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

Nonlinear polarization instability in cubic-quintic plasmonic nanocomposites

Open Access Open Access

Abstract

We report a study of the nonlinear birefringence induced in a metal-dielectric nanocomposite due to the contributions of third- and fifth-order optical nonlinearities. A theoretical model describing the evolution of the light polarization state of a confined laser beam propagating through the nonlinear medium is developed with basis on a pair of coupled dissipative cubic-quintic nonlinear differential equations related to the two orthogonal polarizations of the optical field. As a proof-of-principle experiment we demonstrate the control of the light beam polarization in a silver-nanocolloid by changing the silver nanoparticles volume fraction, f, and the light intensity. A large nonlinear phase-shift (~20π) was observed using a 9 cm long capillary filled with silver nanoparticles suspended in carbon disulfide. Experiments using colloids with 1.0×105f4.5×105 and maximum light intensities of tens of MW/cm2 are performed. In addition, we demonstrate that the modulation instability is highly sensitive to the quintic nonlinearity contribution performed showing good agreement with the experimental results.

© 2017 Optical Society of America

1. Introduction

The nonlinear optical polarization induced by a laser beam propagating through a medium with inversion symmetry can be described by an expansion in powers series of the electric field where the even-order terms are null, unless the symmetry is disturbed. The nonzero expansion coefficients, χ(2N+1), with N=1,2,, called as (2N+1)th order susceptibility, provide a measure of how strong is the light-matter interaction [1]. The majority of articles in the literature are related to the cubic nonlinearity, due to χ(3), which is the parameter that describes phenomena such as third-order self-focusing, coherent anti-Stokes Raman scattering and two-photon absorption, among other phenomena [2]. However, in several cases, the odd higher-order nonlinearities (HON) are essential for characterization of the nonlinear response [3–11]. For instance, novel effects such as liquid light condensates [12], high-harmonic generation [13], filamentation [14,15], stable propagation of two dimensional spatial solitons in isotropic media [16–18] and optical rogue waves generation [19] are associated to HON. In addition, the HON contributions are essential for modelling three-body interactions in Bose-Einstein condensates [20,21], Bose gas with hard-core contact interactions and spin-polarized Fermi gas in the collisional regime [22], higher-order modulation instability [23], stability conditions for optical solitons [24,25], management of the solitary-wave solutions of the nonlinear Schrödinger equation [26] and several other physical situations.

One important consequence of HON is the nonlinear birefringence effect which is manifested by the polarization rotation of an elliptically polarized light beam along its propagation pathway [27]. This effect can be very much influenced by HON when the highly nonlinear materials currently available are excited with moderate laser intensities or even when gases are excited with intense laser pulses [3–11,13,28]. For instance, in [28], the authors were able to identify HON contributions for the nonlinear birefringence in air and determined the nonlinear refractive indices associated to third-, fifth-, seventh-, ninth- and eleventh-order susceptibilities of the main air components (nitrogen, oxygen and argon). Subsequently, a simple formalism was presented for precise characterization of the non-resonant nonlinear birefringence [29]. Moreover, HON play an important role on the modulation instability effects, at moderate or high light intensities, since their contributions are quite sensitive to any perturbation of the optical field. Indeed, recent studies on metamaterials predict that the modulation instability gain is enhanced due to the simultaneous contributions of HON (such as quintic nonlinearity) or saturable nonlinearity and high-order dispersions [30–32]. Those studies show relevant results related to solitons generation due to the modulation instability effects induced by HON.

Nowadays, it is well known that further advances in the telecommunications area require in-depth studies of light propagation in highly nonlinear optical fibers [33]. Indeed, the third-order nonlinear birefringence has been widely studied with important applications in all-optical devices [34–36]. However, an important limitation of optical fibers in all-optical circuits is related to the long propagation distances required to obtain induced polarization rotation. To mitigate this problem, highly nonlinear-core waveguides were developed allowing an increase of the polarization rotation of the coupled light field with small propagation distances [37]. A particular example was reported using an optical capillary where the core was filled with nitrobenzene that is highly nonlinear [38]. Therefore large nonlinear phase shifts of Δϕ(NL)12π in the infrared were obtained for propagation distances of ~10 cm.

On the other hand, metal-dielectric nanocomposites (MDNCs) consisting of a dielectric host containing metal nanoparticles emerged as nonlinear systems with high optical susceptibility and ultrafast response [8–11,39–42] that can be used as excellent platforms to study many associated HON phenomena. The nonlinear response of the MDNCs is described in terms of effective nonlinear susceptibilities that include contributions of the metal nanoparticles and their host. In particular, silver-nanocolloids have been studied under different conditions, by varying intrinsic parameters (such as: size, shape, environment and volume fraction of metal NPs) and extrinsic parameters (such as: laser wavelength, intensity, pulses duration and repetition rate), exhibiting important HON contributions. The exploitation of HON allowed enhancement of the nonlinear response [42,43] and even suppression of nonlinear absorption effects [44] by using a nonlinearity management procedure, which consists in varying the nanoparticles volume fraction, f, and the incident laser intensity [42–44]. In this sense, metal-nanocolloids are strong candidates to investigate the contributions of HON to nonlinear birefringence.

In the present paper, we demonstrate large nonlinear phase-shifts contributed by the third- and fifth-order nonlinearities when picosecond laser pulses propagate in a 9 cm long capillary filled with silver nanoparticles suspended in liquid carbon disulfide (CS2). Control of the nonlinear response was obtained by varying the nanoparticles volume fraction from f=1.0×105 to f=4.5×105, obtaining Δϕ(NL)20π for peak intensities of tens of MW/cm2. In this range of nanoparticles concentration and laser intensity, silver-nanocolloids behave like cubic-quintic media where, depending on the values of f, the nonlinear response is dominated by the third- or fifth-order nonlinearity. The experimental results were modeled by two coupled differential nonlinear equations that describe the evolution of the right- and left-circular optical field polarization in the samples. The gain spectra of the modulation instability were obtained, showing that the instabilities increase significantly due to the presence of the quintic nonlinearity.

The paper is organized as follows: Section 2 describes the samples preparation and identifies their linear and nonlinear optical parameters, as well as describes the experimental setups used. Section 3 presents details of the theory used for modeling the experimental results. Section 4 presents the experiments and numerical simulations made. A summary of the results is presented in Section 5.

2. Experimental details

Silver-nanocolloids were prepared following the chemical synthesis route described in [42–44]. As a result, a pristine colloid was obtained with a non-homogeneous size distribution of silver nanoparticles. Subsequently, the pristine colloid was subjected to laser ablation using the second harmonic beam from a Q-switched Nd:YAG laser (10 Hz, 8 ns, 85 mJ/pulse) for 1 hour in order to obtain a homogeneous nanoparticles size distribution by photofragmentation. During the laser ablation process the colloid was slowly stirred according to the procedure described in [45].

Optical absorbance spectra of the samples, before and after photofragmentation of the original nanoparticles synthesized, were measured using a commercial spectrophotometer. The absorption feature related to the localized surface plasmons resonance (LSPR) centered at ~400 nm was observed. As in our previous experiments [42–44] we observed a smaller LSPR linewidth after photofragmentation that indicates a homogeneous distribution of the silver nanoparticles sizes. This was corroborated by transmission electron microscopy that showed a homogeneous distribution of spherical nanoparticles with average diameter of (6.0±3.0)nm.

The setup used to study the intensity-dependent birefringence is illustrated by Fig. 1(a). The second harmonic of a Q-switched and mode-locked Nd: YAG laser (532 nm, 80 ps, 10 Hz) was used to excite the samples with maximum pulse energy of 10 µJ. A system composed of a Glan prism (P) located between two λ/2 plates was used to control the incident power and to rotate the polarization axis of a linearly polarized laser pulse by the azimuth angle θ. A 40x microscope objective (L1) was used to couple the laser light into a fused-silica hollow capillary (ncladding=1.46) with inner (outer) diameter of 2 µm (285 µm). The capillary with length of 9 cm had at the entrance and exit faces two reservoirs with the input and output sides of the capillary touching the reservoirs windows. The silver-nanocolloid was filled in one of the reservoirs, and using a strong air pressure the colloid was conveyed through the capillary to the other reservoir, in order to guarantee that the capillary is completely filled with the metal colloid (no air bubbles were observed in the end of the filling process). A 20x microscope objective (L2) was used to collimate the laser light after the propagation along the capillary. The laser beam was split by a polarizing beam splitter cube (PBS) to separate the vertical (V) and horizontal (H) polarization components that were monitored by fast detectors D1 and D2, respectively. A reference detector (RD) was used to compensate for the laser intensity fluctuations. Figure 1(b) displays the inner diameter of the capillary, in a fragment of 5 mm, showing small asymmetries that induce linear birefringence in the sample. The inset of Fig. 1(b) shows an image of the capillary core (length: 1 mm), obtained using an optical microscope.

 figure: Fig. 1

Fig. 1 (a) The experimental setup: polarizer (P), beam splitter (BS), spherical lenses with f = 5 cm (L), 40x microscope objective (L1), 20x microscope objective (L2), polarizing beam splitter cube (PBS) and reference detector (RD). The transmitted light with vertical and horizontal polarization was captured in the fast detectors D1 and D2, respectively. (b) Inner diameter of capillary, in a portion of 5 mm, showing small asymmetries. The inset is an optical microscope image of a small section of the hollow capillary core (length: 1 mm).

