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

Fresnel phase matching: Exploring the frontiers between ray and guided wave quadratic nonlinear optics

Open Access Open Access

Abstract

Fresnel phase matching is a convenient and universal way to phase match nonlinear three-wave mixing by total internal reflection in isotropic materials like common semiconductors. This technique makes use of the large relative phase lag between the interacting waves at total internal reflection, and was suggested by the nonlinear optics pioneers in the 70’s; it has been worked out by several teams since then but, quite unexpectedly, has never succeeded in producing enough parametric gain to achieve optical parametric oscillation. We show that this failure stems mostly from a basic law of nonlinear reflection, which leads to a spatial walk-off between the pump and the generated parametric waves, resulting in unexpected destructive interference patterns between the waves while bouncing back and forth between the interfaces. Ray tracing or plane wave analysis gives an incomplete representation of the phenomenon while highly multimodal nonlinear guided wave theory reconciles the different views. Very good agreement between the presented theory and experiments is demonstrated in gallium arsenide samples.

©2008 Optical Society of America

1. Introduction

Mid-infrared tunable sources are of considerable interest for applications in environmental monitoring. Indeed, the main toxic agents and pollutants present strong absorbing lines in the spectral region from 3 to 12 µm. These lines are moreover highly differentiated from one species to the other, so that each pollutant has its own spectral signature (or so-called spectral fingerprint). The spectroscopic interest is reinforced by the high transmissivity of the atmosphere in the corresponding spectral band, allowing for instance LIDAR applications.

Optical parametric conversion appears as a promising solution to cover this wide spectral region: one possible scheme is the down conversion of near-2 µm wave into the mid-infrared, which could be obtained either by difference frequency generation or by parametric fluorescence, leading to optical parametric oscillation. Semiconductors of the technological mainstream (such as gallium arsenide GaAs, indium phosphide InP or zinc selenide ZnSe) are attractive candidates for optical parametric conversion of mid-infrared waves. Indeed, they are transparent and display very low dispersion in the near- and mid-infrared region; they possess high nonlinear 2nd-order coefficients χ (2), and benefit from a mature technology. Moreover, they can stand a high incident energy flux. However, these materials are optically isotropic, so that no natural birefringence phase-matching scenario is possible. Nevertheless, because of the importance of the subject, various solutions have been proposed to get an efficient conversion, among which: artificial birefringence, modal phase-matching, as well as several quasi-phase-matching techniques [13].

In their founding paper, Armstrong et al suggested using the relative phase change between the parametric and the fundamental waves upon total internal reflection to compensate their phase mismatch [4]. This principle was demonstrated by Boyd and Patel as well as by Komine et al by phase matching second harmonic generation in isotropic semiconductors (GaAs, ZnSe) [5,6]. This approach was generalized to three-wave mixing by our group [7]. We showed that this technique displayed a certain number of advantages compared to the previous ones given in references [13]. Firstly, the technique is poorly demanding on technology; secondly, unmatched spectral tuning can be obtained in the mid-infrared, (experimentally from 8 to 13 µm tunability with a single wafer while theory predicts tunability even in the terahertz range). Rather good conversion efficiency was obtained by the different groups though, quite unexpectedly, no one ever succeeded in producing enough parametric gain to achieve optical parametric oscillation. In this paper, we show that this failure stems mostly from a basic law of nonlinear reflection already discovered by Bloembergen and his co-workers [8]. This latter phenomenon leads to a non collinear propagation of the pump and the generated parametric waves. As a consequence, the generated beams walk-off during propagation in the slab, resulting in unexpected destructive interference patterns between the waves while bouncing back and forth between the interfaces.

In Section 2, we briefly recall the basic principles of Fresnel phase matching (FPM). We present in Section 3 a second harmonic generation (SHG) experiment that shows that the theoretical models presented so far are not sufficient to describe the FPM interaction. In particular, in Section 4, an intuitive description of the influence of nonlinear total internal reflection on the parametric conversion yield is shown. This part will serve as a guideline to the development of the multimodal theory which is given in Section 5. Here, an analytical theory based on the infinitely confining and highly confining waveguide assumptions is first developed. This allows us to show the main physical insights of FPM. Numerical models, based on this guided wave theory, of the SHG experiment are then presented and performed. Finally, the comparison between these theoretical predictions and the experimental results are presented in Section 6.

2. Principles of Fresnel phase matching

Fresnel phase-matching makes use of the differential Fresnel phase shift Δϕ F experienced by three interacting waves (ω 1>ω 2>ω 3) at total internal reflection on the air-semiconductor interface (Fig. 1). This Fresnel phase shift Δϕ F may be very large (particularly near the critical angle). It can thus compensate for any dispersion phase-mismatch Δk=k 1-k 2-k 3, and allows a quasi-phase-matched growth of the conversion signal: the wave vectors i k at pulsation ωi are given by k i=(ω ni/c)e i where c is the light velocity, i e and ni are the unitary direction vector and the materials optical index at angular frequency ωi respectively.

 figure: Fig. 1.

Fig. 1. Scheme of the Fresnel phase-matching configuration for SHG and DFG: geometry of the Fresnel phase-matched plate: (a) side view for geometrical description; and (b) view from above for angle definition.

