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

Theory of nonlinear pulse propagation in silicon-nanocrystal waveguides

Open Access Open Access

Abstract

We develop a comprehensive theory of the nonlinear propagation of optical pulses through silica waveguides doped with highly nonlinear silicon nanocrystals. Our theory describes the dynamics of arbitrarily polarized pump and Stokes fields by a system of four generalized nonlinear Schrödinger equations for the slowly varying field amplitudes, coupled to the rate equation for the number density of free carriers. In deriving these equations, we use an analytic expression for the third-order effective susceptibility of the waveguide with randomly oriented nanocrystals, which takes into account both the weakening of the nonlinear optical response of silicon nanocrystals due to their embedment in fused silica and the change in the tensor properties of the response due to the modification of light interaction with electrons and phonons inside the silicon-nanocrystal waveguide. In order to facilitate the use of our theory by experimentalists, and for reasons of methodology, we provide a great deal of detail on the mathematical treatment throughout the paper, even though the derivation of the coupled-amplitude equations is quite straightforward. The developed theory can be applied for the solving of a wide variety of specific problems that require modeling of nonlinear optical phenomena in silicon-nanocrystal waveguides.

© 2013 Optical Society of America

1. Introduction

In a lapse of two decades, the concept of silicon photonics proposed by Richard Soref [1] in the mid 1980s has culminated in the realization of high-performance passive and active microscale optical devices [27]. One of the main advantages offered by such devices is their low fabrication cost and the possibility to integrate them with standard microelectronics components using the mature complementary semiconductor–metal–oxide (CMOS) technology [8]. Another merit is the tight optical confinement, which significantly enhances the efficiency of all nonlinear optical processes, and the small linear absorption in silicon at infrared wavelengths. Being optically dense and strongly nonlinear material, silicon is ideal for the integration of the five key elements enabling optical information processing: waveguides [911], modulators [1215], amplifiers [1622], lasers [2326], and detectors [27]. The functionalities of the nonlinear silicon elements are made possible by the ultrafast Kerr effect and stimulated Raman scattering—stemming from the third-order optical nonlinearity—and the fifth-order free-carrier and thermo-optic effects [2835].

The optical functionalities realized with bulk silicon can be also achieved with low-dimensional silicon in the form of spherical nanocrystals (NCs) embedded in a SiO2 matrix [3640]. Indeed, since the refractive index of the Si-NCs/SiO2 composite may be anywhere between 1.45 and 2.2 depending on the excess of silicon, this composite can confine and guide light better than the pure silica fibers do [41, 42]. More importantly, the nonlinear effects associated with the third-order electron and phonon nonlinearities are significantly enhanced in Si NCs as compared to bulk silicon. Recent measurements have shown that the nonlinear Kerr coefficient and Raman gain of the Si-NCs/SiO2 composite may exceed those of silicon by factors of 100 and 10000 [37, 43, 44], respectively, even when the volume fraction of the NCs does not exceed 10%. Because of an indirect bandgap, neither bulk nor low-dimensional silicon can naturally emit light in the telecommunication band. However, by simply coupling Si NCs to Er3+ ions doping the SiO2 matrix, one can realize the desired emission from the NCs and thus create from them a promising material for light emitting devices [39, 45, 46]. All these features allow us to expect that the Si-NCs/SiO2 composite will eventually replace the silicon-on-insulator (SOI) technology in creating the key elements of silicon photonics.

In contrast to a large number of experimental studies of the Si-NCs/SiO2 composite over the past years [12, 37, 38, 43, 4750], there have been only a few theoretical works centered at the nonlinear propagation through this material [36, 41, 42, 51, 52]. The imbalance between theory and experiment is especially striking in light of the advanced numerical and analytical models of the nonlinear light–matter interaction developed for silicon [3, 4, 53], and the numerous theoretical studies of the nonlinear optical phenomena in SOI waveguides [5462]. The reason behind this disproportion is the absence of the unified theoretical platform for modeling nonlinear optical propagation through the Si-NCs/SiO2 composite. The demand for such a platform is even higher than it was for silicon, as it is extremely computationally challenging to solve Maxwell’s equations for a macroscopic sample of the composite with hundreds of thousands of tiny NCs treated individually.

In this paper, we present the first theory of nonlinear pulse propagation through a Si-NC-doped silica waveguide. Our theory assumes arbitrary orientation of the NCs in space and describes this situation using an effective third-order susceptibility of the Si-NCs/SiO2 composite [36]. The paper is organized as follows. In Section 2, we derive a system of coupled nonlinear equations for the slowly varying amplitudes of the pump and Stokes pulses inside a nonlinear waveguide. The system shows that the source of the space and time evolution of the amplitudes is the waveguide’s nonlinear material polarization. Section 3 is therefore devoted to the calculation of the electronic, Raman, and free-carrier contributions to the nonlinear polarization induced by the pump and Stokes fields inside the Si-NCs/SiO2 composite. The free-carrier contribution depends on the generation rate of electron–hole pairs inside the NCs due to the effect of two-photon absorption, whose efficiency is governed by field intensity and thus depends on the effective mode area (EMA) of the waveguide. The ambiguity in defining EMA is emphasized in Section 4, where we provide two alternative definitions of EMA and explain the often ignored difference between them. The results of Sections 2–4 are merged in Section 5 to give the desired coupled amplitude equations governing the nonlinear propagation of the two pulses. These equations are then simplified for the case of continuous waves, before Section 6 summarizes our results and concludes the paper.

2. Nonlinear propagation equations

The general equations governing propagation of optical field through a nonlinear waveguide may be derived by treating the nonlinear material polarization as a small perturbation coupling otherwise independent propagating modes of the same waveguide operating in the linear regime. Let us consider the propagation of narrowband pump (μ = p) and Stokes (μ = s) pulses, whose spectra are centered around frequencies ωp and ωs. In this case, the unperturbed propagating modes of the linear waveguide are the solutions to Maxwell’s equations for the pump and Stokes fields

×E˜μ(0)(r,ω)=iωμ0H˜μ(0)(r,ω)
and
×H˜μ(0)(r,ω)=iωε0εL(r,ω)E˜μ(0)(r,ω),
where εL(r, ω) is the transverse profile of linear permittivity and r is the two-dimensional radius vector perpendicular to the waveguide axis z.

The solution to Maxwell’s equations for the perturbed fields can be expanded over the complete set of propagating modes eμν(r, ωμ)eμνz and hμν(r, ωμ)eμνz satisfying Eq. (1) as

E˜μ(r,ω)=νa˜μν(z,ωωμ)eμν(r,ωμ)Nμνeiβμνz
and
H˜μ(r,ω)=νa˜μν(z,ωωμ)hμν(r,ωμ)Nμνeiβμνz,
where βμνβν(ωμ) is the real propagation constant of mode ν evaluated at the carrier frequency ωμ of field μ and we assume the modal profiles eμν and hμν to be normalized according to the condition
(eμν*×hμν+c.c.)dr=δνν(eμν*×hμν+c.c.)dr=4Nμν,
where the integration is taken over the entire area of the xy plane and the constants Nμν are implicitly defined. With this normalization, one finds that the total power carried by field μ is given by
Pμ=12Re(E˜μ×H˜μ*)dr=ν|a˜μν|2.
As in Eq. (3), the double integral here is evaluated over the entire transverse plane.

If P˜μNL(r,ω) is the sum of material polarizations stemming from all types of nonlinearities in the waveguide, then the perturbed fields μ and μ must obey the equations

×E˜μ(r,ω)=iωμ0H˜μ(r,ω)
and
×H˜μ(r,ω)=iωε0εL(r,ω)E˜μ(r,ω)iωP˜μNL(r,ω).
The perturbed and unperturbed fields are also interrelated through the reciprocity theorem [3, 59, 63]
z(E˜μ(0)×H˜μ*+E˜μ*×H˜μ(0))dr=(E˜μ(0)×H˜μ*+E˜μ*×H˜μ(0))dr.