Download Full Size | PDF

Two experiments were performed to study the nonlinear birefringence induced in the samples. In one experiment made to analyze the optical transmission as a function of the optical field polarization, the laser beam intensity was chosen (by adjusting the first λ/2 plate and the polarizer) and the incident polarization direction was rotated by the second λ/2 plate using a motorized rotation stage, with steps of ~3.6 degrees. The second experiment, aiming the investigation of the transmittance intensity dependence was performed by rotating the first λ/2 plate using another motorized rotation stage with intensity steps of ~350 kW/cm2, maintaining fixed the incident electric field polarization.

In order to evaluate the contributions of the third- and fifth-order susceptibilities on the polarization instability effect, we used pure CS2 (host) and silver-nanocolloids with four different volume fractions, managed to display specific nonlinear responses. The corresponding values of α0, β(2), χxxxx(3) and χxxxxxx(5) are displayed in Table 1, as obtained from previous experiments [42–44]. Sample A (S-A), corresponding to pure CS2 (f=0), exhibits only focusing third-order nonlinearity, for the intensities used in this work. The other four samples (S-B, S-C, S-D and S-E), containing silver nanoparticles with 1.0×105f4.5×105, display contributions of third- and fifth-order susceptibilities which depend on the f value [42–44]. Notice that S-B represents a cubic-quintic (self-focusing) medium where the nonlinear refraction is dominated by the positive signal of Re[χxxxx(3)]. On the other hand, S-D and S-E correspond to cubic-quintic (defocusing-focusing) media because their nonlinear responses exhibit a negative value of Re[χxxxx(3)] and positive value of Re[χxxxxxx(5)]. However, S-C exemplifies a pure refractive quintic medium since Re[χxxxx(3)]=0, due to the destructive interference of the third-order susceptibilities of the host and the nanoparticles [42,43]. The change from the self-focusing to self-defocusing behavior in each case is due to the balance between the contributions of positive third-order refractive index of the host (CS2) and negative third-order refractive index of the silver nanoparticles, as detailed in [43].

Tables Icon

Table 1. Linear absorption coefficienta, second-order dispersion coefficienta, third and fifth-order susceptibilitiesb for pure CS2 (S-A) and silver-colloids with different nanoparticles volume fraction.

3. Nonlinear birefringence in a capillary filled with an isotropic cubic-quintic medium

3.1 General description

The complex electric field associated with an optical wave with arbitrary polarization can be written as

E(r,t)=12(x^Ex+y^Ey)exp(iω0t)+c.c,
where Ex and Ey are the complex amplitudes of the field components in the x- and y-axis with the carrier frequency ω0.

The nonlinear light propagation inside the colloid hosted by the capillary can be modeled by using the Helmholtz equation given by

2E(r,ω)+k02ε(ω)E(r,ω)=0,
where E and ε represent the frequency-dependent field and the dielectric tensor, respectively. 2 is the Laplacian operator, k0=ω0/c, ω0 is the laser frequency and c is the speed of light in vacuum.

As follows from Eqs. (1) and (2) the field components obeys the Helmholtz equation:

2Eμ(r,ω)+k02εμσ(ω)Eσ(r,ω)=0,
with μ and σ equal to x or y. In the principal-axis of the system, the dielectric tensor is represented as a diagonal matrix (εμσεμσδμ,σ), with εxx=εx and εyy=εy being the only nonzero components. This complex dielectric function (εμ) is associated to the refractive index, nμ, and the absorption coefficient, αμ, of the material through the expression εμ=(nμ+iαμ2k0)2. For media exhibiting nonlinear response, the dielectric function can be expressed in terms of the linear and nonlinear contributions as εμ(ω)=εμL(ω)+εμNL(ω), where the nonlinear term is related to the nonlinear optical polarization by:
PμNL(ω)=ε0εμNL(ω)Eμ(ω),
where εμNL(ω) contains the contributions of third- and higher-order susceptibilities, as can be observed below. In particular, silver nanoparticles suspended in liquid CS2 behave as a cubic-quintic medium in the experimental conditions used here, according to [42,43], and therefore, the nonlinear polarization is described by the sum of the third- and fifth-order contributions, PμNL=Pμ(3)+Pμ(5),with:
Pμ(3)(ω)=3ε0{2χxxyy(3)(ω)Eμ(ω)[E(ω)E*(ω)]+χxyyx(3)(ω)Eμ*(ω)[E(ω)E(ω)]},
Pμ(5)(ω)=10ε0{103χxxyyxx(5)(ω)|Eμ(ω)|2[E(ω)E(ω)]Eμ*(ω)+53χxxyyyy(5)(ω)σ=x,y|Eσ(ω)|4Eμ(ω)},
where χ(3) and χ(5) are the third- and fifth-order susceptibilities. A detailed explanation how to obtain Eqs. (5) and (6), as well as higher-order nonlinear polarizations are given in [29]. By considering Eqs. (4)-(6) we obtain an expression for the dielectric function in the form:

εμNL(ω)=3[(2χxxyy(3)(ω)+χxyyx(3)(ω))|Eμ(ω)|2+2χxxyy(3)(ω)σ=x,y|Eσ(ω)|2(1δμ,σ)]+10[103χxxyyxx(5)(ω)|Eμ(ω)|4+53χxxyyyy(5)(ω)(|Eμ(ω)|4+σ=x,y|Eσ(ω)|4(1δμ,σ))]+[3χxyyx(3)(ω)+10103χxxyyxx(5)(ω)|Eμ(ω)|2]σ=x,y[Eσ(ω)]2(1δμ,σ)Eμ*(ω)Eμ(ω).

The Helmholtz equation [Eq. (3)] can be solved by the separation of variables method for each component of the electric field:

Eμ(r,ω)=F(x,y)Aμ(z,ωω0)exp(iβ0,μz),
where F(x,y) is the transverse mode pattern supported by the capillary, Aμ(z,ωω0) is the slowly varying amplitude and β0,μ=βμ(ω0) is the propagation constant for μ=x,y. Substituting Eq. (8) in Eq. (3) and separating the terms dependent upon the propagation distance (z) and the transverse coordinates (x,y), we obtain
2iβ0,μAμz+(βμ2β0,μ2)Aμ=0,
2Fx2+2Fy2+[εμ(ω)k02βμ2]F=0,
where the slowly varying envelope approximation [(2Aμ/z2)<<β0,μ(Aμ/z)] was made to obtain Eq. (9). βμ=βμ(ω) represents the frequency-dependent wavenumber, where the zero-order term of the series corresponds to β0,μ. The dielectric function in Eq. (10) can be approximated by
εμ=(n0,μ+ΔNμ)2(n0,μ)2+2(n0,μ)(ΔNμ),
where ΔNμ=iα0,μ/(2k0)+[Δnμ+iΔαμ/(2k0)] contains a dissipative term and a small perturbation induced by the nonlinearity, n0,μ(α0,μ) is the linear refractive index (linear absorption coefficient) and Δnμ(Δαμ) represents the contributions of the nonlinear refractive index (nonlinear absorption coefficient). It is important to note that HON are included in the first order perturbation, ΔNμ, as shown in Eq. (12). However (ΔNμ)2 and high-order perturbation were neglected in Eq. (11) because, for the intensities used in this work, the contributions to the linear dielectric function are negligible since [(ΔNμ)2 is ~5 orders of magnitude smaller than (ΔNμ)]. The linear and nonlinear dielectric functions are given by εμL=[n0,μ+iα0,μ/(2k0)] and εμNL=2(n0,μ)[Δnμ+iΔαμ/(2k0)], and by substituting in Eqs. (7) and (8), we have
(Δnμ+iΔαμ2k0)=32n0,μ|F|2[(2χxxyy(3)+χxyyx(3))|Aμ|2+2χxxyy(3)σ=x,y|Aσ|2(1δμ,σ)]+102n0,μ|F|4[103χxxyyxx(5)|Aμ|4+53χxxyyyy(5)(|Aμ|4+σ=x,y|Aσ|4(1δμ,σ))]+12n0,μ|F|2[3χxyyx(3)+10103χxxyyxx(5)|F|2|Aμ|2]σ=x,y[Aσ(ω)]2(1δμ,σ)Aμ*Aμexp[2i(β0,σβ0,μ)z].
βμ can be determined by solving Eq. (10) using the first-order approximation (βμβμ+Δβμ), where F and the eigenvalue βμ2βμ2+2βμ(Δβμ) are represented by
F=F(0)+ξF(1)+,
βμ2=(βμ)2+ξ2βμ(Δβμ)+,
where ξ is a perturbation parameter which can vary over a continuous range of values from 0 (no perturbation) to 1 (full perturbation). The term (Δβμ)2 was disregarded by the same reason as (ΔNμ)2, as mentioned earlier. Inserting Eqs. (13) and (14) into the eigenvalue Eq. (10), and using Eq. (11), we obtain