Download Full Size | PDF

Let us consider the general three waves interaction scheme: ω 1-ω 2ω 3 (and correlatively ω 3+ω 2ω 1). Two cases can be analysed:

- ω 3=ω 2=ω and ω 1=2 ω: this represents the second harmonic generation (SHG) case treated by [5,6]. Such case is easier to analyse and will be used here in order to highlight the physics involved in the process.

- ω 1 and ω 2 are, for instance, output waves of a nearly degenerated optical parametric oscillator of pump angular frequency ω 0(ω 1ω 2ω 0/2) and ω 3 is the difference frequency generated (DFG) angular frequency. This case is of a more important practical importance but is more complex to analyse.

The two pumping waves are injected into the plate of thickness t by one of the slanted faces (Fig. 1). Due to the high optical index of the material, a total internal reflection situation is easily reached (internal reflection angleθ). The three waves then experience multiple bounces all through the plate. The global phase shift Δϕ between two successive bounces is the sum of the three following contributions:

Δϕ=ΔkLb+ΔϕF+δϕ.

Lb being the distance between two successive bounces given by Lb=t cosθ. ΔϕF is given by ΔϕF=ϕ 1 F-ϕ 2 F-ϕ 3 F, where ϕiF is the Fresnel phase shift experienced by the wave of angular frequency ωi at total internal reflection at the semiconductor-air interface which depends on θ and on the different polarizations of the waves [9]. There is a third possible contribution: the effective nonlinear coefficient d eff may change value after reflection. It may be a mere change of sign, which introduces an additional phase shift δϕ=π.

The Fresnel phase shift ΔϕF depends thus on (i) the reflection angle θ, which must be greater than the critical angle for each wave ωi(θθc=arcsin(n ext/n(ωi)), n ext being the optical index of the outside medium i.e. n ext=1 in this paper) ; (ii) the polarization of the three waves. This last dependency is responsible for the so-called Fresnel birefringence.

Because of the large variation of ΔϕF as a function of θ, Eq. (1) indicates that almost any phase mismatch Δk may be compensated, which makes Fresnel phase matching a very versatile and universal phase matching technique [7].

As demonstrated in reference [7], computation of the interacting process provides the proportionality factor between the incoming pump waves intensities I in 1 and I in 2, and the outgoing wave intensity I out 3:

I3out=Z0ω322c2(Ndeff)2n1n2n3(sin(ΔkLb2)sin(Δk2))2(sin(NΔϕ2)Nsin(Δϕ2))2I1inI2in,

where Z0 is the vacuum impedance (Z0=377Ω). The global parametric process efficiency η is the thus product of two main parts. Firstly, there is the parametric conversion on each path Lb between two bounces, so that the first term of the conversion yield, writes [sin(ΔkLb/2)/(Δk/2)]2. Then, the individual fields generated along all the paths will interfere with each other. This introduces the term: {sin(NΔϕ/2)/[Nsin(Δϕ/2)]}2, where N is the number of bounces inside the plate. In reference [7], we showed that resonant phase matching (ΔkLb=(2 p+1)π) and non resonant phase matching (ΔkLb≠(2 p+1)π), with p integer, were both interesting cases, which increase the versatility of the technique. As shown in Fig. 2, resonant phase matching corresponds to the optimal condition where the distance Lb between two successive bounces is exactly an odd number of coherence lengths Λc(=πk) of the nonlinear process.

 figure: Fig. 2.

Fig. 2. Resonant Fresnel phase-matching: the distance Lb between two successive bounces is optimized (Lb=(2p+1) Λc, where p is an integer and Λc the coherence length of the nonlinear interaction).

Download Full Size | PDF

Equation (2) also indicates that, when the Fresnel phase matching condition is met, the signal should grow quadratically with the number of bounces N, so that large conversion efficiencies could be obtained for long samples. The sole condition for Fresnel phase-matching (FPM) is:

Δϕ(θ)=ΔkLb(θ)+ΔϕF(θ)+δϕ=0[2π].

However, such a large conversion yield has never been achieved in such a way that optical parametric oscillation could be obtained. In Ref. [7], we have investigated some possible limitations of this efficiency: (i) the surface roughness and (ii) the Goos-Hänchen effect which can lead to a spatial walk off between the waves. However, these effects could not explain a lack of a factor greater than 10 between the ray theory and the experimental data that were reported in [10] and are recalled in Sec. 3. We show here that another effect introduces a more sever limitation: the nonlinear reflection law at the interfaces.

3. Second harmonic generation experiment

In order to investigate in a more systematic way the discrepancies between the results given by the ray theory presented in Sec. 2 and those obtained experimentally, we carried out several second harmonic generation (SHG) experiments in a fixed FPM configuration varying only the nonlinear material sample length while other parameters were kept constant.

 figure: Fig. 3.

Fig. 3. Experimental set-up: (a) nonlinear material samples geometry, the only varying parameter is the length L, (b) SHG experimental setup, (c) single coherence length intensity calibration experiment.