Using the vector identity ∇ · (f × g) = g · (∇ × f) − f · (∇ × g), together with Eqs. (1) and (5), it is possible to express the gradient in Eq. (6) by means of the scalar product

(E˜μ(0)×H˜μ*+E˜μ*×H˜μ(0))=iω(E˜μ(0)P˜μNL*).
By introducing this expression into Eq. (6), substituting in the resulting relation both the unperturbed solutions
E˜μ(0)=eμν(r,ωμ)Nμνeiβν(ω)zandH˜μ(0)=hμν(r,ωμ)Nμνeiβν(ω)z
and the perturbed ones given in Eq. (2), we find using Eq. (3) that
z{a˜μν(z,ωωμ)ei[βν(ωμ)βν(ω)]z}=iω4Nμνeμν*(r,ωμ)P˜μNL(r,ω)eiβν(ω)zdr.
The evaluation of the derivative in this equation gives
(z+in=11n!βνω|ωμ(ωωμ)n)a˜μν(z,ωωμ)=iω4Nμνeμν*(r,ωμ)P˜μNL(r,ω)eiβν(ωμ)zdr.

By multiplying both sides of Eq. (10) by ei(ωωμ)t and integrating with respect to ω, we arrive at the following time-domain equation:

(z+n=1in+1n!βνω|ωμntn)aμν(z,t)=eiβμνz4Nμνeμν*(r,ωμ)PμNL(r,t)teiωμtdr,
in which we have set
aμν(z,t)=12π+a˜μν(z,ω)eiωtdωandPμNL(r,t)=12π+P˜μNL(r,ω)eiωtdω.

If we now represent the polarization vector as a product of the slowly varying amplitude and the rapidly oscillating exponential, PμNL(r,t)=PωμNL(r,t)eiωμt, then Eq. (11) acquires the form

aμνz+n=1in+1βμν(n)n!naμνtn=iωμ4Nμν(1+iωμt)eiβμνz(eμν*PωμNL)dr,
where βμν(n)=βν/ω|ω=ωμ is the nth-order dispersion parameter at the frequency ωμ. The term with the time derivative on the right-hand side of this equation accounts for the effect of self steepening, which is significant for ultrashort optical pulses [64].

Equation (13) shows that the variation of the amplitudes aμν of pump and Stokes fields is due to the nonlinear material polarization of Si-NCs/SiO2 waveguide. This equation is similar to those derived in Refs. [59, 63, 65].

In this paper, we shall calculate the nonlinear polarization of the Si-NCs/SiO2 waveguide under the frame of the effective medium theory [47, 66, 67]. According to this theory, the effective permittivity of the Si-NCs/SiO2 composite is related to the permittivity ε1 of silicon and the permittivity ε2 of silica as [68]εeff = (1/4) [u + (u2 + 8ε1ε2)1/2], where u = (3f − 1)ε1 + (2 − 3f)ε2 and f is the volume fraction of Si NCs in the composite. The dispersion of εeff needs to be taken into account when one calculates the parameters βμν(n), but may generally be ignored in the cases where εeff explicitly appears in the coupled amplitude equations. Specifically, for narrowband pump and Stokes pulses in the 1.55-μm spectral region, the effective permittivity of the Si-NCs/SiO2 composite may be calculated by taking ε1 ≈ 12.1 and ε2 ≈ 2.1. An allowance for the size dependency of the NCs’ permittivity can be made using ε1 extracted from the spectroscopic ellipsometry data presented in Ref. [48].

3. Nonlinear polarization of Si-NCs/SiO2 waveguide

The nonlinear material polarization of the Si-NCs/SiO2 waveguide may originate from the non-linearities of both silicon crystallites and silica matrix. Since the nonlinear effects in Si NCs are typically much stronger than those in SiO2, one may safely neglect the third-order susceptibility of silica, provided the volume fraction of the NCs is larger than 0.1% [36]. With this simplification, the total material polarization is a sum of three terms [3, 4, 53, 69]

PωμNL(r,t)=PωμK(r,t)+PωμR(r,t)+PωμFC(r,t),
which represent the electronic (Kerr), Raman, and free-carrier contributions.

The electronic polarization stems from the nonlinear interaction of an optical field with electronic clouds of silicon atoms and leads to an almost instantaneous (with a response time less than 100 fs) change in the refractive index of Si NCs (Kerr effect), as well as to the two-photon absorption (TPA) [12]. The Raman polarization allows for the interaction of light with optical phonons and is responsible for the stimulated Raman scattering (SRS) and lasing [1719, 26]. Free electrons and holes generated inside the NCs via TPA also change the refractive index of the NCs, which is usually referred to as the effect of free-carrier dispersion (FCD), and result in the free-carrier absorption (FCA) of the pump and Stokes fields [19, 29]. The effects of FCD and FCA are included in the last polarization term in Eq. (14).

Consider the three types of polarization contributions separately, while assuming that the mean NC size is large enough for the quantum confinement to not increase the bandgap of silicon significantly.

3.1. Electronic contribution

The Kerr polarization induced in the Si-NCs/SiO2 waveguide with randomly oriented nanocrystals is given by the tensor product

PωμK(r,t)=ε0χK(3)(ωμ;ωμ,ωμ,ωμ)Eωμ(r,t)Eωμ*(r,t)Eωμ(r,t),
where Eωμ(r, t) is the slowly varying amplitude of the electric field, which may be calculated by Fourier-transforming Eq. (2), and the third-order susceptibility tensor is of the form [36]
χK(3)(ωμ;ωμ,ωμ,ωμ)=χμ(8+7ρ45(δklδmn+δkmδln+δknδlm)+1ρ9δklδlmδmn),
where {k, l, m} = {x, y, z}, δij is the Kronecker delta, and ρ is the nonlinear anisotropy factor (ρ ≈ 1.27 near the 1.55 μm wavelength). Owing to a uniform distribution of Si-NC orientations in space, the Kerr tensor of the Si-NCs/SiO2 composite is no longer described with respect to the crystallographic basis, but rather is given in the reference frame associated with the nonlinear waveguide.

The strength of the electronic nonlinearity is determined by the complex parameter [4, 56, 69]

χμ=cε0εeff[n2+iβTPA/(2kμ)]ξ,
where n2 and βTPA are the nonlinear Kerr parameter and TPA coefficient of a Si NC, kμ = ωμ/c, and
ξ=1f(εeffε1)2=[(3f1)εeff+ε2]2f(u2+8ε1ε2)
is the third-order susceptibility attenuation factor [36].

Both the Kerr parameter and the TPA coefficient at 1.55 μm are enhanced in Si NCs when compared to bulk silicon. The Kerr parameter may be enhanced in the Si-NCs/SiO2 composite by as much as two orders of magnitude and lie in the range (2±1)×10−12 cm2/W [37, 38, 44], while the reported spread of the TPA coefficient is larger and ranges 5 to 170 cm/GW [37, 42, 44, 70]. The values of βTPA and n2 depend on the doping of Si NCs [44] and may vary with powers and repetition rates of optical pulses, due to the band-filling-induced saturation of the nonlinear absorption [71]. In borrowing these parameters from the experimental papers, it is important to distinguish between their values related to a single Si NC and the entire composite. For example, Spano et al.[71] found the intensity-dependent TPA coefficient of the Si-NCs/SiO2 composite with f = 8% to be given by the expression βTPA(c)=βTPA(0)/[1+(I/Is)2], where βTPA(0)=(7.0±0.6)cm/GW and Is = 4070 GW/cm2. The TPA coefficient of a single NC in this case may be estimated as βTPA=βTPA(c)/ξ.

The electronic contribution to the nonlinear material polarization can be written entirely in vector form as

PωμK=ε0χμ[8+7ρ45(2|Eωμ|2Eωμ+Eωμ2Eωμ*)+1ρ9ηEη2Eη*η^],
where Eη is the ηth component of the vector Eωμ and η̂ is the unit vector in the direction of the ηth Cartesian axis.