ξ0:[(2x2+2y2+n0,μ2k02)βμ2]F(0)=0,
ξ1:[(2x2+2y2+n0,μ2k02)βμ2]F(1)+[2(n0,μ)(ΔNμ)k022βμΔβμ]F(0)=0.

In order to find an expression for F(0), Eq. (15) was written in cylindrical coordinates (ρ,ϕ,z) and by using the separation of variables F(0)(ρ,ϕ)=F(0)(ρ)exp(imϕ), it was possible to recognize Eq. (15) as the modified Bessel differential equation: 2F(0)ρ2+1ρF(0)ρ+(n0,μ2k02βμ2m2ρ2)F(0)=0. Then, a solution valid for the region filled with the nonlinear medium, corresponding to the capillary core (ncore=n0,μ) is

Fcore(0)(ρ)=C1Jm(ρn0,μ2k02βμ2),
where Jm(x) is the first kind m-order Bessel function. For the region formed by the capillary wall (n0,μ=ncladding), F(0) should decay exponentially with ρ. A function that describes such behavior is the modified Bessel function of second kind, given by

Fcladding(0)(ρ)=C2Km(ρβμ2ncladding2k02).

The values of C1 and C2 are determined by considering the boundary conditions.

From Eq. (16), corresponding to the first-order of perturbation, we get

Δβμ=k0(ΔNμ)|F(0)|2dxdy|F(0)|2dxdy,
where F(0) is given by Eqs. (17) and (18), and ΔNμ was defined right after Eq. (11). Therefore, to the first order in the perturbation, the nonlinearity (ΔNμ) does not affect the modal distribution [Eqs. (17) and (18)], but modifies the wavenumber through Δβμ and consequently the evolution of the field amplitude, Aμ. In this way, by considering the first-order approximation of the wavenumber (βμβμ+Δβμ) into Eq. (9) we obtain the expression
Aμz=i(βμ+Δβμβ0,μ)Aμ,
where the approximation βμ2β0,μ22β0,μ(βμβ0,μ) was used. Nevertheless, as an exact functional form of βμ is generally unknown, an expansion in Taylor series around the frequency ω0 is made to obtain the more specialized expression
βμ(ω)=β0,μ+βμ(1)(ωω0)+12βμ(2)(ωω0)2+O[(ωω0)3],
where βμ(m)=(dmβμ/dωm)ω=ω0, m=1,2, represent the dispersion terms and the last term represents higher-order contributions that in the present case can be neglected. Then, substituting Eq. (21) in Eq. (20) and using Eqs. (12) and (19), we obtain two coupled differential equations that describe the evolution of the two polarization components along a capillary filled with a cubic-quintic isotropic media. Both equations can be summarized as:
Aμz+βμ(1)Aμt+iβ(2)22Aμt2+α02Aμ=ik02n0,μ3F(1){[(2χxxyy(3)+χxyyx(3))|Aμ|2+2χxxyy(3)σ=x,y(1δμ,σ)|Aσ|2]Aμ+χxyyx(3)[σ=x,y(1δμ,σ)(Aσ)2]Aμ*exp[2i(β0,σβ0,μ)z]}+ik02n0,μ10F(2){[103χxxyyxx(5)|Aμ|4+53χxxyyyy(5)[|Aμ|4+σ=x,y(1δμ,σ)|Aσ|4]]Aμ+103χxxyyxx(5)[σ=x,y(1δμ,σ)|Aμ|2(Aσ)2]Aμ*exp[2i(β0,σβ0,μ)z]},
with μ=x,y,F(1)=(|F(0)|4dxdy)/(|F(0)|2dxdy)2 and F(2)=(|F(0)|6dxdy)/(|F(0)|2dxdy)3. The modal birefringence of the capillary is expressed by Δβ0=β0,xβ0,y=2πλ|nxny|, which leads to different group velocities for the two polarization components (βx(1)βy(1)). However, β(2) and α0 are assumed to be the same for both polarization components.

For convenience, we rewrite Eq. (22) using the circularly polarized components A±=212(A¯x±iA¯y), where A¯x=Axexp(iΔβz/2) and A¯y=Ayexp(iΔβz/2), and after performing some manipulations we obtain:

A±z+12[β+(1)A±t+β(1)At]+i2β(2)2A±t2+α02A±=i2(Δβ0)A+i3ω0n0cF(1)[χxxyy(3)(|A+|2+|A|2)+χxyyx(3)|A|2]A±+i5ω02n0cF(2){103χxxyyxx(5)[|A++A|2(A++A)*|A+A|2(A+A)*]A+56χxxyyyy(5)[|A++A|4+|A+A|4]}A±,
with β±(1)=βx(1)±βy(1).

3.2 Nonlinear birefringence in capillaries filled with silver-nanocolloids

Silver nanocolloids exhibit fast electronic nonlinear response but their nonlinear optical properties depend on the shape, size and volume fraction of the silver nanoparticles, the laser frequency (which may excite resonant or non resonant processes), and the mismatch between the dielectric functions of the nanoparticles and their host [47–50]. In particular, the complex nonlinear susceptibilities, which correspond to non resonant processes of electronic origin, obey the relationship: χxxyy(3)=χxyyx(3)=1/3χxxxx(3) and χxxyyxx(5)=χxxyyyy(5)=1/5χxxxxxx(5) [29].

For the case of a capillary filled with silver nanocolloids, the induced birefringence is due to the refractive index shift (Δn) induced by the nanoparticles nonlinearity. For light peak intensity of tens of MW/cm2 (used in this work), we have Δn<105, corresponding to a weak birefringence. Therefore, we may assume βx(1)βy(1)=β(1), and since the difference between the group velocities, vg=[β(1)]1, in both directions is negligible, we can introduce vg=[β(1)]1 as an intrinsic variable by performing the transformation of Eq. (23) to the pulse reference frame (retarded frame) with (z,t)(z,τ=tz/vg). Thus, we obtain

A±z+i2β(2)2A±τ2+α02A±=i2(Δβ0)A+iω0n0cF(1)χxxxx(3)[(|A+|2+|A|2)+|A|2]A±+i5ω012n0cF(2)χxxxxxx(5){4[|A++A|2(A++A)*|A+A|2(A+A)*]A+[|A++A|4+|A+A|4]}A±,
where the dispersion coefficients for silver nanocolloids can be obtained by using the Maxwell-Garnett model [51]. Then, the effective dielectric function for a macroscopically isotropic medium is given by
εeff(λ,f)=εh(λ)[1+3Θ(λ)f1Θ(λ)f],
with Θ(λ)=[εNP(λ)εh(λ)][εNP(λ)+2εh(λ)]1, where εNP and εhare the linear dielectric functions of the silver nanoparticles and the host, respectively, and f is the volume fraction occupied by the silver nanoparticles. For the experiments reported in this article, liquid CS2 was the solvent (host) and its dielectric function according to [52] has the form

εh(λ)=[nCS2(λ)]2=[1.580826+1.52389×102λ2+4.8578×104λ4+8.2863×105λ6+1.4619×105λ8]2.