Download Full Size | PDF

The experimental setup and the obtained results were recently presented in [10], and they are only briefly recalled in the present paper. The experiments were carried out using the 4 µm idler beam emitted by a PPLN optical parametric oscillator (OPO) as the fundamental radiation of the SHG process. The OPO is pumped by a Nd:YAG laser with a 10 kHz repetition rate, the pulse duration at 4 µm is typically 150 ns and the emitted idler beam is nearly diffraction limited (M 2≤1.3).

The experimental set-up is shown in Fig. 3. The p-polarized pump beam is focalised into a 100 µm waist inside GaAs samples of different lengths. The crystals facets are bevelled with an angle of 28.5° (Fig.3(a)), and mounted on a rotation stage, allowing us to inject the pump beam at the expected resonant Fresnel phase matching angle of 32.5°. The remaining 4 µm pump beam is filtered out using glass filters at the output of the plate. The 2 µm second harmonic (SH) signal is detected using a thermoelectrically cooled HgCdTe detector (Fig. 3(b)).

In order to facilitate the comparison with the modelling results (see Sec. 5.2), the measured SH intensity is normalized to the SH intensity that is generated in a single coherence length. This normalising intensity was determined by maximizing the SHG in a GaAs prism in a single-pass configuration with the same propagation angle relatively to the crystallographic axes as in the GaAs plates (Fig. 3(c)). The obtained ratio thus represents the square of an efficient number of bounces in the plate.

The measured normalized SH intensity is plotted as a function of the propagation angle and the crystal length in Fig. 4, as well as the calculated signal expected from the ray theory. First, let us notice that, apart from a 0.3° error due to both the uncertainty on the bevel angle and on the angle rotation measurement, the FPM condition arises at the expected angle, and with the correct acceptance angle (Fig. 4(a)). In terms of sample length influence, we can observe two regimes: for short crystal length (<15 mm), the generated intensity varies quadratically with the crystal length. In this case, the measured number of bounces is around 1.8-fold smaller than the number of bounces expected from the ray theory. For longer crystals, the SHG signal does not follow a quadratic law any more and clearly exhibits saturation that is not predicted by the ray theory, leading to a measured SHG signal up to 6- fold smaller (Fig. 4(b)).

 figure: Fig. 4.

Fig. 4. Experimental results of 4→2 µm FPM SHG experiment: SH intensity as a function of the internal angle (a), SHG intensity normalized to one coherence length intensity as a function of the sample length (b).

Download Full Size | PDF

In order to rule out diffraction effects as a possible cause of the unusual variation of the SH-wave intensity as a function of the fundamental-wave intensity, we also recorded the experimental pump and SHG profiles by use of a SWIR camera. As shown in Fig. 5, we can notice that the beam quality of the pump is not degraded by the propagation inside the plate and that the SHG beam is also of good spatial quality. This is consistent with the fact that the Rayleigh distance of the pump beam in GaAs is larger or at least equal to the length of our samples.

 figure: Fig. 5.

Fig. 5. Spatial profiles of the pump beam (using an infrared Pyrocam camera) (a) and SHG beam (using a SWIR camera) (b), at the output of the GaAs crystal. Note that the profiles are only given here for a qualitative appreciation of the beam quality, not to give an actual value of the beams sizes.

Download Full Size | PDF

4. Nonlinear reflection law

As noted by Bloembergen and Pershan, when a beam of light is reflected onto nonlinear materials, the laws of reflections must take into account the conservation of the component of the wave momentum parallel to the boundary [8]. This condition implies that, besides the collinear field generated in the bulk, at the interface the nonlinear polarisation driving term radiates an “homogeneous” SH field that satisfies the following generalized Snell-Descartes law, which is visualized in Fig. 6 for a SHG situation:

nωsinθω=n2ωsinθ2ω.

This equation shows that, because of optical dispersion, the only condition for the SH wave to be a plane wave is to be in a situation where the pump wave incidence angle θω and the SHG incidence angle θ 2ω are not equal and must satisfy Eq. (4). In this manner, the reflected SH wave is collinear with the “homogeneous” SH wave generated at the interface and a plane wave propagation is preserved. The required small variation in incidence angle δθ is trivially obtained by a first order expansion, so that:

δθ(δnn)tanθω,

where δn=n 2ω-nω is the optical dispersion and nn 2ωn ω.

 figure: Fig. 6.

Fig. 6. Nonlinear total internal reflection law

Download Full Size | PDF

A more elaborate approach is to unfold the zig-zag motion of the waves in the plate, as represented in Fig. 7. The incident collinear waves ω 1 and ω 2 propagate at an angle θ relatively to the virtual stack. The periodicity of the stack introduces a reciprocal lattice vector which has to be determined. The exact calculation is done in the Appendix A. It leads to a vector conservation law given by:

k1+KF=k2+k3
KF=ΔϕFt,

where K F is the reciprocal lattice vector due to the periodic Fresnel phase shift at the interfaces. This equation is a generalisation of the vector equation given by Fejer et al in reference [11] and leads to the same expression as (4). Equation (6) shows that, indeed, the nonlinear wave is non collinear relatively to the pump waves and that FPM Eq. (2) is an approximation which does not take the non collinearity of the waves into account.

 figure: Fig. 7.

Fig. 7. Unfolding of the zig-zag motion of the waves at total internal reflection (a). Because of the nonlinear law of reflection, the waves suffer from a non collinear propagation due to the periodicity of the structure (b).

