandgaps of the photonic crystal with the rhombic transverse structure (see Fig. 2) and self-defocusing nonlinearity. In the second bandgap, the top branch is unstable, while the two lower ones are stable. (b) The largest instability growth rate of the gap solitons as functions of the propagation constant, β. The solid and dotted curves in the first bandgap in (b) correspond, respectively, to the bottom and top portions of the P(β) curve in (a). The curve in the second bandgap in (b) corresponds to the top P(β) branch in the same bandgap in (a). A typical example of stable solitons in the first bandgap is shown in panel (c) for β = 3.52 and P = 60.00. (d) Examples of unstable solitons with β = 3.9 in the first bandgap for P = 8.5687 and P = 26.6719 (top and bottom rows, respectively).
Fig. 11
Fig. 11 (Color online) (a) and (b): The evolution of unstable solitons corresponding to the top and bottom rows in Fig. 10 is shown in the cross section of |Ψ(x,y)|2 drawn through y = 0. (c): The same for the unstable pattern from Fig. 12(b).
Fig. 12
Fig. 12 (Color online) Typical examples of solitons, found in the second finite bandgap of the photonic crystal with the rhombic structure and self-defocusing nonlinearity, for β = 2.1. In (a), the top and bottom rows display examples of stable solitons, with P = 5.9460 and P = 13.4476, which correspond, respectively, to the lower and middle P(β) curves in the second bandgap in Fig. 10. (b) A typical example of an unstable soliton with P = 53.0894, which corresponds to the upper curve in the second bandgap in Fig. 10.
Fig. 13
Fig. 13 (Color online) Examples of unstable vortices found, respectively, in the first and second finite bandgaps of the model with the rhombic structure, with β = −4.25, P = 30.4432 (a), and β = 2.1, P = 64.2146 (b).
Fig. 14
Fig. 14 (Color online) (a) The power-vs.-propagation-constant dependence for solitons supported by the self-focusing nonlinearity in the model with the rhombic transverse structure. (b) A typical example of an unstable soliton, with β = 8.0 and P = 48.0398. (c) Decay of this soliton in direct simulations.

Equations (12)

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

i Ψ z + Ψ x x + Ψ y y + W ( x , y ) ( 1 + σ | Ψ | 2 ) Ψ = 0 .
W ( x , y ) = { U , L n < x < D + L n and L n < y < D + L n , 0 , D + L n < x < L ( 1 + n ) and / or D + L n < y < L ( 1 + n ) ,
β Φ + Φ x x + Φ y y + W ( x , y ) ( Φ + σ Φ 3 ) = 0 .
P ( β ) + + | Ψ ( x , y ) | 2 d x d y ,
[ ( x + i k x ) 2 + ( y + i k y ) 2 + W ( x , y ) ] φ ( x , y ) = β φ ( x , y )
Ψ ( x , y , z ) = e i β z { Φ 0 ( x , y ) + [ v ( x , y ) w ( x , y ) ] e λ z + [ v * ( x , y ) w * ( x , y ) ] e λ * z }
L ψ = λ ψ ,
L = i ( G 0 2 + G 1 2 + G 2 G 0 ) , ψ = ( v w ) G 0 = 1 2 ( Φ 0 2 Φ 0 * 2 ) σ W ( x , y ) G 1 = β + W ( x , y ) + 2 σ W ( x , y ) | Φ 0 | 2 1 2 σ W ( x , y ) ( Φ 0 2 + Φ 0 * 2 ) G 2 = β + W ( x , y ) + 2 σ W ( x , y ) | Φ 0 | 2 + 1 2 σ W ( x , y ) ( Φ 0 2 + Φ 0 * 2 )
8.315 < β < 8.398 ,
4.15 < β < 4.00 , 24.14 < P < 25.69 ,
3.70 < β < 3.50 , 30.52 < P < 75.20 ,
4.30 < β < 4.27 , 10.35 < P < 10.63.
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.