On the other hand, the wavelength-dependent dielectric function of the silver nanoparticles, in the visible light region, can be adequately described by using the Drude's free-electron model [53] given by

εNP(λ)=(1λ2λp2)+i(12πcτrλ3λp2),
where λp2=ρe2/πm0c2, ρ and e are the density and charge of the conduction electrons, m0 is their effective mass, and τr is the relaxation time of the electrons in the metal. Hence, the dielectric function can be written as εNP(λ)=(152.51λ2)+i(0.899λ3) where we consider the values of λp=0.138μm and τr=31fs obtained in [54] by fitting the reflection and transmission measurements of a silver thin film using the Drude free-electron theory. Therefore, by substituting Eqs. (26) and (27) in the effective dielectric function expression [Eq. (25)], and defining the effective refractive index by neff(λ,f)=212Re[εeff(λ,f)]+|εeff(λ,f)|, we obtain the dispersion coefficients

β(1)(λ,f)=1c[neff(λ,f)λd[neff(λ,f)]dλ],
β(2)(λ,f)=λ32πc2d2[neff(λ,f)]dλ2.

For instance, when λ=532nm and f=105 we determined β(1)=6.06ns/m and β(2)=0.717ps2/m. Since high-order dispersion of silver nanocolloids is not reported in the literature its contribution was neglected in the deduction of Eq. (24).

An important point to highlight here is that the Z-scan experiments, previously performed for the nonlinear characterization of silver nanocolloids, were limited to a single wavelength. Therefore, in our analysis it is not possible to identify the type of non resonant processes contributing to the effective third- and fifth-order susceptibilities, shown in Table 1. Thus, for our particular case, the second and third term of the right hand in the Eq. (24) represent the contributions of all possible non resonant cubic effects such as stimulated Raman, Rayleigh-wing scattering as well as non resonant fifth-order effects.

3.3 Linear Stability Analysis

In order to analyze the effects of temporal-modulation instability (TMI) of a short laser pulse propagating in a cubic-quintic medium we performed a typical linear perturbation analysis of the pulse amplitude. For simplicity, we assume that the field polarization is oriented along the fast axis and thus Ax=0. Therefore, considering a Gaussian laser pulse profile, the field amplitude is represented by

A±(z,τ)=±iB(z,τ)exp(α0z2iΔβ02z)exp(τ2τ02),
where B(z,τ)=B0exp[Λ(τ)z+iΦ(τ)z] is the unperturbed pulse amplitude. B0 is the incident power at z=0, τ0 is the initial pulse duration, Λ(τ) and Φ(τ) are given by

Λ(τ)=B02ω0c[3F(1)Im(χ(3))+20B02F(2)Im(χ(5))exp(2τ2τ02)]exp(2τ2τ02),
Φ(τ)=12β(2)[2τ02(12τ2τ02)]+B02ω0c[3F(1)Re(χ(3))+20B02F(2)Re(χ(5))exp(2τ2τ02)]exp(2τ2τ02).

The modulated amplitude of the laser pulse may be written as the superposition of perturbed and unperturbed amplitudes, as

a±(z,τ)=±i{B0exp[Λ(τ)z]+B1,±(z,τ)}exp(α0z2iΔβ02z)exp[iΦ(τ)z]exp(τ2τ02),
where B1(z,τ) is the complex perturbed beam amplitude. A linear stability analysis was performed in order to examine the evolution of the perturbation B1(z,τ) and the stability of the solution (33). Thus, by introducing Eq. (33) in Eq. (24) and linearizing as a function of B1,±(z,τ), we obtain

i{B1,±z+iΦB1,±i2(Δβ0)(B1,+B1,)}12β(2){2B1,±τ2+2[izΦτ2ττ02]B1,±τ+izB1,±[2Φτ24ττ02Φτ+iz(Φτ)2]+2τ02[2τ2τ021]B1,±}=ω0c|B0|2F(1)χxxxx(3){5B1,±+2[B1,++B1,]}exp(2Λzα0z2τ2τ02)5ω012cF(2)χxxxxxx(5){16|B0|4[3B1,±+2(B1,++B1,+B1,)]+48|B0|2B02[B1,++B1,]*}exp(4Λz2α0z4τ2τ02).

Solution of Eq. (34) were found by neglecting further variations in the pulse shape and considering the perturbed wave amplitude to be a sinusoidal function of z and τ, that is, B1,±=u±exp[i(KzΩτ)]+iv±exp[i(KzΩτ)], with K and Ω being the wave number and frequency of the perturbed wave amplitude, respectively. Nontrivial solutions are obtained only when the perturbation satisfies the following dispersion relation

K2+[2β(2)Ω(zΦτ)]KM=0,
where

M=Re(N2)2β(2)Ω[zΦτRe(N)2ττ02Im(N)]+[40ω0cB02|B0|2F(2)exp(4Λz2α0z4τ2τ02)]2Re[(χxxxxxx5)2],
N=Φ+12β(2){Ω[Ω+2i(izΦτ2ττ02)]+iz[2Φτ24ττ02Φτ+iz(Φτ)2]+2τ02[2τ2τ021]}3ω0c|B0|2exp(2Λzα0z2τ2τ02)[3F(1)χxxxx(3)+20|B0|2F(2)χxxxxxx(5)exp(2Λzα0z2τ2τ02)].

The local gain spectra of TMI as a function of the frequency shift along the fast axis was obtained from the imaginary part of K [g(Ω,z)=2Im(K)], as illustrated in Fig. 2. We remark that the local gain depends on the z position because the field amplitude varies along the propagation distance due to dissipative terms. The TMI gain curves were calculated for a peak intensity of 42MW/cm2 that corresponds to the pulse intensity at τ=0.1τ0, after a propagation distance of 1.5 mm. In the normal-dispersion regime (β(2)>0), the S-A sample (self-focusing cubic medium) presents a very small modulation instability gain compared to the other samples, and for this reason its gain spectrum is not shown in Fig. 2. In fact, numerical simulations were performed for pulse propagation with distance of 9 cm (capillary length) showing that the modulation instability effect inside S-A is negligible, as shown below (see Fig. 3). However, under the same conditions, side-bands in the local gain spectrum of S-B, which is also a focusing medium, are observed displaying a growth of the modulation instability attributed to the quintic nonlinearity which is positive and larger than in S-A. The local gain curve obtained for S-C (focusing refractive quintic media without cubic contribution) clearly corroborates that the growth rate of TMI strongly depends on the quintic nonlinearity contribution, since the refractive cubic nonlinearity is null, as indicated in Table 1. Although in the S-D and S-E, the refractive cubic nonlinearity is negative (defocusing), the refractive quintic nonlinearity dominates the effects of modulation instability since its contribution is enhanced due to the larger dependence with the electric field. As in the case of metamaterials discussed in [30] distortions in the sidebands are due to the linear losses, but enhanced by nonlinear absorption contributions. To obtain an analytical expression for the total gain accumulated after a propagation distance L, it is necessary to integrate the local gain with respect to z in the interval from 0 to L [55]. Notice however that by performing the numerical simulation based on the nonlinear propagation equations this was automatically considered.

 figure: Fig. 2

Fig. 2 Local gain spectra of modulation instability versus the frequency shift along the fast axis, for the five samples.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Numerical pulse shape evolution for sample A (S-A), sample B (S-B), sample C (S-C), sample D (S-D) and sample E (S-E), with input intensity of 60 MW/cm2. Pulse duration: 80 ps. Propagation length: 9 cm.

Download Full Size | PDF

4. Results and discussions

The evolution of the x- and y-components of the electric field, influenced by the nonlinear birefringence of the 9 cm long capillary filled with silver nanoparticles suspended in CS2, was analyzed by solving numerically the two coupled equations [Eq. (24)], using the compact finite-difference method based on the Crank–Nicolson scheme [56].

Figure 3 shows numerical results that represent the pulse shape evolution, for the five samples, with input intensity of 60 MW/cm2. For all cases, the input beam has a Gaussian profile with pulse duration of 80 ps. A temporal window of 400 ps, with steps of 0.1 ps, was used for the numerical simulations. For the S-A sample (black line) the pulse propagates keeping its Gaussian shape, but suffering a small broadening from 80 ps to ~89 ps. Due to the small nanoparticles concentration the contribution of the fifth-order nonlinearity is negligible in this case. On the other hand, the pulse propagation inside sample S-B (red line) is influenced by TMI, in accordance with the gain spectrum obtained in Fig. 2. In this case the nonlinear refractive index n2Re[χxxxx(3)] of S-B is smaller than for sample S-A [see Table 1] and the main contribution for TMI is due to n4Re[χxxxxxx(5)], as previously deduced from the linear stability analysis. To corroborate our interpretation, we studied the pulse propagation in a pure quintic refractive medium (blue line), corresponding to sample S-C. The TMI effect increased due to the larger n4 than in the samples S-A and S-B. Moreover, larger TMI effect is observed in the samples S-D (green line) and S-E (pink line), corresponding to more concentrated silver nanocolloids. Intensity losses along the propagation, in all samples, are due to linear and nonlinear absorptions. It is important to emphasize that both nonlinear refraction and absorption terms contribute to the stable or unstable propagation of optical fields, under suitable conditions. For instance, the nonlinear absorption, i.e. the imaginary part of the susceptibilities, contributes to the stability of a beam propagating in a self-focusing medium (S-A and S-B) by arresting the catastrophic collapse, as shown in [16,17]. Moreover, in defocusing cubic media (S-C, S-D and S-E), the diffraction of a beam is compensated by the focusing quintic nonlinearity [18].