Download Full Size | PDF

Non collinear angles of 1 to 2 degrees are thus expected in our experimental conditions, which is indeed quite important. This non collinearity leads to two effects. The first one is a spatial walk-off which decreases the overlap between the interacting beams. This effect has been evaluated using the same expression as for the birefringence walk-off correction term developed in reference [12] and leads typically to an efficiency decrease of a factor 2. It cannot alone explain the large factor 10 observed with respect to the experiments. Moreover, as shown in Fig. 8, the pump and parametric beams may recombine after a certain number of bounces: the generated beam then fills the plate completely. This recombination may lead to complex destructive interferences which may explain the small efficiency experimentally determined. The number of bounces necessary for such a recombination may be roughly evaluated. As shown in Fig. 8, the recombination occurs when the SH wave generated at the beginning of the plate and the fundamental wave experience numbers of bounces that differ form 2 for a given propagation length, so that Nsinθω≈(N+2)sin(θω+δθ). This number N rec of bounces is then given by:

Nrec2tanθωδθ2nδn.

This means that, after N rec≈150 bounces which corresponds typically to 15 mm long samples, the two beams can begin to overlap and destructive interferences are likely to occur. This might partly explain the discrepancy between experimental results and nonlinear plane wave theory. A more detailed analysis is now described.

 figure: Fig. 8.

Fig. 8. Recombination of the pump and SHG waves after a few bounces

Download Full Size | PDF

5. Guided wave multimode analysis of Fresnel Phase Matching

Different physical phenomena point to the inadequacy of a plane wave analysis for Fresnel phase matching. Indeed the spatial extensions of the waves are involved in two different aspects: (i) the non collinearity of the waves due to the nonlinear law of reflection, (ii) the Goos-Hänchen shift between the waves. Both effects induce a spatial walk off and interferences between the waves. It is thus natural to resort to a guided wave analysis which takes both phenomena into account in a built-in way. At first sight, it may seem that the guided wave approach is not useful because of the large number of modes which are involved in the wave propagation in the guide. Typically, the pump wavelengths are in the 0.6 µm to 10 µm range in the materials, while the sample thickness covers the 200 µm to 400 µm range, so that we are envisaging many hundred of modes. In fact, it is shown in Appendix B that only a small number of modes are actually involved (typically a few tens) under usual experimental conditions. For simplicity’s sake, and in order to validate this nonlinear guided wave theory with respect to the plane wave one, let us first describe the waveguide analysis of SHG process in an isotropic sample (DFG process is a mere generalization) for a simple case. In a second paragraph, we briefly describe the generalized modal numerical calculation used for modelling of the Fresnel phase matching experiments.

5.1 Analytical description of second harmonic generation process

We shall first recall the main points of this theory which are relevant to our situation. The waves are propagating in the z direction and the polarization vectors are oriented in the y direction for the s (TE) waves and in the (x,z) plane for the p (TM) one (see Fig. 9 for the geometry). In order to highlight the main physical insights of this theory, we first analytically present the simplest TEω-TE2ω or sss case.

 figure: Fig. 9.

Fig. 9. Fresnel phase-matching waveguide geometry.

Download Full Size | PDF

For such a configuration, the pump and SH electric fields distributions functions are respectively given by:

Eyω(x,z,t)=lAlω(z)Elω(x)ei(ωtβlωz)ey+cc,

and

Ey2ω(x,z,t)=mAm2ω(z)Em2ω(x)ei(2ωtβm2ωz)ey+cc.

In these expressions, E ωl (x)are the l th order mode TE wave eigenfunctions, solutions of the linear Maxwell equations. Their expressions can be found in references [2, 13]. The propagation constant βωl is solution of the implicit equation:

tan(αlωt)=2κlωαlω((αlω)2(κlω)2),

where t is the waveguide thickness. αlω=(nωkω)2(βlω)2 and κlω=(βlω)2(kω)2 are the transverse and the evanescent wavenumbers, respectively. In a ray tracing interpretation, it corresponds to a zigzag angle θωl given by sinθωl=βωl/(nωkω) (see Fig. 6). The Aωl(z) are the slowly varying amplitude of the modes. Their value at z=0 is obtained by the projection of the input Gaussian field on the waveguide modes and we note lθ the value of l for which |Aωl(0)| is maximum for a given incidence angle θ. As it is shown in Appendix B, few modes (llθ) are involved in this projection (from 10 to 20). Let us insist that, by construction, expressions (8) and (9) take into account, in a built-in way, the Goos–Hänchen shift as well as the Fresnel phase shift.

In the undepleted pump approximation (Aωl(z)≈Aωl(0)≈Aωl), the mode coupling equation which describes the energy transfer between the fundamental and the harmonic waves is easily derived (see Ref. [2]):

ddzAm,y2ω(z)=iωε0p0l'lχy,y,y(2)Sl,l',m(Al,yωAl',yω)eiδβl,l',mz;

p0 is a normalizing power constant which is such that the amplitude of the Poynting vector Π given by

=1μ0+E.Bdx=βlω2ωμ0+El,yω2dx=p0

corresponds to a power of 1 W/m in the y direction. S l,l’,m is the overlap integral between the modes:

Sl,l',m=+El,yω(x)El',yω(x)Em,y2ω(x)dx,

and δβ l,l’,m are the mismatch in the propagation constants:

δβl,l',m=βlω+βl'ωβm2ω.