In the rest of the paper, we shall restrict ourselves to the special case of only two unperturbed transverse modes (eμx and eμy) in the expansion of Eq. (2), assuming them to be polarized along the x and y axes. In this case, the orthogonality relations (eμνeμν)=eμν2δνν and (eμνeμν*)=|eμν|2δνν, the expansion

Eωμ=ν=x,yaμνeμνNμνeiβμνz,
and Eq. (19) enable one to obtain
eiβμνzNμν(eμν*PωμK)dr=ε0χμ[8+7ρ45(2aμννΓνν(μ)|aμν|2+aμν*νaμν2Λνν(μ)e2i(βμνβμν)z)+1ρ9Γνν(μ)aμν|aμν|2],
where
Γνν(μ)=1NμνNμν|eμν|2|eμν|2dr
and
Λνν(μ)=1NμνNμνeμν*2eμν2dr.

It is easy to prove using Maxwell’s equations that (hμν) = (βμν/μ0ωμ)ẑ × eμν for the transverse electric modes, so that the normalization condition yields

Nμν=βμν2μ0ωμ|eμν|2dr.

3.2. Raman contribution

The Raman polarization of the Si-NCs/SiO2 waveguide with randomly oriented nanocrystals can be represented as

PωμR(r,t)=eiωμttdt1tdt2tdt3ε0χR(3)(tt1,tt2,tt3)×(Eωμ(r,t1)Eωμ*(r,t2)Eωμ(r,t3)ei(ωμt1ωμt2+ωμt3)+Eωμ(r,t1)Eωμ*(r,t2)Eωμ(r,t3)ei(ωμt1ωμt2+ωμt3)),
where {μ, μ′} = {p, s}, μμ′, and the effective susceptibility tensor is given by the expressions [3, 4, 36, 69]
χR(3)(t1,t2,t3)=12[δ(t1t2)δ(t3)klmn+δ(t1)δ(t2t3)knml]ξH(t2),
klmn=2945(δkmδln+δknδlm)1645δklδmn29δklδlmδmn,
and
H(t)=2χRΓRΩR(ΩR2ΓR2)1/2et/τ2sin(t/τ1),
where χR is the peak value of the Raman susceptibility of an individual Si NC, 2ΓR = 2/τ2 is the amplification bandwidth, and 1/τ1=(ΩR2ΓR2)1/2ΩR is the Stokes shift.

Recent experiments on SRS in the Si-NCs/SiO2 composite have shown [43] that the Raman gain of the composite may be 10000 times larger than that of bulk silicon (∼ 10−17 m2/V2) [3]. Since the third-order susceptibility of Si NCs is attenuated by the factor ξ in the composite, this gives us an estimate ξχR ∼ 10−13 m2/V2. The peak position and bandwidth of gain spectrum in Si NCs also change significantly with respect to their values 2ΓR ≈ 105 GHz and ΩR = 15.6 THz in bulk silicon. In particular, for spherical NCs of 2 nm in diameter, the gain peak broadens up to 2 THz and red shifts by about 0.6 THz [72, 73]. The downshift and broadening of the Raman spectrum with the change in size of Si NCs can be calculated numerically using the phenomenological theory of Richter et al.[74], developed by Campbell and Fauchet [73, 75].

The origin of giant Raman gain in Si NCs is currently unclear and warrants further investigation. One possible reason behind this effect may be the resonant field enhancement inside the nanocrystals caused by the interaction of light with optical phonons and the formation of the localized phonon-polariton modes [76, 77]. The enhancement occurs within the reststrahlen band of phonon-polariton dispersion and is analogous to the plasmon-induced field enhancement in the vicinity of metallic nanoparticles.

By introducing Eq. (25a) into Eq. (24) and evaluating the integrals, we find that

PωμR(r,t)=ε0ξtdt1H(tt1)klmnEωμ(r,t1)Eωμ*(r,t1)Eωμ(r,t)ei(ωμωμ)(tt1).
In deriving this expression, we have omitted the nonresonant terms in Eq. (24) and taken into account that the remaining two terms are equal. Now using Eqs. (20) and (25b), we get
eiβμνzε0ξNμν(eμν*PωμR)dr=2945νΛννμμexp(iβμνμνμνμνz)aμν(t)taμν*(t1)aμν(t1)H(tt1)eiωμμ(tt1)dt1+2945νΨννμμexp(iβμνμνμνμνz)aμν(t)taμν*(t1)aμν(t1)H(tt1)eiωμμ(tt1)dt11645νΓννμμaμν(t)taμν*(t1)aμν(t1)H(tt1)eiωμμ(tt1)dt129Λννμμaμνttaμν*(t1)aμν(t1)H(tt1)eiωμμ(tt1)dt1,
where μμ′, βμνμνμνμν=βμν+βμνβμνβμν, ωμμ = ωμωμ,
Λννμμ=(eμν*eμν*)(eμνeμν)NμνNμνNμνNμνdr,
Ψννμμ=(eμν*eμν)(eμνeμν*)NμνNμνNμνNμνdr,
and
Γννμμ=1NμνNμν|eμν|2|eμν|2dr.

3.3. Free-carrier effects

The last contribution to the nonlinear polarization, stemming from free-carrier effects, may be written as follows:

PωμFC(r,t)=2ζε0neff[ΔnFC+ic/(2ωμ)ΔαFC]Eωμ(r,t),
where
ζ=neffn1=(ε1εeff)1/2(3f1)εeff+ε2u2+8ε1ε2
is the linear susceptibility attenuation factor, neff=εeff, and n1=ε1. The amounts of changes to the refractive index and absorption coefficient of Si NCs are related to the number density N of the TPA-generated electron–hole pairs as [4, 56]
ΔnFC=σn(ω0/ωμ)2(1+ςN0.2)N0.8andΔαFC=σα(ω0/ωμ)2N,
where σn = 1.35 × 10−22 m2.4, ς = 6.53 × 10−6 m0.6, σα = 14.5 × 10−21 m2, and ω0 = 2πc/(1.55 μm).

With the above form of the nonlinear polarization, one finds that

eiβμνzNμν(eμν*PωμFC)dr=4(ζ/c)(neff/nμν)[ΔnFC+ic/(2ωμ)ΔαFC]aμν,
where the modal refractive index is defined as nμν = βμν/kμ.

The number density of electron–hole pairs entering Eq. (31) grows with light intensity and may be calculated from the rate equation

Nt=Nτcμ12h¯ωμAeffPμz,
where ∂Pμ/∂z is the rate of mode μ power dissipation due to the TPA and τc is the effective free-carrier lifetime. The same average effective mode area (EMA), Aeff, for all modes have been assumed in this equation for the sake of simplicity. As will be discussed below, this assumption has to be abandoned if one needs to write the propagation equations in terms of the average field intensities inside the Si-NCs/SiO2 composite.

By assuming continuous pump and Stokes waves and neglecting the time derivatives in Eq. (13), it may be readily shown using Eq. (21) that

Pμz=ν(aμνaμν*z+aμν*aμνz)=14ξc2ε02εeffβTPAν(8+7ρ45ν2Γνν(μ)|aμν|2|aμν|2+13+2ρ45Γνν(μ)|aμν|4).
In evaluating the power dissipation rate, we have taken into account that Λνν(μ)=Γνν(μ) and omitted the spatially varying terms, which arise due to the waveguide birefringence (nμνnμν) and do not contribute to the TPA-induced power dissipation.

4. Effective mode area

It should be recognized that the parameters Γννμμ, Λννμμ, and Ψννμμ allow one to introduce three types of EMAs, dependent on the frequencies and polarizations of both the pump and Stokes fields. For simplicity, we define the EMA of mode ν at ωμ in the common fashion as [4, 10, 78]

Aeffμν=(|eμν|2dr)2/|eμν|4dr.
Using this definition, we adopt the following definition of the average EMA:
Aeff=μ,ν(Aeffμν)1/4.

When using Eqs. (33), (35), and (36), one should keep in mind that the quantities |aμν|2/Aeffμν and |aμν|2/Aeff are not modal intensities. In order to treat the nonlinear propagation in terms of the average mode intensities inside the nonlinear composite, one needs to use a different EMA [51],

𝒜effμν=𝒜NL|eμν|2dr/NL|eμν|2dr,
in which 𝒜NL is the cross section area of the nonlinear core of the waveguide (filled with Si-NCs/SiO2 composite) and the symbols NL and ∞ denote integrations over the nonlinear core and the entire transverse plane, respectively. It is easy to see that μν=|aμν|2/𝒜effμν gives the average intensity of field μ at ων, i.e., the ratio of the power carried by this field through the waveguide core to the core area 𝒜NL.