Figure 4 shows the normalized transmittance as a function of the incident polarization azimuth angle, θ, for all samples. From top to bottom, the incident peak intensities correspond to 6, 24, 42 and 60 MW/cm2, in each row. For I ≤ 6 MW/cm2 [first row of Fig. 4], all samples behave as linear isotropic media. As a consequence, the normalized experimental transmittance exhibits a cos2θ dependence (black circles) for the vertical (V)-polarization (captured in the detector D1) and a sin2θ dependence (red squares) for the horizontal (H)-polarization (captured in the detector D2). Black and red lines represent the normalized transmittance for the V- and H-polarization, obtained by numerical solution of Eq. (24) and using the coefficients of Table 1. At I = 24 MW/cm2 [second row of Fig. 4], the transmittance response as a function of θ shows significant nonlinear contributions, dominated by third-order nonlinearity. For the cubic and cubic-quintic self-focusing samples, S-A and S-B, respectively, the transmittance response displays a lower variation with θ which increases with χ(3), in comparison to the first row of Fig. 4. The quintic self-focusing sample (S-C) transmittance shows small variation with θ at 24 MW/cm2 because the contributions of χ(3) is null and χ(5) is very small. In contrast, the cubic-quintic (defocusing-focusing) samples (S-D and S-E) show a larger dependence as a function of the incident polarization azimuth angle. With the increase of the incident intensity, the refractive index variation between the slow and fast axis of the capillary increases or decreases depending of the sign of total nonlinear susceptibility and the direction of the incident field polarization in the transverse plane. In this way, rotation of the incident polarization direction, at high intensities, generate multiple regions where the variation of the nonlinear birefringence increases or decreases the polarization instability effects. As a consequence, regions of small and large response are simultaneously observed for high intensities by varying the incident polarization azimuth angle, as shown in the third and fourth rows of Fig. 4 that corresponds to intensities of 42 MW/cm2 and 60 MW/cm2, respectively. Note that in the last two rows of Fig. 4, the fifth-order contribution is very important to increase the modulation instability effect, in agreement with the Fig. 2. Numerical simulations of Eq. (24), represented by the solid lines, were made using as initial condition a 80 ps Gaussian pulse. However, the relative orientation of the capillary fast- and slow-axis in relation to the laboratory frame (x, y) was treated as a free parameter. Values of Δβ0 between 0.07 and 0.12, which correspond to |nxny|106, were used to obtain a better fit of the experimental data. These values are reasonable since the refractive index variation produced by the nonlinearities, Δn=n2I+n4I2, are of the same order of magnitude.

 figure: Fig. 4

Fig. 4 Normalized transmittance as a function of the incident polarization azimuth angle, θ, for (a) sample A, (b) sample B, (c) sample C, (d) sample D and (e) sample E. From top to bottom, the incident peak intensities are 6, 24, 42 and 60 MW/cm2, in each row. Black circles and red squares correspond to vertical and horizontal polarization transmittance, respectively. Solid lines were obtained from numerical solutions of Eq. (24) and the parameters of Table 1.

Download Full Size | PDF

Figure 5 shows the transmittance behavior, in the V-polarization (black circles) and H-polarization (red squares), for incident peak intensities between 0.1 MW/cm2 and 70 MW/cm2. An incident azimuth angle of 43° in relation to V-polarization was used for all samples. For non-birefringent materials or materials exhibiting only linear birefringence, the transmittance remains constant for different intensities. However, Fig. 5(a), corresponding to S-A (pure CS2), shows modulation of the transmittance response with increasing of the intensity, induced by the cubic nonlinearity. Each oscillation observed in the transmittance signal corresponds to ~2π phase-shift. It is possible to observe that with the addition of silver nanoparticles [Figs. 5(b)–5(e)], the modulation increases due to the large effective nonlinear susceptibility of the samples. It is worth noting that the fifth-order nonlinearities are essential for the increasing of the modulation, as shown in Fig. 5(c), which corresponds to a refractive quintic medium with Re[χ(3)]=0. A maximum nonlinear phase shift of ~20π was observed for intensities of 70 MW/cm2, using the S-E sample, as shown in Fig. 5(e). The black and red solid lines represent the numerical simulations of Eq. (24), showing a good agreement with the experimental results. The blue dashed lines in Figs. 5(b)–5(e) display the transmittance behavior neglecting the Re[χ(5)] contribution, showing that the modulation due to the nonlinear birefringence is highly modified due to the fifth-order nonlinearity.

 figure: Fig. 5

Fig. 5 Vertical (black circles) and horizontal (red squares) polarization transmittance as a function of the incident peak intensities variation, for (a) sample A, (b) sample B, (c) sample C, (d) sample D and (e) sample E. Black and red solid lines were obtained from numerical solutions of Eq. (24) and the parameters of Table 1. Blue dashed lines represent the numerical solutions of Eq. (24) neglecting the contribution of Re(χ(5)).

Download Full Size | PDF

Small discrepancies between the experimental and theoretical results are due to the capillary used for all experiments that supports few propagation modes (V-number is ~9), while in the theoretical description it was assumed, for simplicity, that the propagation occurs in a single-mode capillary. Therefore, coupling effects between the different spectral modes were neglected in the numerical simulations. The free parameters used in the simulation, such as the focusing angle of the beam entering in the capillary and the linear refractive index dependence with f are not expect to contribute for relevant discrepancies between the experimental and numerical results.

5. Summary

In summary, the nonlinear birefringence effect due to cubic and quintic nonlinearities was investigated in a 9 cm long capillary filled with silver nanoparticles suspended in CS2. For a given optical field amplitude the intensity-dependent birefringence was varied by changing the volume fraction occupied by the silver nanoparticles from 0 (pure CS2) to 4.5×105. Two experimental schemes were used to analyze the transmittance response as a function of the incident polarization azimuth angle and the light intensity, in order to identify the contributions of the third- and fifth-order susceptibilities on the polarization instability effect. To compare with the experimental results, a model describing the evolution of the optical field polarization was developed considering the refractive and dissipative contributions due to the third- and the fifth-order susceptibility. The model describes how the gain spectrum of the modulation instability increases significantly with the presence of the quintic nonlinearity. In addition, numerical simulations were performed considering the dispersion coefficients as well as the linear and nonlinear susceptibilities for the colloids studied, showing good agreement with the experimental results.

Finally we want to recall that the nonlinear birefringence effect is attractive for exploitation in all-optical switches. For the samples used in this work, a large nonlinear phase-shift (~20π) was observed in diluted silver-nanocolloids with f=4.5×105. However, more concentrated samples will present larger nonlinear phase-shift due to the increased third- and fifth-order nonlinearities and may present further interesting behavior not observed in the present experimental conditions. Also we recall that by using the nonlinearity management procedure of [42–44], it is possible to improve the figures-of-merit of all-optical switches based on metal-dielectric nanocomposites [44].

Funding

Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq); Fundação de Amparo à Ciência e Tecnologia do Estado de Pernambuco (FACEPE); National Institute of Photonics (INCT de Fotônica– INFO); PRONEX/CNPq/FACEPE; CAPES-COFECUB Program.