Integration of Eq. (11) is trivial and leads to:

Am,x2ω(z)=ωε0p0χyyy(2)l'lSl,l',m(Al,yωAl',yω)L((exp(iδβl,l',mL)1)δβl,l',mL).

The normalization constant in Eq. (12) is such that the total amount of harmonic power generated by unit length along y axis is given by the summation over each mode contribution m:

P2ω(z)=mAm,y2ω(z)2.

A careful examination of Eqs. (15) and (16) shows that the energy transfer is efficient when three conditions are met.

(i) Firstly, as expected, a sinc (δβ l,l’,m L/2) term appears in the conversion Eq. (15). In order to obtain good conversion efficiency, TE modes at pulsations ω and 2ω must be almost phase matched, i.e.:

δβl,l',m=βlω+βl'ωβm2ω2kω(nωsinθlωn2ωsinθm2ω)0,

This latter approximation is valid since ll’. We can notice here that Eq. (17) is similar to Eq. (4) established in the frame of plane-wave theory. Moreover, it is important to highlight the fact that condition (17) is different in nature to the phase matching condition obtained in single mode waveguides [14]. For the latter, since single modes are involved, Eq. (17) is very stringent, as far as the waveguide thickness and the phase-matched wavelengths are concerned. This is why the tunability of nonlinear single mode waveguide is so restricted. However, in the former case, many sets of modes (l,l’,m) are involved in the parametric process so that Eq. (17) is far less stringent than in the single mode waveguides. We are in fact interested with the sets of modes with large coherence lengths Δl,l’,m=π/δβ l,l’,m, i.e. which satisfy the following inequality δβ l,l’,mπ/L or Δl,l’,mL. We will show that, in many situations, FPM can be obtained over a large span of wavelengths and sample thicknesses.

(ii) The element of the nonlinear susceptibility tensor (2) χyyy must not be zero. We must stress the fact that, in this guided wave approach, there is no more such consideration as the change of nonlinear coefficient at total internal reflection. This is taken into account by the value of the effective nonlinear coefficient which, in the general case, involves more than a single nonlinear susceptibility tensor element.

(iii) The overlap integral S l,l’,m between the modes must be optimized. As we will see below, this condition mostly points to the parity of the interacting waves.

In order to derive rules of thumbs which will allow us to further validate this nonlinear guided wave theory with respect to the plane wave one and derive approximate results which will be confirmed by comprehensive numerical calculations, we discuss the guided wave multimode analysis of FPM by use of expressions of the overlap integral, the phase mismatch and the mode amplitudes for highly or infinitely confining waveguides (HCW and ICW respectively).

One should note that, since the analytic expressions of mode spatial profiles Eωl (x) inside the nonlinear slab involve only trigonometric functions [2], analytic expressions of overlap integrals S l,l’,m can be derived in the general case as well. However, the corresponding expressions are complicated and their values do not differ significantly from those calculated in the ICW case. The expression of overlap integrals derived in this latter case are thus considered in the following discussion to gain some physical insights. It can be written as follows

Sl,l',m(1)m(1)l+l'(m+1)2Δ82π2tβ1ωβl'ωβm2ω(ωμ0p0)32,

where the mode mismatch l+l’-m=Δ is an odd integer (see Appendix C). The overlap integral is thus strongly enhanced for small Δ values (|Δ|=1) when ml+l’≈2 l, i.e. nωsinθ ωIn 2ω sinθ 2ω m, which is a rough expression of the momentum conservation at reflection. Larger Δ values correspond to higher non-collinearity angle between the waves and thus to lower spatial overlap between the interacting waves.

The analytical expression of mode amplitudes Aωl can also be derived for ICW (see Appendix B)

Alω=2iIωμ0βltp0π2wsinθeF(l)2[Erfi(F(l)isinθ2tw0)Erfi(F(l)+isinθ2tw0)],

where w0 is the incident Gaussian beam waist and F(l)is a function given by

F(l)=12w0sinθ(kωcosθlπt).

The expression of the mode amplitude (19) shows that the distribution of modes involved in the Gaussian beam propagation is centred near lθ≈ (cosθ)kωt/π, which is nothing else than the condition for waveguiding for an infinitely confining waveguide. Moreover, the number of excited modes Δl is roughly given by Δl=4tsinθ/(πw 0), meaning that only few modes (typically 10) are usually involved for typical experimental conditions (see Appendix B).

The ICW approximation is not suitable to derive useful analytical expressions of phase mismatch δβ l,l’,m because it overlooks the effect of the Fresnel phase lag at the waveguide interfaces. We thus evaluate δβ l,l’,m by use of a more sophisticated approach that relies on the Snyder–De la Rue approximation which is particularly valid in a HCW situation [15]. This yields (see Appendix D):

δβl,l',m=nωkωcosθtβlω[2nωcosθlnω21(1Kl+1Kl'1Km)ΔπtcosθlΛcπ],

where

βlω=(nωkω)2(lπt)2,

is the propagation constant of mode l in an ICW and

Kl={1forTEpolarizationnω2forTMpolarization

describes the TE and TM index contrast in a HCW, i.e. the effect of Fresnel birefringence.