5. Coupled amplitude equations

By combining Eqs. (13), (14), (21), (27), and (31)(34), we arrive at the set of coupled equations for the slowly varying amplitudes aμν and free-carrier density N. These equations may be simplified by neglecting the effects of self steepening and assuming that no phase matching occurs between the waves of different frequencies and polarizations. In this case, the coupled amplitude equations acquire the form of the generalized nonlinear Schrödinger equation

aμνz+n=1in+1βμν(n)n!naμνtn+αμν2aμν=18ξc2ε02εeff(βTPA2in2kμ)(8+7ρ452Γνν(μ)|aμν|2+29+16ρ45Γνν(μ)|aμν|2)aμν+8iε0ωμ45ξΓννμμaμν(t)taμν*(t1)aμν(t1)H(tt1)eiωμμ(tt1)dt14iε0ωμ45ξΓννμμaμν(t)taμν*(t1)aμν(t1)H(tt1)eiωμμ(tt1)dt1ζneffnμν(ω0ωμ)2(iσnkμ(1+ςN0.2)+σα2N0.2)N0.8aμν,
while the rate equation becomes
Nt=Nτc+ξc2ε02εeff4Aeffμ,νβTPA2h¯ωμ(8+7ρ452Γνν(μ)|aμν|2+29+16ρ45Γνν(μ)|aμν|2)|aμν|2.
Here we have collected the terms with ν = ν′ and explicitly written out the remaining terms with ν′ ≠ ν. We have also employed the identity Λννμμ=Ψννμμ=Γννμμ and phenomenologically added the terms αμνaμν/2, which account for waveguide losses through the linear absorption coefficients aμν.

There are several features of the derived equations that clearly distinguish them from the analogous equations for SOI waveguides [3, 56, 59, 63, 65, 79]. First, the factors ξ and ζ entering Eqs. (38) and (39) are responsible for the weakening of the nonlinear effects of Si NCs due to their embedment in fused silica. For small filling factors (f ≲ 0.01), the extent of weakening may be estimated using the approximate expressions ξ ≈ (3ε2)4f/(ε1 + 2ε2)4 ≈ 0.022 f and ζ9fε2ε1ε2/(ε1+2ε2)20.36f. We wish to iterate that the parameters βTPA, n2, χR, σn, and σα are the characteristics of individual nanocrystals, whereas the products ξβTPA, ξn2, ξχR, ζσn, and ζσα characterize the Si-NCs/SiO2 composite as a whole. Second, the strengths of self-phase modulation and TPA are reduced by about 45/(8 + 7ρ) ≈ 2.7 times, while the effects of cross-phase modulation and cross-TPA are enhanced by a factor of (26+16ρ)/45 ≈ 1.1, as compared to their values in SOI waveguides fabricated along the [01̄1] direction [3, 4, 79, 80]. Third, the second integral term on the right-hand side of Eq. (38) appears due to a modification of the interaction between the optical field and phonons inside the Si-NCs/SiO2 composite and is completely absent in the equations governing the nonlinear propagation through silicon. It should be also emphasized that the coupled amplitude equations written in a coordinate basis aligned with the waveguide edges are uniform, as the information on crystallographic directions in silicon is completely lost due to the random orientation of the crystallites in the composite.

Equations (38) and (39) provide a sound theoretical foundation for modeling nonlinear optical phenomena in Si-NCs/SiO2 waveguides and constitute the main result of this paper.

5.1. A continuous-wave regime

In the CW regime, when the amplitudes aμν are independent of time, the above set of equations can be simplified even further, by evaluating the two integral terms responsible for the Raman interaction as

4iε0ωμ45ξH˜(ωμμ)(2Γννμμ|aμν|2Γννμμ|aμν|2)aμν,
where
H˜(ω)=0H(t)eiωtdt=2χRΓRΩRΩR2+2iΓRωω2
is the Raman gain profile and χR is related to the the Raman gain coefficient gR as [4, 69]χR = 2ε0εeffc2gR/ωμ.

We can also formally solve Eq. (39) and introduce its solution into Eq. (38) to get a set of four coupled equations solely for the field amplitudes. In doing so, it is more convenient to use an approximate relation ΔnFC ≈ −σ̄n(ω0/ωμ)2N, with σ̄n = 5.3 × 10−27 m3[4] for typical free-carrier densities in the range 1022 to 1023 m−3 [the measurements of Spano et al.[71] give ζσ̄n = (1.2 ± 0.3) × 10−28 m3 for f = 8%]. Some algebra shows that Eqs. (38) and (39) in this case yield

lnaμνz+αμν2=ξ(βTPA2in2kμ)(8+7ρ452γνν(μ)Iμν+29+16ρ45γνν(μ)Iμν)+3245ξg˜R(ωμμ)(2γννμμIμνγννμμIμν)ζξτcneffnμν(ω0ωμ)2(σα2+iσ¯nkμ)×μ,νβTPA2h¯ωμ(8+7ρ452γνν(μ)Iμν+29+16ρ45γνν(μ)Iμν)Iμν,
where Iμν = |aμν|2/Aeff, γνν(μ)=γννμμ,
γννμμ=neff2nμνnμνAeff|eμν|2|eμν|2dr|eμν|2dr|eμν|2dr,
and
g˜R(ω)=2igRΓRΩRΩR2+2iΓRωω2.
For exact resonance, ωpωs = ΩR, the last expression gives g̃RR) = gR.

As we mentioned earlier, the quantities Iμν entering Eq. (42) do not represent modal intensities, as they are defined through the average EMA of the waveguide. One may rewrite the propagation equations in terms of the average field intensities ℐμν, by simply replacing Aeff with 𝒜effμν in the definition of the confinement factors γννμμ.

Although it is now possible to use the developed theory for modeling various nonlinear effects in silicon-nanocrystal waveguides, a thorough analysis of even the simplest propagation scenario (involving one continuous wave) is rather complicated and warrants a comprehensive discussion. We therefore intend to present such an analysis in future publications on nonlinear optics in the Si-NCs/SiO2 waveguides.

6. Conclusions

We have developed a general theory of two-frequency pulse propagation through silicon-nanocrystal-doped silica waveguides in the presence of the Kerr effect, stimulated Raman scattering, two-photon absorption (TPA), and the associated nonlinear optical phenomena. The dynamics of arbitrarily polarized pump and Stokes pulses was described using the system of four coupled nonlinear equations for the slowly varying amplitudes of the pulses, and the rate equation for the number density of TPA-generated free carriers. The equations were derived by assuming a uniform distribution of nanocrystal axes in space and treating the third-order nonlinear response of the waveguide within the framework of the effective medium theory. They allow for the modification of light interaction with electron and phonon subsystems of silicon nanocrystals, and take into account the weakening of the nanocrystals’ nonlinear response due to their embedment in silica matrix. The mathematics in the paper was dealt with in much detail, for methodological reasons and to facilitate the application of the developed theory by experimentalists. Our theory may be used to study a wide variety of nonlinear phenomena in silicon-nanocrystal silica waveguides, including the polarization rotation induced in optical pulses by self- and cross-phase modulation, Raman amplification and lasing, as well as TPA-induced optical modulation.

Acknowledgments

This work is sponsored by the Australian Research Council, through its Discovery Early Career Researcher Award DE120100055.