References and links

1. N. Bloembergen, Nonlinear Optics (W.A. Benjamin, 1965).

2. R. W. Boyd, Nonlinear Optics (Academic, 2008).

3. K. Dolgaleva, H. Shin, and R. W. Boyd, “Observation of a microscopic cascaded contribution to the fifth-order nonlinear susceptibility,” Phys. Rev. Lett. 103(11), 113902 (2009). [CrossRef]   [PubMed]  

4. N. J. Dawson and J. H. Andrews, “The local-field factor and microscopic cascading: a self-consistent method applied to confined systems of molecules,” J. Phys. At. Mol. Opt. Phys. 45(3), 035401 (2012). [CrossRef]  

5. P. Yosia and Shum, “Optical bistability in periodic media with third-, fifth-, and seventh-order nonlinearities,” J. Lightwave Technol. 25(3), 875–882 (2007). [CrossRef]  

6. C. Schnebelin, C. Cassagne, C. B. de Araújo, and G. Boudebs, “Measurements of the third- and fifth-order optical nonlinearities of water at 532 and 1064 nm using the D4σ method,” Opt. Lett. 39(17), 5046–5049 (2014). [CrossRef]   [PubMed]  

7. G. Boudebs, S. Cherukulappurath, H. Leblond, J. Troles, F. Smektala, and F. Sanchez, “Experimental and theoretical study of higher-order nonlinearities in chalcogenide glasses,” Opt. Commun. 219(1-6), 427–433 (2003). [CrossRef]  

8. E. L. Falcão-Filho, C. B. de Araújo, and J. J. Rodrigues Jr., “High-order nonlinearities of aqueous colloids containing silver nanoparticles,” J. Opt. Soc. Am. B 24(12), 2948–2956 (2007). [CrossRef]  

9. J. Jayabalan, “Origin and time dependence of higher-order nonlinearities in metal nanocomposites,” J. Opt. Soc. Am. B 28(10), 2448–2455 (2011). [CrossRef]  

10. J. Jayabalan, A. Singh, R. Chari, S. Khan, H. Srivastava, and S. M. Oak, “Transient absorption and higher-order nonlinearities in silver nanoplatelets,” Appl. Phys. Lett. 94(18), 181902 (2009). [CrossRef]  

11. S. Wang, Y. Zhang, R. Zhang, H. Yu, H. Zhang, and Q. Xiong, “High-order nonlinearity of surface plasmon resonance in Au nanoparticles: paradoxical combination of saturable and reverse-saturable absorption,” Adv. Opt. Mat. 3(10), 1342–1348 (2015). [CrossRef]  

12. H. Michinel, M. J. Paz-Alonso, and V. M. Pérez-García, “Turning light into a liquid via atomic coherence,” Phys. Rev. Lett. 96(2), 023903 (2006). [CrossRef]   [PubMed]  

13. D. L. Weerawarne, X. Gao, A. L. Gaeta, and B. Shim, “Higher-order nonlinearities revisited and their effect on harmonic generation,” Phys. Rev. Lett. 114(9), 093901 (2015). [CrossRef]   [PubMed]  

14. M. Kolesik, E. M. Wright, and J. V. Moloney, “Femtosecond filamentation in air and higher-order nonlinearities,” Opt. Lett. 35(15), 2550–2552 (2010). [CrossRef]   [PubMed]  

15. Z. Wang, C. Zhang, J. Liu, R. Li, and Z. Xu, “Femtosecond filamentation in argon and higher-order nonlinearities,” Opt. Lett. 36(12), 2336–2338 (2011). [CrossRef]   [PubMed]  

16. E. L. Falcão-Filho, C. B. de Araújo, G. Boudebs, H. Leblond, and V. Skarka, “Robust two-dimensional spatial solitons in liquid carbon disulfide,” Phys. Rev. Lett. 110(1), 013901 (2013). [CrossRef]   [PubMed]  

17. A. S. Reyna, K. C. Jorge, and C. B. de Araújo, “Two-dimensional solitons in a quintic-septimal medium,” Phys. Rev. A 90(6), 063835 (2014). [CrossRef]  

18. A. S. Reyna and C. B. de Araújo, “Guiding and confinement of light induced by optical vortex solitons in a cubic-quintic medium,” Opt. Lett. 41(1), 191–194 (2016). [CrossRef]   [PubMed]  

19. B. G. O. Essama, J. Atangana, B. M. Frederick, B. Mokhtari, N. C. Eddeqaqi, and T. C. Kofane, “Rogue wave train generation in a metamaterial induced by cubic-quintic nonlinearities and second-order dispersion,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 90(3), 032911 (2014). [CrossRef]   [PubMed]  

20. F. Kh. Abdullaev and M. Salerno, “Gap-Townes solitons and localized excitations in low-dimensional Bose-Einstein condensates in optical lattices,” Phys. Rev. A 72(3), 033617 (2005). [CrossRef]  

21. I. Danshita, D. Yamamoto, and Y. Kato, “Cubic-quintic nonlinearity in superfluid Bose-Bose mixtures in optical lattices: Heavy solitary waves, barrier-induced criticality, and current-phase relations,” Phys. Rev. A 91(1), 013630 (2015). [CrossRef]  

22. A. Minguzzi, P. Vignolo, M. L. Chiofalo, and M. P. Tosi, “Hydrodynamic excitations in a spin-polarized Fermi gas under harmonic confinement in one dimension,” Phys. Rev. A 64(3), 033605 (2001). [CrossRef]  

23. M. Erkintalo, K. Hammani, B. Kibler, C. Finot, N. Akhmediev, J. M. Dudley, and G. Genty, “Higher-Order Modulation Instability in Nonlinear Fiber Optics,” Phys. Rev. Lett. 107(25), 253901 (2011). [CrossRef]   [PubMed]  

24. M. Shen, H. Zhao, B. Li, J. Shi, Q. Wang, and R. K. Lee, “Stabilization of vortex solitons by combining competing cubic-quintic nonlinearities with a finite degree of nonlocality,” Phys. Rev. A 89(2), 025804 (2014). [CrossRef]  

25. S. Loomba, R. Pal, and C. N. Kumar, “Bright solitons of the nonautonomous cubic-quintic nonlinear Schrödinger equation with sign-reversal nonlinearity,” Phys. Rev. A 92(3), 033811 (2015). [CrossRef]  

26. A. Choudhuri, H. Triki, and K. Porsezian, “Self-similar localized pulses for the nonlinear Schrödinger equation with distributed cubic-quintic nonlinearity,” Phys. Rev. A 94(6), 063814 (2016). [CrossRef]  

27. G. P. Agrawal, Nonlinear Fiber Optics (Academic, 2013).

28. V. Loriot, E. Hertz, O. Faucher, and B. Lavorel, “Measurement of high order Kerr refractive index of major air components,” Opt. Express 17(16), 13429–13434 (2009). [CrossRef]   [PubMed]  

29. G. Stegeman, D. G. Papazoglou, R. Boyd, and S. Tzortzakis, “Nonlinear birefringence due to non-resonant, higher-order Kerr effect in isotropic media,” Opt. Express 19(7), 6387–6399 (2011). [CrossRef]   [PubMed]  

30. M. Saha and A. K. Sarma, “Modulation instability in nonlinear metamaterials induced by cubic–quintic nonlinearities and higher order dispersive effects,” Opt. Commun. 291, 321–325 (2013). [CrossRef]  

31. Y. Liu, Y. L. Xue, and C. Yu, “Modulation instability induced by cross-phase modulation in negative index materials with higher-order nonlinearity,” Opt. Commun. 339, 66–73 (2015). [CrossRef]  

32. C. G. L. Tiofack, A. Mohamadou, K. Alim, K. Porsezian, and T. C. Kofane, “Modulational instability in metamaterials with saturable nonlinearity and higher-order dispersion,” J. Mod. Opt. 591(11), 972–979 (2012).

33. A. M. Glass, D. J. Di Giovanni, T. A. Strasser, A. J. Stentz, R. E. Slusher, A. E. White, A. R. Kortan, and B. J. Eggleton, “Advances in fiber optics,” Bell Labs Tech. J. 5(1), 168–187 (2000). [CrossRef]  

34. B. Daino, G. Gregori, and S. Wabnitz, “New all-optical devices based on third-order nonlinearity of birefringent fibers,” Opt. Lett. 11(1), 42–44 (1986). [CrossRef]   [PubMed]  

35. G.-D. Peng and A. Ankiewicz, “All-optical fibre devices using polarization ellipse rotation,” Opt. Quantum Electron. 22(4), 343–350 (1990). [CrossRef]  

36. L. Lefortand and A. Barthelemy, “All-optical transistor action by polarisation rotation during type-II phase-matched second harmonic generation,” Electron. Lett. 31(11), 910–911 (1995). [CrossRef]  

37. J. Y. Lee, L. Yin, G. P. Agrawal, and P. M. Fauchet, “Ultrafast optical switching based on nonlinear polarization rotation in silicon waveguides,” Opt. Express 18(11), 11514–11523 (2010). [CrossRef]   [PubMed]  

38. R. Kashyap and N. Finlayson, “Nonlinear polarization coupling and instabilities in single-mode liquid-cored optical fibers,” Opt. Lett. 17(6), 405–407 (1992). [CrossRef]   [PubMed]  

39. R. A. Ganeev, M. Suzuki, M. Baba, M. Ichihara, and H. Kuroda, “High-order harmonic generation in Ag nanoparticle-containing plasma,” J. Phys. At. Mol. Opt. Phys. 41(4), 045603 (2008). [CrossRef]  

40. E. L. Falcão-Filho, R. Barbosa-Silva, R. G. Sobral-Filho, A. M. Brito-Silva, A. Galembeck, and C. B. de Araújo, “High-order nonlinearity of silica-gold nanoshells in chloroform at 1560 nm,” Opt. Express 18(21), 21636–21644 (2010). [CrossRef]   [PubMed]  

41. E. Almeida, A. C. L. Moreira, A. M. Brito-Silva, A. Galembeck, C. P. de Melo, L. S. de Menezes, and C. B. de Araújo, “Ultrafast dephasing of localized surface plasmons in colloidal silver nanoparticles: the influence of stabilizing agents,” Appl. Phys. B 108(1), 9–16 (2012).