Equation (21) is the main result of this calculation. It shows nicely how FPM is used to enhance tunability, the influence of mode mismatch, birefringence, and the wide angle tuning capability of the technique.

For the TEω-TE2ω case, the propagation constant mismatch is then :

δβl,l',mnωkωcosθtβlω(2nωcosθlnω21ΔπtπΛccosθl).

The first term in bracket represents the Fresnel phase shift, the second term is the modal dispersion and the last one is the phase mismatch between two bounces of the waves. The Fresnel phase matching conditions are fulfilled once, δβ l, l’,m≈0 :

2nωcosθlnω21tπΛccosθlΔπ,

with |Δ| an odd number. It is easily shown that such an equation is an alternative formula to the ray analysis one (3). The possible values of the mode mismatch Δ are given by the condition that Eq. (25) has a solution and the phase matched modes l are then the solutions of Eq. (25). We can also notice here that |Δ| represents the order of quasi-phase matching in the considered three wave mixing process.

Many different terms may appear in the summation of Eq. (15). However, because of the mode overlap dependence on Δ, terms where Δ is minimal are privileged. For any values of nω and Λc(ω), i.e., any frequency, and any large value of thickness t (tc), Eq. (25) possesses solutions, which explains the large tunability of Fresnel phase matching.

The solutions analytically derived in this section confirm the adequacy of the guided wave approach to our physical problem and their coherence with the results found under the ray analysis assumption. They are nevertheless approximations and a numerical analysis of the three waves mixing in the guided wave approach is required to take into account the different contributions summed in Eq. (15), for all polarization cases, and thus come up with calculated results to hold comparison with the experiment that has been presented in Sec. 3.

5.2 Description of the numerical modelling of SHG in a highly multimode guide.

The developed calculation code is a generalization for all states of polarizations of the three-wave mixing process analytically described previously.

Firstly, the eigenmodes fields Eωl(x) for TE modes and Hωl(x) for TM modes are calculated, as well as the propagation constants, at the pump and the SH wavelength, given the geometrical parameters considered in the calculations, such as: wavelengths, thickness, incidence angle, pump waist and polarization state.

The incident Gaussian pump waves are then numerically decomposed on these eigenmodes. In order to minimize the calculation duration, a filtering process is implemented to keep only the waveguide modes that contain a significant amount of the incident power (around 10–20 modes contain the total incident power for incident waists in the 100 µm range and 400 µm thick plates).

Let us recall here that [2, 13]:

ETEω(x,z,t)=lAlTE,ω(z)Elω(x)ei(ωtβlTE,ωz)ey+cc

for a TE wave and,

HTMω(x,z,t)=lKlω(z)Hlω(x)ei(ωtβlTM,ωz)ey+cc,

so that

ETMω(x,z,t)=l(Al,xTM,ω(z)El,xTM,ω(x)ex+Al,zTM,ω(z)El,zTM,ω(x)ez)ei(ωtβlTM,ωz)+cc

for a TM wave.

In a next step, the coefficients of the eigenmodes of the generated TE and TM second harmonic waves are calculated, under the approximation of undepleted pump. The coupling equations to calculate each mode coefficient, which are generalization of Eq. (15), are derived from Maxwell equations and presented in reference [2].

ddzAm,yTE,2ω(z)=iω4p0PxNL(ω,x,z)EmTE,2ω(x)dx

for a TE SH wave,

ddzKp,yTM,2ω(z)=14ε0n2ω2[ddzPxNL(ω,x,z)ddzPzNL(ω,x,z)]HpTM,2ω(x)dx

for a TM SH wave,

with P NL the nonlinear driving polarization which involves the fundamental wave.

For each case (TE or TM SH wave), coordinates changes have to be operated to calculate the nonlinear polarization in the crystallographic coordinates of the crystal. Indeed, in order to explore our experimental conditions, the mixing in a <110> oriented GaAs plate (as described in Fig. 10) is considered. The overlap integrals S l,l’,m also involved in the nonlinear polarization term are calculated by use of exact analytical expressions to minimize the calculation time. For each polarization state, we can then derive the coefficients of each SH eigenmode fields [2], and thus obtain the SH power as a function of the guide length as well as the spatial profile of the SH wave propagating along the guide.

6. Numerical modelling results and comparison with experiments

6.1 Numerical results for a 4 µm→2 µm SHG process

In order to hold comparisons with the experiments presented in Sec. 3, the numerical calculations presented here are based on the 4 µm→2 µm SHG process.

As a first step before carrying out nonlinear mixing calculation, we check that the pump wave propagation is properly described using our calculation. In this purpose, we consider the propagation of a TM polarized input Gaussian beam with a 100 µm waist into a 400 µm thick gallium arsenide slab for two different incidence angles: either along the longitudinal direction the plate (90° injection angle) or injected at a 32.5° angle.