References and links

1. R. Soref and J. Lorenzo, “All-silicon active and passive guided-wave components for λ = 1.3 and 1.6 μm,” IEEE J. Quantum Electron. 22, 873–879 (1986). [CrossRef]  

2. J. Leuthold, C. Koos, and W. Freude, “Nonlinear silicon photonics,” Nat. Photonics 4, 535–544 (2010). [CrossRef]  

3. R. M. Osgood Jr., N. C. Panoiu, J. I. Dadap, X. Liu, X. Chen, I.-W. Hsieh, E. Dulkeith, W. M. Green, and Y. A. Vlasov, “Engineering nonlinearities in nanoscale optical systems: Physics and applications in dispersion-engineered silicon nanophotonic wires,” Adv. Opt. Photonics 1, 162–235 (2009). [CrossRef]  

4. Q. Lin, O. J. Painter, and G. P. Agrawal, “Nonlinear optical phenomena in silicon waveguides: Modeling and applications,” Opt. Express 15, 16604–16644 (2007). [CrossRef]   [PubMed]  

5. R. A. Soref, “The past, present, and future of silicon photonics,” IEEE J. Sel. Top. Quantum Electron. 12, 1678–1687 (2006). [CrossRef]  

6. G. T. Reed and A. P. Knights, Silicon Photonics: An Introduction (John Wiley, Hoboken, 2004). [CrossRef]  

7. L. Pavesi and D. Lockwood, eds., Silicon Photonics, vol. 94 of Topics in Applied Physics (Springer-Verlag, Berlin, 2004).

8. M. Paniccia, “Integrating silicon photonics,” Nat. Photonics 4, 498–499 (2010). [CrossRef]  

9. J. I. Dadap, N. C. Panoiu, X. Chen, I.-W. Hsieh, X. Liu, C.-Y. Chou, E. Dulkeith, S. J. McNab, F. Xia, W. M. J. Green, L. Sekaric, Y. A. Vlasov, and J. R. M. Osgood, “Nonlinear-optical phase modification in dispersion-engineered Si photonic wires,” Opt. Express 16, 1280–1299 (2008). [CrossRef]   [PubMed]  

10. C. Koos, L. Jacome, C. Poulton, J. Leuthold, and W. Freude, “Nonlinear silicon-on-insulator waveguides for all-optical signal processing,” Opt. Express. 15, 5976–5990 (2007). [CrossRef]   [PubMed]  

11. X. Chen, N. C. Panoiu, I. Hsieh, J. I. Dadap, and R. M. Osgood, “Third-order dispersion and ultrafast-pulse propagation in silicon wire waveguides,” IEEE Photon. Technol. Lett. 18, 2617–2619 (2006). [CrossRef]  

12. A. Martinez, J. Blasco, P. Sanchis, J. V. Galan, J. Garcia-Ruperez, E. Jordana, P. Gautier, Y. Lebour, S. Hernandez, R. Spano, R. Guider, N. Daldosso, B. Garrido, J. M. Fedeli, L. Pavesi, and J. Marti, “Ultrafast all-optical switching in a silicon-nanocrystal-based silicon slot waveguide at telecom wavelengths,” Nano Lett. 10, 1506–1511 (2010). [CrossRef]   [PubMed]  

13. C. Manolatou and M. Lipson, “All-optical silicon modulators based on carrier injection by two-photon absorption,” J. Lightwave Technol. 24, 1433–1439 (2006). [CrossRef]  

14. R. Jones, A. Liu, H. Rong, M. Paniccia, O. Cohen, and D. Hak, “Lossless optical modulation in a silicon waveguide using stimulated Raman scattering,” Opt. Express 13, 1716–1723 (2005). [CrossRef]   [PubMed]  

15. D. J. Moss, L. Fu, I. Littler, and B. J. Eggleton, “Ultrafast all-optical modulation via two-photon absorption in silicon-on-insulator waveguides,” Electron. Lett. 41, 320–321 (2005). [CrossRef]  

16. I. D. Rukhlenko, M. Premaratne, I. L. Garanovich, A. A. Sukhorukov, and G. P. Agrawal, “Analytical study of pulse amplification in silicon Raman amplifiers,” Opt. Express 18, 18324–18338 (2010). [CrossRef]   [PubMed]  

17. M. Krause, H. Renner, and E. Brinkmeyer, “Silicon Raman amplifiers with ring-resonator-enhanced pump power,” IEEE J. Sel. Top. Quantum Electron. 16, 216–225 (2010). [CrossRef]  

18. I. D. Rukhlenko, C. Dissanayake, M. Premaratne, and G. P. Agrawal, “Maximization of net optical gain in silicon-waveguide Raman amplifiers,” Opt. Express 17, 5807–5814 (2009). [CrossRef]   [PubMed]  

19. I. D. Rukhlenko, M. Premaratne, C. Dissanayake, and G. P. Agrawal, “Continuous-wave Raman amplification in silicon waveguides: Beyond the undepleted pump approximation,” Opt. Lett. 34, 536–538 (2009). [CrossRef]   [PubMed]  

20. M. Krause, H. Renner, S. Fathpour, B. Jalali, and E. Brinkmeyer, “Gain enhancement in cladding-pumped silicon Raman amplifiers,” IEEE J. Quantum Electron. 44, 692–704 (2008). [CrossRef]  

21. M. A. Foster, A. C. Turner, J. E. Sharping, B. S. Schmidt, M. Lipson, and A. L. Gaeta, “Broad-band optical parametric gain on a silicon photonic chip,” Nature 441, 960–963 (2006). [CrossRef]   [PubMed]  

22. J. I. Dadap, R. L. Espinola, R. M. Osgood, S. J. McNab, and Y. A. Vlasov, “Spontaneous Raman scattering in ultrasmall silicon waveguides,” Opt. Lett. 29, 2755–2757 (2004). [CrossRef]   [PubMed]  

23. D. Liang and J. E. Bowers, “Recent progress in lasers on silicon,” Nat. Photonics 4, 511–517 (2010). [CrossRef]  

24. H. Rong, S. Xu, O. Cohen, O. Raday, M. Lee, V. Sih, and M. Paniccia, “A cascaded silicon Raman laser,” Nat. Photonics 2, 170–174 (2008). [CrossRef]  

25. M. Krause, H. Renner, and E. Brinkmeyer, “Analysis of Raman lasing characteristics in silicon-on-insulator waveguides,” Opt. Express 12, 5703–5710 (2004). [CrossRef]   [PubMed]  

26. O. Boyraz and B. Jalali, “Demonstration of a silicon Raman laser,” Opt. Express 12, 5269–5273 (2004). [CrossRef]   [PubMed]  

27. M. W. Geis, S. J. Spector, M. E. Grein, J. U. Yoon, D. M. Lennon, and T. M. Lyszczarz, “Silicon waveguide infrared photodiodes with > 35 GHz bandwidth and phototransistors with 50 AW-1 response,” Opt. Express 17, 5193–5204 (2009). [CrossRef]   [PubMed]  

28. W. Astar, J. B. Driscoll, X. Liu, J. I. Dadap, W. M. J. Green, Y. A. Vlasov, G. M. Carter, and R. M. Osgood, “Conversion of 10 Gb/s NRZ-OOK to RZ-OOK utilizing XPM in a Si nanowire,” Opt. Express 17, 12987–12999 (2009). [CrossRef]   [PubMed]  

29. I. D. Rukhlenko, M. Premaratne, and G. P. Agrawal, “Analytical study of optical bistability in silicon-waveguide resonators,” Opt. Express 17, 22124–22137 (2009). [CrossRef]   [PubMed]  

30. H. K. Tsang and Y. Liu, “Nonlinear optical properties of silicon waveguides,” Semicond. Sci. Technol. 23, 064007 (2008). [CrossRef]  

31. M. A. Foster, A. C. Turner, M. Lipson, and A. L. Gaeta, “Nonlinear optics in photonic nanowires,” Opt. Express 16, 1300–1320 (2008). [CrossRef]   [PubMed]  

32. I.-W. Hsieh, X. Chen, X. Liu, J. I. Dadap, N. C. Panoiu, C.-Y. Chou, F. Xia, W. M. Green, Y. A. Vlasov, and R. M. Osgood, “Supercontinuum generation in silicon photonic wires,” Opt. Express 15, 15242–15249 (2007). [CrossRef]   [PubMed]  

33. R. Espinola, J. Dadap, R. Osgood Jr., S. McNab, and Y. Vlasov, “C-band wavelength conversion in silicon photonic wire waveguides,” Opt. Express 13, 4341–4349 (2005). [CrossRef]   [PubMed]  

34. M. W. Geis, S. J. Spector, R. C. Williamson, and T. M. Lyszczarz, “Submicrosecond submilliwatt silicon-on-insulator thermooptic switch,” IEEE Photon. Technol. Lett. 16, 2514–2516 (2004). [CrossRef]  

35. M. Dinu, F. Quochi, and H. Garcia, “Third-order nonlinearities in silicon at telecom wavelengths,” Appl. Phys. Lett. 82, 2954–2956 (2003). [CrossRef]  

36. I. D. Rukhlenko, W. Zhu, M. Premaratne, and G. P. Agrawal, “Effective third-order susceptibility of silicon-nanocrystal-doped silica,” Opt. Express 20, 26275–26284 (2012). [CrossRef]   [PubMed]  

37. L. Pavesi and R. Turan, eds., Silicon Nanocrystals: Fundamentals, Synthesis and Applications (WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, 2010).