42. A. S. Reyna and C. B. de Araújo, “Nonlinearity management of photonic composites and observation of spatial-modulation instability due to quintic nonlinearity,” Phys. Rev. A 89(6), 063803 (2014). [CrossRef]  

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

44. A. S. Reyna and C. B. de Araújo, “An optimization procedure for the design of all-optical switches based on metal-dielectric nanocomposites,” Opt. Express 23(6), 7659–7666 (2015). [CrossRef]   [PubMed]  

45. A. M. Brito-Silva, L. A. Gómez, C. B. de Araújo, and A. Galembeck, “Laser ablated silver nanoparticles with nearly the same size in different carrier media,” J. Nanomater. 2010, 142897 (2010). [CrossRef]  

46. V. Besse, G. Boudebs, and H. Leblond, “Determination of the third- and fifth-order optical nonlinearities: the general case,” Appl. Phys. B 116(4), 911–917 (2014). [CrossRef]  

47. L. A. Gómez, C. B. de Araújo, A. M. Brito-Silva, and A. Galembeck, “Solvent effects on the linear and nonlinear optical response of silver nanoparticles,” Appl. Phys. B 92(1), 61–66 (2008). [CrossRef]  

48. L. A. Gómez, C. B. de Araújo, A. M. Brito-Silva, and A. Galembeck, “Influence of stabilizing agents on the nonlinear susceptibility of silver nanoparticles,” J. Opt. Soc. Am. B 24(9), 2136–2140 (2007). [CrossRef]  

49. Y. M. Wu, L. Gao, and Z. Y. Li, “The influence of particle shape on nonlinear optical properties of metal–dielectric composites,” Phys. Status Solidi, B Basic Res. 220(2), 997–1008 (2000). [CrossRef]  

50. R. Sato, M. Ohnuma, K. Oyoshi, and Y. Takeda, “Experimental investigation of nonlinear optical properties of Ag nanoparticles: Effects of size quantization,” Phys. Rev. B 90(12), 125417 (2014). [CrossRef]  

51. J. E. Sipe and R. W. Boyd, “Nonlinear susceptibility of composite optical materials in the Maxwell Garnett model,” Phys. Rev. A 46(3), 1614–1629 (1992). [CrossRef]   [PubMed]  

52. Y. Xu, X. Chen, and Y. Zhu, “Modeling of micro-diameter-scale liquid core optical fiber filled with various liquids,” Opt. Express 16(12), 9205–9212 (2008). [CrossRef]   [PubMed]  

53. J. M. Ziman, Principles of the Theory of Solids (Cambridge University, 1969).

54. P. B. Johnson and R. W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6(12), 4370–4379 (1972). [CrossRef]  

55. M. N. Z. Abou’ou, P. T. Dinda, C. M. Ngabireng, B. Kibler, and F. Smektala, “Impact of the material absorption on the modulational instability spectra of wave propagation in high-index glass fibers,” J. Opt. Soc. Am. B 28(6), 1518–1528 (2011). [CrossRef]  

56. S. Wang and L. Zhang, “An efficient split-step compact finite difference method for cubic–quintic complex Ginzburg–Landau equations,” Comput. Phys. Commun. 184(6), 1511–1521 (2013). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 (a) The experimental setup: polarizer (P), beam splitter (BS), spherical lenses with f = 5 cm (L), 40x microscope objective (L1), 20x microscope objective (L2), polarizing beam splitter cube (PBS) and reference detector (RD). The transmitted light with vertical and horizontal polarization was captured in the fast detectors D1 and D2, respectively. (b) Inner diameter of capillary, in a portion of 5 mm, showing small asymmetries. The inset is an optical microscope image of a small section of the hollow capillary core (length: 1 mm).
Fig. 2
Fig. 2 Local gain spectra of modulation instability versus the frequency shift along the fast axis, for the five samples.
Fig. 3
Fig. 3 Numerical pulse shape evolution for sample A (S-A), sample B (S-B), sample C (S-C), sample D (S-D) and sample E (S-E), with input intensity of 60 MW/cm2. Pulse duration: 80 ps. Propagation length: 9 cm.
Fig. 4
Fig. 4 Normalized transmittance as a function of the incident polarization azimuth angle, θ, for (a) sample A, (b) sample B, (c) sample C, (d) sample D and (e) sample E. From top to bottom, the incident peak intensities are 6, 24, 42 and 60 MW/cm2, in each row. Black circles and red squares correspond to vertical and horizontal polarization transmittance, respectively. Solid lines were obtained from numerical solutions of Eq. (24) and the parameters of Table 1.
Fig. 5
Fig. 5 Vertical (black circles) and horizontal (red squares) polarization transmittance as a function of the incident peak intensities variation, for (a) sample A, (b) sample B, (c) sample C, (d) sample D and (e) sample E. Black and red solid lines were obtained from numerical solutions of Eq. (24) and the parameters of Table 1. Blue dashed lines represent the numerical solutions of Eq. (24) neglecting the contribution of Re( χ ( 5 ) ).

Tables (1)

Tables Icon

Table 1 Linear absorption coefficienta, second-order dispersion coefficienta, third and fifth-order susceptibilitiesb for pure CS2 (S-A) and silver-colloids with different nanoparticles volume fraction.

Equations (37)