Figure 10 shows the interference pattern between the excited waveguide modes along the slab which fully describes the propagation of the Gaussian beam in both cases. Indeed, as shown in Fig. 10(a) for an on-axis propagation, the natural diffraction of the Gaussian beam is accurately rendered as well as the zigzag propagation shown in Fig. 10(b) where one can notice the interference fringes occurring at each reflection in the overlap of upward and downward beams. Besides, as shown in Fig. 10(c), it is confirmed that the incident pump power is actually coupled to only few modes as expected with the analytical treatment in the ICW case presented in Sec. 5.1.

 figure: Fig. 10.

Fig. 10. Calculation of the pump wave intensity, (a) along the longitudinal direction of the guide, (b) zigzag propagation, (c) modal distribution (intensity) of the incoming pump for the zigzag case geometry.

Download Full Size | PDF

As a second step, the consistency of the SHG model is validated in the straightforward case of a pump beam propagating along the plate axis where the expected result is well-known. For a TM fundamental beam at 4 µm, we calculate the generated beam for both TE and TM polarizations. Figure 11(a) shows the generated TE beam at 2 µm. As expected in the case of phase mismatched SHG, the SH beam intensity displays a sinusoidal modulation along its propagation axis. The corresponding SHG coherence length of 28 µm corroborates the one calculated by use of the well-known textbook plane wave formula: Λc=λ/[4(n 2ω-nω)]. For the TM case presented in Fig. 11(b), the SH intensity profile exhibits the same periodicity along the propagation axis. But it also displays a modulation along the transverse axis where it cancels out along the longitudinal wafer axis. We also notice that in this case, the maximal SH intensity is orders of magnitude smaller than for the TE case. These atypical features are actually fully explained by the fact that the effective nonlinear coefficient cancels out for θ=90° (d eff=d 143cosθcosϕ(-sin2 θ+cos2 θsin2 ϕ), according to the angular notations of Fig. 9).

 figure: Fig. 11.

Fig. 11. Calculation of the SHG TE (a) and TM (b) waves generated with a TM pump wave propagating along a 300 µm guide axis (n.b. the TM intensity is 7 orders of magnitude lower than the TE one).

Download Full Size | PDF

Then, we calculate the Fresnel phase matched SHG process that corresponds to the experimental configuration investigated in Sec. 4.2, i.e. a 4-µm TM-polarized fundamental beam generating a 2-µm TE-polarized SH beam at an internal propagation angle of 32.5° in a 400-µm thick GaAs slab. The spatial profile of the calculated SH beam is shown in Fig. 12 in the case of a 4-mm long GaAs sample. As expected, one can observe: (i) a zigzag propagation of the generated beam at the right angle, (ii) a periodic modulation of the SH intensity along each zigzag leg that is consistent with the coherence length of our high-order (precisely 17) quasi-phase matching process, (iii) a step-growth of the SH intensity at each bounce inside the plate, meaning that the Fresnel phase matching condition is satisfied.

 figure: Fig. 12.

Fig. 12. Calculation of a TE SHG wave for a TM pump wave injected at a 32.5° angle inside a 400 µm thick 4 mm long guide.

Download Full Size | PDF

Similar calculations are also run for longer crystals in order to evaluate the SHG signal as a function of the sample length. The obtained results are discussed hereafter in Sec. 6.2 where they are compared to experiments on actual GaAs plates.

6.2 Experimental results and comparison

Figure 13 shows the results of this model applied to our experimental case, with a 30 mm long GaAs sample. Clearly the non collinear SHG wave completely fills the guide along its generation.

 figure: Fig. 13.

Fig. 13. Calculation of a TE SHG wave for a TM pump wave injected at a 32.5° angle inside a 400 µm thick 30 mm long guide.

Download Full Size | PDF

Figure 14 shows the calculation of the SHG efficiency as a function of the crystal length, with and without the 0.5% loss per bounces that were experimentally measured at 4 µm. The calculation clearly show that the conversion efficiency does not grow quadratically with the crystal length above 10 mm and even display a saturation around 20 mm. The agreement with the experimental result is rather fair and shows a strong amelioration compared to the plane wave approach.

This confirms the relevance of the guided-wave approach and demonstrates that the saturation of the nonlinear efficiency in the considered FPM scheme can be satisfactorily accounted by such approach that includes the main linear and nonlinear propagation phenomena limiting the nonlinear conversion.

 figure: Fig. 14.

Fig. 14. Comparison of the measured and calculated SHG intensity as a function of the crystal length. Note that the point dispersion on the waveguide calculation line is an aliasing effect.

Download Full Size | PDF

7. Conclusion

New theoretical concepts are presented, which allow a better understanding of Fresnel phase matching. A guided wave approach is described, which allows us to intrinsically take into account all the physical processes involved, in particular limitations such as walk-off effects, due to Goos-Hänchen shift and nonlinear reflection at the interfaces. Our model was validated by second harmonic generation experiments in the mid-infrared. The guided wave description presented here thus allows us to model and generalize this phase matching technique to any nonlinear material plate thickness, from nearly single mode to highly multimode waveguides, to any optical wavelengths. Moreover, this model can be applied to the optimization of Fresnel birefringence phase matched plates for the infrared, allowing optimum structures to be designed and tested in the near future. We also envision to apply our calculation to the generation of terahertz waves by difference frequency generation of close infrared lines. Our model would be there an essential tool, since the semiconductor would behave as a strongly multimode guide for the pumps and a nearly single mode guide for the terahertz wave.