38. L. Khriachtchev, ed., Silicon Nanophotonics: Basic Principles, Present Status and Perspectives (Pan Stanford, Singapore, 2009).

39. V. A. Belyakov, V. A. Burdov, R. Lockwood, and A. Meldrum, “Silicon nanocrystals: Fundamental theory and implications for stimulated emission,” Adv. Opt. Technol. 2008, 279502 (2008).

40. F. Iori, E. Degoli, R. Magri, I. Marri, G. Cantele, D. Ninno, F. Trani, O. Pulci, and S. Ossicini, “Engineering silicon nanocrystals: Theoretical study of the effect of codoping with boron and phosphorus,” Phys. Rev. B 76, 085302 (2007). [CrossRef]  

41. I. D. Rukhlenko and M. Premaratne, “Optimization of nonlinear performance of silicon-nanocrystal cylindrical nanowires,” IEEE Photonics J. 4, 952–959 (2012). [CrossRef]  

42. F. D. Leonardis and V. M. N. Passaro, “Dispersion engineered silicon nanocrystal slot waveguides for soliton ultrafast optical processing,” Adv. OptoElectron. 2011, 751498 (2011).

43. L. Sirleto, M. A. Ferrara, T. Nikitin, S. Novikov, and L. Khriachtchev, “Giant Raman gain in silicon nanocrystals,” Nat. Commun. 3, 1220 (2012). [CrossRef]   [PubMed]  

44. K. Imakita, M. Ito, R. Naruiwa, M. Fujii, and S. Hayashi, “Enhancement of ultrafast nonlinear optical response of silicon nanocrystals by boron-doping,” Opt. Lett. 37, 1877–1879 (2012). [CrossRef]   [PubMed]  

45. R. J. Kashtiban, U. Bangert, I. F. Crowe, M. Halsall, A. J. Harvey, and M. Gass, “Study of erbium-doped silicon nanocrystals in silica,” J. Phys.: Conference Series 241, 012097 (2010). [CrossRef]  

46. R. J. Walters, G. I. Bourianoff, and H. A. Atwater, “Field-effect electroluminescence in silicon nanocrystals,” Nat. Mater. 4, 143–146 (2005). [CrossRef]   [PubMed]  

47. T. Nikitin, R. Velagapudi, J. Sainio, J. Lahtinen, M. Räsänen, S. Novikov, and L. Khriachtchev, “Optical and structural properties of SiOx films grown by molecular beam deposition: Effect of the Si concentration and annealing temperature,” J. Appl. Phys. 112, 094316–094316 (2012). [CrossRef]  

48. J. Wei, J. Price, T. Wang, C. Hessel, and M. C. Downer, “Size-dependent optical properties of Si nanocrystals embedded in amorphous SiO2 measured by spectroscopic ellipsometry,” J. Vac. Sci. Technol. B 29, 04D112 (2011). [CrossRef]  

49. T. Nikitin, K. Aitola, S. Novikov, M. Räsänen, R. Velagapudi, J. Sainio, J. Lahtinen, K. Mizohata, T. Ahlgren, and L. Khriachtchev, “Optical and structural properties of silicon-rich silicon oxide films: Comparison of ion implantation and molecular beam deposition methods,” Phys. Status Solidi (a) 208, 2176–2181 (2011). [CrossRef]  

50. L. Ding, T. P. Chen, Y. Liu, C. Y. Ng, and S. Fung, “Optical properties of silicon nanocrystals embedded in a SiO2 matrix,” Phys. Rev. B 72, 125419 (2005). [CrossRef]  

51. I. D. Rukhlenko, M. Premaratne, and G. P. Agrawal, “Effective mode area and its optimization in silicon-nanocrystal waveguides,” Opt. Lett. 37, 2295–2297 (2012). [CrossRef]   [PubMed]  

52. F. Trani, D. Ninno, and G. Iadonisi, “Role of local fields in the optical properties of silicon nanocrystals using the tight binding approach,” Phys. Rev. B 75, 033312 (2007). [CrossRef]  

53. C. M. Dissanayake, I. D. Rukhlenko, M. Premaratne, and G. P. Agrawal, “Raman-mediated nonlinear interactions in silicon waveguides: Copropagating and counterpropagating pulses,” IEEE Photonics Technol. Lett. 21, 1372–1374 (2009). [CrossRef]  

54. I. D. Rukhlenko, M. Premaratne, and G. P. Agrawal, “Nonlinear propagation in silicon-based plasmonic waveguides from the standpoint of applications,” Opt. Express 19, 206–217 (2011). [CrossRef]   [PubMed]  

55. I. D. Rukhlenko, M. Premaratne, and G. P. Agrawal, “Analytical study of optical bistability in silicon ring resonators,” Opt. Lett. 35, 55–57 (2010). [CrossRef]   [PubMed]  

56. B. A. Daniel and G. P. Agrawal, “Vectorial nonlinear propagation in silicon nanowire waveguides: Polarization effects,” J. Opt. Soc. Am. B 27, 956–965 (2010). [CrossRef]  

57. I. D. Rukhlenko, C. Dissanayake, M. Premaratne, and G. P. Agrawal, “Optimization of Raman amplification in silicon waveguides with finite facet reflectivities,” IEEE J. Sel. Top. Quantum Electron. 16, 226–233 (2010). [CrossRef]  

58. I. D. Rukhlenko, I. Udagedara, M. Premaratne, and G. P. Agrawal, “Effect of free carriers on pump-to-signal noise transfer in silicon Raman amplifiers,” Opt. Lett. 35, 2343–2345 (2010). [CrossRef]   [PubMed]  

59. M. D. Turner, T. M. Monro, and S. Afshar V., “A full vectorial model for pulse propagation in emerging waveguides with subwavelength structures part II: Stimulated Raman scattering,” Opt. Express 17, 11565–11581 (2009). [CrossRef]   [PubMed]  

60. L. Yin, J. Zhang, P. M. Fauchet, and G. P. Agrawal, “Optical switching using nonlinear polarization rotation inside silicon waveguides,” Opt. Lett. 34, 476–478 (2009). [CrossRef]   [PubMed]  

61. I. D. Rukhlenko, M. Premaratne, C. Dissanayake, and G. P. Agrawal, “Nonlinear pulse evolution in silicon waveguides: An approximate analytic approach,” J. Lightwave Technol. 27, 3241–3248 (2009). [CrossRef]  

62. D. Dimitropoulos, B. Houshmand, R. Claps, and B. Jalali, “Coupled-mode theory of Raman effect in silicon-on-insulator waveguides,” Opt. Lett. 28, 1954–1956 (2003). [CrossRef]   [PubMed]  

63. S. Afshar V. and T. M. Monro, “A full vectorial model for pulse propagation in emerging waveguides with subwavelength structures part I: Kerr nonlinearity,” Opt. Express 17, 2298–2318 (2009). [CrossRef]   [PubMed]  

64. R. W. Boyd, Nonlinear Optics, 3rd ed. (Academic Press, San Diego, 2008).

65. X. Chen, N. C. Panoiu, and R. M. Osgood Jr., “Theory of Raman-mediated pulsed amplification in silicon-wire waveguides,” IEEE J. Quantum Electron. 42, 160–170 (2006). [CrossRef]  

66. S. N. Volkov, J. J. Saarinen, and J. E. Sipe, “Effective medium theory for 2D disordered structures: A comparison to numerical simulations,” J. Mod. Opt. 59, 954–961 (2012). [CrossRef]  