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

E( r,t )= 1 2 ( x ^ E x + y ^ E y )exp( i ω 0 t )+c.c,
2 E( r,ω )+ k 0 2 ε( ω )E( r,ω )=0,
2 E μ ( r,ω )+ k 0 2 ε μσ ( ω ) E σ ( r,ω )=0,
P μ NL ( ω )= ε 0 ε μ NL ( ω ) E μ ( ω ),
P μ (3) ( ω )=3 ε 0 { 2 χ xxyy (3) ( ω ) E μ ( ω )[ E( ω ) E * ( ω ) ]+ χ xyyx (3) ( ω ) E μ * ( ω )[ E( ω )E( ω ) ] },
P μ (5) ( ω )=10 ε 0 { 10 3 χ xxyyxx (5) ( ω ) | E μ ( ω ) | 2 [ E( ω )E( ω ) ] E μ * ( ω ) + 5 3 χ xxyyyy (5) ( ω ) σ=x,y | E σ ( ω ) | 4 E μ ( ω ) },
ε μ NL ( ω )=3[ ( 2 χ xxyy (3) ( ω )+ χ xyyx (3) ( ω ) ) | E μ ( ω ) | 2 +2 χ xxyy (3) ( ω ) σ=x,y | E σ ( ω ) | 2 ( 1 δ μ,σ ) ] +10[ 10 3 χ xxyyxx (5) ( ω ) | E μ ( ω ) | 4 + 5 3 χ xxyyyy (5) ( ω )( | E μ ( ω ) | 4 + σ=x,y | E σ ( ω ) | 4 ( 1 δ μ,σ ) ) ] +[ 3 χ xyyx (3) ( ω )+10 10 3 χ xxyyxx (5) ( ω ) | E μ ( ω ) | 2 ] σ=x,y [ E σ ( ω ) ] 2 ( 1 δ μ,σ ) E μ * ( ω ) E μ ( ω ) .
E μ ( r,ω )=F( x,y ) A μ ( z,ω ω 0 )exp( i β 0,μ z ),
2i β 0,μ A μ z +( β μ 2 β 0,μ 2 ) A μ =0,
2 F x 2 + 2 F y 2 +[ ε μ ( ω ) k 0 2 β μ 2 ]F=0,
ε μ = ( n 0,μ +Δ N μ ) 2 ( n 0,μ ) 2 +2( n 0,μ )( Δ N μ ),
( Δ n μ +i Δ α μ 2 k 0 )= 3 2 n 0,μ | F | 2 [ ( 2 χ xxyy (3) + χ xyyx (3) ) | A μ | 2 +2 χ xxyy (3) σ=x,y | A σ | 2 ( 1 δ μ,σ ) ] + 10 2 n 0,μ | F | 4 [ 10 3 χ xxyyxx (5) | A μ | 4 + 5 3 χ xxyyyy (5) ( | A μ | 4 + σ=x,y | A σ | 4 ( 1 δ μ,σ ) ) ] + 1 2 n 0,μ | F | 2 [ 3 χ xyyx (3) +10 10 3 χ xxyyxx (5) | F | 2 | A μ | 2 ] σ=x,y [ A σ ( ω ) ] 2 ( 1 δ μ,σ ) A μ * A μ exp[ 2i( β 0,σ β 0,μ )z ].
F= F (0) +ξ F (1) +,
β μ 2 = ( β μ ) 2 +ξ2 β μ ( Δ β μ )+,
ξ 0 : [ ( 2 x 2 + 2 y 2 + n 0,μ 2 k 0 2 ) β μ 2 ] F (0) =0,
ξ 1 : [ ( 2 x 2 + 2 y 2 + n 0,μ 2 k 0 2 ) β μ 2 ] F (1) +[ 2( n 0,μ )( Δ N μ ) k 0 2 2 β μ Δ β μ ] F (0) =0.
F core ( 0 ) ( ρ )= C 1 J m ( ρ n 0,μ 2 k 0 2 β μ 2 ),
F cladding ( 0 ) ( ρ )= C 2 K m ( ρ β μ 2 n cladding 2 k 0 2 ).
Δ β μ = k 0 ( Δ N μ ) | F ( 0 ) | 2 dxdy | F ( 0 ) | 2 dxdy ,
A μ z =i( β μ +Δ β μ β 0,μ ) A μ ,
β μ ( ω )= β 0,μ + β μ (1) ( ω ω 0 )+ 1 2 β μ (2) ( ω ω 0 ) 2 +O[ ( ω ω 0 ) 3 ],
A μ z + β μ (1) A μ t +i β (2) 2 2 A μ t 2 + α 0 2 A μ =i k 0 2 n 0,μ 3 F (1) { [ ( 2 χ xxyy (3) + χ xyyx (3) ) | A μ | 2 +2 χ xxyy (3) σ=x,y ( 1 δ μ,σ ) | A σ | 2 ] A μ + χ xyyx (3) [ σ=x,y ( 1 δ μ,σ ) ( A σ ) 2 ] A μ * exp[ 2i( β 0,σ β 0,μ )z ] } +i k 0 2 n 0,μ 10 F (2) { [ 10 3 χ xxyyxx (5) | A μ | 4 + 5 3 χ xxyyyy (5) [ | A μ | 4 + σ=x,y ( 1 δ μ,σ ) | A σ | 4 ] ] A μ + 10 3 χ xxyyxx (5) [ σ=x,y ( 1 δ μ,σ ) | A μ | 2 ( A σ ) 2 ] A μ * exp[ 2i( β 0,σ β 0,μ )z ] },
A ± z + 1 2 [ β + (1) A ± t + β (1) A t ]+ i 2 β (2) 2 A ± t 2 + α 0 2 A ± = i 2 ( Δ β 0 ) A +i 3 ω 0 n 0 c F ( 1 ) [ χ xxyy (3) ( | A + | 2 + | A | 2 )+ χ xyyx (3) | A | 2 ] A ± +i 5 ω 0 2 n 0 c F ( 2 ) { 10 3 χ xxyyxx (5) [ | A + + A | 2 ( A + + A ) * | A + A | 2 ( A + A ) * ] A + 5 6 χ xxyyyy (5) [ | A + + A | 4 + | A + A | 4 ] } A ± ,
A ± z + i 2 β (2) 2 A ± τ 2 + α 0 2 A ± = i 2 ( Δ β 0 ) A +i ω 0 n 0 c F ( 1 ) χ xxxx (3) [ ( | A + | 2 + | A | 2 )+ | A | 2 ] A ± +i 5 ω 0 12 n 0 c F ( 2 ) χ xxxxxx (5) { 4[ | A + + A | 2 ( A + + A ) * | A + A | 2 ( A + A ) * ] A +[ | A + + A | 4 + | A + A | 4 ] } A ± ,
ε eff ( λ,f )= ε h ( λ )[ 1+ 3Θ( λ )f 1Θ( λ )f ],
ε h ( λ )= [ n C S 2 ( λ ) ] 2 = [ 1.580826+ 1.52389× 10 2 λ 2 + 4.8578× 10 4 λ 4 + 8.2863× 10 5 λ 6 + 1.4619× 10 5 λ 8 ] 2 .
ε NP ( λ )=( 1 λ 2 λ p 2 )+i( 1 2πc τ r λ 3 λ p 2 ),
β (1) ( λ,f )= 1 c [ n eff ( λ,f )λ d[ n eff ( λ,f ) ] dλ ],
β (2) ( λ,f )= λ 3 2π c 2 d 2 [ n eff ( λ,f ) ] d λ 2 .
A ± ( z,τ )=±iB( z,τ )exp( α 0 z 2 i Δ β 0 2 z )exp( τ 2 τ 0 2 ),
Λ( τ )= B 0 2 ω 0 c [ 3 F ( 1 ) Im( χ ( 3 ) )+20 B 0 2 F ( 2 ) Im( χ ( 5 ) )exp( 2 τ 2 τ 0 2 ) ]exp( 2 τ 2 τ 0 2 ),
Φ( τ )= 1 2 β ( 2 ) [ 2 τ 0 2 ( 1 2 τ 2 τ 0 2 ) ] + B 0 2 ω 0 c [ 3 F ( 1 ) Re( χ ( 3 ) )+20 B 0 2 F ( 2 ) Re( χ ( 5 ) )exp( 2 τ 2 τ 0 2 ) ]exp( 2 τ 2 τ 0 2 ).
a ± ( z,τ )=±i{ B 0 exp[ Λ( τ )z ]+ B 1,± ( z,τ ) }exp( α 0 z 2 i Δ β 0 2 z )exp[ iΦ( τ )z ]exp( τ 2 τ 0 2 ),
i{ B 1,± z +iΦ B 1,± i 2 ( Δ β 0 )( B 1,+ B 1, ) } 1 2 β (2) { 2 B 1,± τ 2 +2[ iz Φ τ 2τ τ 0 2 ] B 1,± τ +iz B 1,± [ 2 Φ τ 2 4τ τ 0 2 Φ τ +iz ( Φ τ ) 2 ]+ 2 τ 0 2 [ 2 τ 2 τ 0 2 1 ] B 1,± } = ω 0 c | B 0 | 2 F ( 1 ) χ xxxx (3) { 5 B 1,± +2[ B 1,+ + B 1, ] }exp( 2Λz α 0 z2 τ 2 τ 0 2 ) 5 ω 0 12c F ( 2 ) χ xxxxxx (5) { 16 | B 0 | 4 [ 3 B 1,± +2( B 1,+ + B 1, + B 1, ) ] +48 | B 0 | 2 B 0 2 [ B 1,+ + B 1, ] * }exp( 4Λz2 α 0 z4 τ 2 τ 0 2 ).
K 2 +[ 2 β (2) Ω( z Φ τ ) ]KM=0,
M=Re( N 2 )2 β (2) Ω[ z Φ τ Re( N ) 2τ τ 0 2 Im( N ) ] + [ 40 ω 0 c B 0 2 | B 0 | 2 F ( 2 ) exp( 4Λz2 α 0 z4 τ 2 τ 0 2 ) ] 2 Re[ ( χ xxxxxx 5 ) 2 ],
N=Φ+ 1 2 β (2) { Ω[ Ω+2i( iz Φ τ 2τ τ 0 2 ) ]+iz[ 2 Φ τ 2 4τ τ 0 2 Φ τ +iz ( Φ τ ) 2 ]+ 2 τ 0 2 [ 2 τ 2 τ 0 2 1 ] } 3 ω 0 c | B 0 | 2 exp( 2Λz α 0 z2 τ 2 τ 0 2 )[ 3 F ( 1 ) χ xxxx (3) +20 | B 0 | 2 F ( 2 ) χ xxxxxx (5) exp( 2Λz α 0 z2 τ 2 τ 0 2 ) ].
Select as filters


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