67. X. C. Zeng, D. J. Bergman, P. M. Hui, and D. Stroud, “Effective-medium theory for weakly nonlinear composites,” Phys. Rev. B 38, 10970–10973 (1988). [CrossRef]  

68. W. Cai and V. Shalaev, Optical Metamaterials: Fundamentals and Applications (Springer, New York, 2010).

69. C. M. Dissanayake, M. Premaratne, I. D. Rukhlenko, and G. P. Agrawal, “FDTD modeling of anisotropic nonlinear optical phenomena in silicon waveguides,” Opt. Express 18, 21427–21448 (2010). [CrossRef]   [PubMed]  

70. F. Trojanek, K. Neudert, K. Zidek, K. Dohnalova, I. Pelant, and P. Maly, “Femtosecond photoluminescence spectroscopy of silicon nanocrystals,” Physica Status Solidi (c) 3, 3873–3876 (2006). [CrossRef]  

71. R. Spano, N. Daldosso, M. Cazzanelli, L. Ferraioli, L. Tartara, J. Yu, V. Degiorgio, E. Giordana, J. M. Fedeli, and L. Pavesi, “Bound electronic and free carrier nonlinearities in silicon nanocrystals at 1550 nm,” Opt. Express 17, 3941–3950 (2009). [CrossRef]   [PubMed]  

72. M. A. Ferrara, I. Rendina, S. N. Basu, L. D. Negro, and L. Sirleto, “Raman amplifier based on amorphous silicon nanoparticles,” Int. J. Photoenergy 2012, 254946 (2012). [CrossRef]  

73. M. A. Ferrara, I. Rendina, and L. Sirleto, “Stimulated Raman scattering in quantum dots and nanocomposite silicon based materials,” in “Nonlinear Optics,” N. Kamanina, ed. (InTech, Rijeka, 2012), pp. 53–70.

74. H. Richter, Z. P. Wang, and L. Ley, “The one phonon Raman spectrum in microcrystalline silicon,” Solid State Commun. 39, 625–629 (1981). [CrossRef]  

75. I. H. Campbell and P. M. Fauchet, “The effects of microcrystal size and shape on the one phonon Raman spectra of crystalline semiconductors,” Solid State Commun. 58, 739–741 (1986). [CrossRef]  

76. L. Cao, B. Nabet, and J. E. Spanier, “Enhanced Raman scattering from individual semiconductor nanocones and nanowires,” Phys. Rev. Lett. 96, 157402 (2006). [CrossRef]   [PubMed]  

77. R. Hillenbrand, T. Taubner, and F. Keilmann, “Phonon-enhanced light-matter interaction at the nanometre scale,” Nature 418, 159–162 (2002). [CrossRef]   [PubMed]  

78. G. P. Agrawal, Nonlinear Fiber Optics (Academic Press, San Diego, 2007).

79. I. D. Rukhlenko, M. Premaratne, and G. P. Agrawal, “Nonlinear silicon photonics: Analytical tools,” IEEE J. Sel. Top. Quantum Electron. 16, 200–215 (2010). [CrossRef]  

80. I. D. Rukhlenko, I. L. Garanovich, M. Premaratne, A. A. Sukhorukov, G. P. Agrawal, and Y. S. Kivshar, “Polarization rotation in silicon waveguides: Analytical modeling and applications,” IEEE Photonics J. 2, 423–435 (2010). [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.


Equations (52)

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

× E ˜ μ ( 0 ) ( r , ω ) = i ω μ 0 H ˜ μ ( 0 ) ( r , ω )
× H ˜ μ ( 0 ) ( r , ω ) = i ω ε 0 ε L ( r , ω ) E ˜ μ ( 0 ) ( r , ω ) ,
E ˜ μ ( r , ω ) = ν a ˜ μ ν ( z , ω ω μ ) e μ ν ( r , ω μ ) N μ ν e i β μ ν z
H ˜ μ ( r , ω ) = ν a ˜ μ ν ( z , ω ω μ ) h μ ν ( r , ω μ ) N μ ν e i β μ ν z ,
( e μ ν * × h μ ν + c . c . ) d r = δ ν ν ( e μ ν * × h μ ν + c . c . ) d r = 4 N μ ν ,
P μ = 1 2 Re ( E ˜ μ × H ˜ μ * ) d r = ν | a ˜ μ ν | 2 .
× E ˜ μ ( r , ω ) = i ω μ 0 H ˜ μ ( r , ω )
× H ˜ μ ( r , ω ) = i ω ε 0 ε L ( r , ω ) E ˜ μ ( r , ω ) i ω P ˜ μ NL ( r , ω ) .
z ( E ˜ μ ( 0 ) × H ˜ μ * + E ˜ μ * × H ˜ μ ( 0 ) ) d r = ( E ˜ μ ( 0 ) × H ˜ μ * + E ˜ μ * × H ˜ μ ( 0 ) ) d r .
( E ˜ μ ( 0 ) × H ˜ μ * + E ˜ μ * × H ˜ μ ( 0 ) ) = i ω ( E ˜ μ ( 0 ) P ˜ μ N L * ) .
E ˜ μ ( 0 ) = e μ ν ( r , ω μ ) N μ ν e i β ν ( ω ) z and H ˜ μ ( 0 ) = h μ ν ( r , ω μ ) N μ ν e i β ν ( ω ) z
z { a ˜ μ ν ( z , ω ω μ ) e i [ β ν ( ω μ ) β ν ( ω ) ] z } = i ω 4 N μ ν e μ ν * ( r , ω μ ) P ˜ μ NL ( r , ω ) e i β ν ( ω ) z d r .
( z + i n = 1 1 n ! β ν ω | ω μ ( ω ω μ ) n ) a ˜ μ ν ( z , ω ω μ ) = i ω 4 N μ ν e μ ν * ( r , ω μ ) P ˜ μ NL ( r , ω ) e i β ν ( ω μ ) z d r .
( z + n = 1 i n + 1 n ! β ν ω | ω μ n t n ) a μ ν ( z , t ) = e i β μ ν z 4 N μ ν e μ ν * ( r , ω μ ) P μ NL ( r , t ) t e i ω μ t d r ,
a μ ν ( z , t ) = 1 2 π + a ˜ μ ν ( z , ω ) e i ω t d ω and P μ NL ( r , t ) = 1 2 π + P ˜ μ NL ( r , ω ) e i ω t d ω .
a μ ν z + n = 1 i n + 1 β μ ν ( n ) n ! n a μ ν t n = i ω μ 4 N μ ν ( 1 + i ω μ t ) e i β μ ν z ( e μ ν * P ω μ NL ) d r ,
P ω μ NL ( r , t ) = P ω μ K ( r , t ) + P ω μ R ( r , t ) + P ω μ FC ( r , t ) ,
P ω μ K ( r , t ) = ε 0 χ K ( 3 ) ( ω μ ; ω μ , ω μ , ω μ ) E ω μ ( r , t ) E ω μ * ( r , t ) E ω μ ( r , t ) ,
χ K ( 3 ) ( ω μ ; ω μ , ω μ , ω μ ) = χ μ ( 8 + 7 ρ 45 ( δ k l δ m n + δ k m δ ln + δ k n δ l m ) + 1 ρ 9 δ k l δ l m δ m n ) ,
χ μ = c ε 0 ε eff [ n 2 + i β TPA / ( 2 k μ ) ] ξ ,
ξ = 1 f ( ε eff ε 1 ) 2 = [ ( 3 f 1 ) ε eff + ε 2 ] 2 f ( u 2 + 8 ε 1 ε 2 )
P ω μ K = ε 0 χ μ [ 8 + 7 ρ 45 ( 2 | E ω μ | 2 E ω μ + E ω μ 2 E ω μ * ) + 1 ρ 9 η E η 2 E η * η ^ ] ,
E ω μ = ν = x , y a μ ν e μ ν N μ ν e i β μ ν z ,
e i β μ ν z N μ ν ( e μ ν * P ω μ K ) d r = ε 0 χ μ [ 8 + 7 ρ 45 ( 2 a μ ν ν Γ ν ν ( μ ) | a μ ν | 2 + a μ ν * ν a μ ν 2 Λ ν ν ( μ ) e 2 i ( β μ ν β μ ν ) z ) + 1 ρ 9 Γ ν ν ( μ ) a μ ν | a μ ν | 2 ] ,
Γ ν ν ( μ ) = 1 N μ ν N μ ν | e μ ν | 2 | e μ ν | 2 d r
Λ ν ν ( μ ) = 1 N μ ν N μ ν e μ ν * 2 e μ ν 2 d r .
N μ ν = β μ ν 2 μ 0 ω μ | e μ ν | 2 d r .
P ω μ R ( r , t ) = e i ω μ t t d t 1 t d t 2 t d t 3 ε 0 χ R ( 3 ) ( t t 1 , t t 2 , t t 3 ) × ( E ω μ ( r , t 1 ) E ω μ * ( r , t 2 ) E ω μ ( r , t 3 ) e i ( ω μ t 1 ω μ t 2 + ω μ t 3 ) + E ω μ ( r , t 1 ) E ω μ * ( r , t 2 ) E ω μ ( r , t 3 ) e i ( ω μ t 1 ω μ t 2 + ω μ t 3 ) ) ,
χ R ( 3 ) ( t 1 , t 2 , t 3 ) = 1 2 [ δ ( t 1 t 2 ) δ ( t 3 ) k l m n + δ ( t 1 ) δ ( t 2 t 3 ) k n m l ] ξ H ( t 2 ) ,
k l m n = 29 45 ( δ k m δ ln + δ k n δ l m ) 16 45 δ k l δ m n 2 9 δ k l δ l m δ m n ,
H ( t ) = 2 χ R Γ R Ω R ( Ω R 2 Γ R 2 ) 1 / 2 e t / τ 2 sin ( t / τ 1 ) ,
P ω μ R ( r , t ) = ε 0 ξ t d t 1 H ( t t 1 ) k l m n E ω μ ( r , t 1 ) E ω μ * ( r , t 1 ) E ω μ ( r , t ) e i ( ω μ ω μ ) ( t t 1 ) .
e i β μ ν z ε 0 ξ N μ ν ( e μ ν * P ω μ R ) d r = 29 45 ν Λ ν ν μ μ exp ( i β μ ν μ ν μ ν μ ν z ) a μ ν ( t ) t a μ ν * ( t 1 ) a μ ν ( t 1 ) H ( t t 1 ) e i ω μ μ ( t t 1 ) d t 1 + 29 45 ν Ψ ν ν μ μ exp ( i β μ ν μ ν μ ν μ ν z ) a μ ν ( t ) t a μ ν * ( t 1 ) a μ ν ( t 1 ) H ( t t 1 ) e i ω μ μ ( t t 1 ) d t 1 16 45 ν Γ ν ν μ μ a μ ν ( t ) t a μ ν * ( t 1 ) a μ ν ( t 1 ) H ( t t 1 ) e i ω μ μ ( t t 1 ) d t 1 2 9 Λ ν ν μ μ a μ ν t t a μ ν * ( t 1 ) a μ ν ( t 1 ) H ( t t 1 ) e i ω μ μ ( t t 1 ) d t 1 ,
Λ ν ν μ μ = ( e μ ν * e μ ν * ) ( e μ ν e μ ν ) N μ ν N μ ν N μ ν N μ ν d r ,
Ψ ν ν μ μ = ( e μ ν * e μ ν ) ( e μ ν e μ ν * ) N μ ν N μ ν N μ ν N μ ν d r ,
Γ ν ν μ μ = 1 N μ ν N μ ν | e μ ν | 2 | e μ ν | 2 d r .
P ω μ FC ( r , t ) = 2 ζ ε 0 n eff [ Δ n FC + i c / ( 2 ω μ ) Δ α FC ] E ω μ ( r , t ) ,
ζ = n eff n 1 = ( ε 1 ε eff ) 1 / 2 ( 3 f 1 ) ε eff + ε 2 u 2 + 8 ε 1 ε 2
Δ n FC = σ n ( ω 0 / ω μ ) 2 ( 1 + ς N 0.2 ) N 0.8 and Δ α FC = σ α ( ω 0 / ω μ ) 2 N ,
e i β μ ν z N μ ν ( e μ ν * P ω μ FC ) d r = 4 ( ζ / c ) ( n eff / n μ ν ) [ Δ n FC + i c / ( 2 ω μ ) Δ α FC ] a μ ν ,
N t = N τ c μ 1 2 h ¯ ω μ A eff P μ z ,
P μ z = ν ( a μ ν a μ ν * z + a μ ν * a μ ν z ) = 1 4 ξ c 2 ε 0 2 ε eff β TPA ν ( 8 + 7 ρ 45 ν 2 Γ ν ν ( μ ) | a μ ν | 2 | a μ ν | 2 + 13 + 2 ρ 45 Γ ν ν ( μ ) | a μ ν | 4 ) .
A eff μ ν = ( | e μ ν | 2 d r ) 2 / | e μ ν | 4 d r .
A eff = μ , ν ( A eff μ ν ) 1 / 4 .
𝒜 eff μ ν = 𝒜 NL | e μ ν | 2 d r / NL | e μ ν | 2 d r ,
a μ ν z + n = 1 i n + 1 β μ ν ( n ) n ! n a μ ν t n + α μ ν 2 a μ ν = 1 8 ξ c 2 ε 0 2 ε eff ( β TPA 2 i n 2 k μ ) ( 8 + 7 ρ 45 2 Γ ν ν ( μ ) | a μ ν | 2 + 29 + 16 ρ 45 Γ ν ν ( μ ) | a μ ν | 2 ) a μ ν + 8 i ε 0 ω μ 45 ξ Γ ν ν μ μ a μ ν ( t ) t a μ ν * ( t 1 ) a μ ν ( t 1 ) H ( t t 1 ) e i ω μ μ ( t t 1 ) d t 1 4 i ε 0 ω μ 45 ξ Γ ν ν μ μ a μ ν ( t ) t a μ ν * ( t 1 ) a μ ν ( t 1 ) H ( t t 1 ) e i ω μ μ ( t t 1 ) d t 1 ζ n eff n μ ν ( ω 0 ω μ ) 2 ( i σ n k μ ( 1 + ς N 0.2 ) + σ α 2 N 0.2 ) N 0.8 a μ ν ,
N t = N τ c + ξ c 2 ε 0 2 ε eff 4 A eff μ , ν β TPA 2 h ¯ ω μ ( 8 + 7 ρ 45 2 Γ ν ν ( μ ) | a μ ν | 2 + 29 + 16 ρ 45 Γ ν ν ( μ ) | a μ ν | 2 ) | a μ ν | 2 .
4 i ε 0 ω μ 45 ξ H ˜ ( ω μ μ ) ( 2 Γ ν ν μ μ | a μ ν | 2 Γ ν ν μ μ | a μ ν | 2 ) a μ ν ,
H ˜ ( ω ) = 0 H ( t ) e i ω t d t = 2 χ R Γ R Ω R Ω R 2 + 2 i Γ R ω ω 2
ln a μ ν z + α μ ν 2 = ξ ( β TPA 2 i n 2 k μ ) ( 8 + 7 ρ 45 2 γ ν ν ( μ ) I μ ν + 29 + 16 ρ 45 γ ν ν ( μ ) I μ ν ) + 32 45 ξ g ˜ R ( ω μ μ ) ( 2 γ ν ν μ μ I μ ν γ ν ν μ μ I μ ν ) ζ ξ τ c n eff n μ ν ( ω 0 ω μ ) 2 ( σ α 2 + i σ ¯ n k μ ) × μ , ν β TPA 2 h ¯ ω μ ( 8 + 7 ρ 45 2 γ ν ν ( μ ) I μ ν + 29 + 16 ρ 45 γ ν ν ( μ ) I μ ν ) I μ ν ,
γ ν ν μ μ = n eff 2 n μ ν n μ ν A eff | e μ ν | 2 | e μ ν | 2 d r | e μ ν | 2 d r | e μ ν | 2 d r ,
g ˜ R ( ω ) = 2 i g R Γ R Ω R Ω R 2 + 2 i Γ R ω ω 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.