Appendix A : Reciprocal lattice vector for Fresnel phase matching

Let three waves of amplitude Ei(r)ei(ωitki·r) interact in a nonlinear materials in a Fresnel phase matching configuration. We shall treat the case ω 1-ω 2ω 3. The generalization to any kind of second order process (SFG, SHG) is trivial. The geometry is shown in Fig. 7: pump beams ω 1 and ω 2 propagate in the (x,z) plane at an angle of incidence θ-δθ and we anticipate a reflection angle of θ for the difference-frequency wave. The z axis is chosen to be collinear to the difference-frequency wave vector k 3. The phase of the difference-frequency must thus be constant along x axis to be a plane wave.

The nonlinear conversion differential equation thus reads:

dE3dz=iω3n3cχeff2E1E2*ei(Δkz+Δkx),

with Δk =(k 1-k 2)cosδθ-k 1 and Δk =(k 1-k 2)sinδθ.

In the coordinates (O, x, z), the boundary of domain number is given by the equation (see Fig. 7): z=-xtanθ+ℓt/cosθ

The difference-frequency field increment along domain number is thus given by the following integral:

E3(z+1)ei(+1)ϕ3=E3(z)eiϕ3eiΔϕFzz+1iω3n3cχeff2E1E2*ei(Δkz+iΔkx)dz,

with ΔϕF=ϕ 1-ϕ 2-ϕ 3+δϕ NL.

In order to conveniently integrate the nonlinear equation, we introduce the variable changing z′=z+xtanθ+ℓt/cosθ so that the integral can be rewritten:

E3(z+1)exp(i(+1)ϕ3)=E3(z)exp(iϕ3)exp(iΔϕF)0tcosθiω3n3cχeff2E1E2*exp(i(Δkz'+(ΔkΔktanθ)x+Δktcosθ))dz.

Since plane-wave solutions are considered, the difference-frequency field increment must not depend on coordinate x, which leads to the following condition:

Δktanθ+Δk=0.

After integration, Eq. (A3) thus yields :

E3(z+1)exp(i(+1)ϕ3)=E3(z)exp(iϕ3)
iω32n3ctcosθsin⁡c(Δkt2cosθ)exp(iΔkt2cosθ)χeff2E1E2*exp(i(Δktcosθ+ΔϕF)).

And the difference-frequency field after propagating through N domains reads:

E3(zN)exp(iNϕ3)=iω32n3ctcosθsin⁡c(Δkt2cosθ)exp(iΔkt2cosθ)χeff(2)E1E2*
×l=0N1exp(i(Δktcosθ+ΔϕF))

One thus obtains :

E3(N)2=(ω32n3ctcosθχeff(2))2E12E22sinc2(Δkt2cosθ)sin2[N2(Δktcosθ+ΔϕF)]sin2[12(Δktcosθ+ΔϕF)].

In this way, one recovers the Fresnel quasi-phase matching condition :

Δk+ΔϕFtcosθ=0.

Inserting the above equation in condition (A4), (A4) can be rewritten as

Δk+ΔϕFtsinθ=0.

It is thus possible to interpret Fresnel quasi-phase matching with the following vector conservation law :

k1+KF=k2+k3,

where |K F|=|ΔϕF|/t is the reciprocal lattice vector for Fresnel phase matching.

Appendix B : Mode amplitudes of a Gaussian beam in the infinitely confining waveguide (ICW) approximation

Let us consider a Gaussian beam propagating in an ICW at an angle of incidence θ and focused at z=0. We neglect any effects related to the beam inclination relatively to the entrance facet. The waist of the beam is w 0 at z=0 (see Fig. 6). The Gaussian beam at z=0 is thus described by:

Einc(x)=Iexp((xw0sinθ)2ikωxcosθ),

where I is the intensity of the beam. The projection of the Gaussian beam onto the eigenmodes El,yω(x)=2ωμ0βltp0sc(lπtx) is given by:

Alω=2ωμ0βltp0It2t2exp((xw0sinθ)2ikωxcosθ)exp(ilπtx)dx.

In fact, Eq. (B2) is a short expression from which the projection on (i) even modes (l odd and sc=cos) is the real part of (B2) and (ii) on odd modes modes (l even and sc=sin) is the imaginary part. This expression is easily calculated and leads to :

Alω=2iIωμ0βltp0π2wsinθexp(F(l)2)
×[Erfi(F(l)isinθ2tw0)Erfi(F(l)+isinθ2tw0)]

where

F(l)=12w0sinθ(kwcosθlπt),

and Erfi(z)=-iErf(iz), Erf(z) being the error function.

From expression, (B3), it appears that the mode amplitude is maximal for F(l)≈0 i.e. for kω cosθ/t which is nothing else than the condition for wave guiding for an infinitely confining waveguide. Moreover, the e -1 FWHM of the AωI distribution in l is given by the condition:

F(l)=12w0sinθ(kωcosθlπt)=±1,

so that the number of modes involved in the Gaussian beam propagation is roughly given by:

Δl=4πtw0sinθ.

This latter result shows that, indeed only few modes are expected to participate to the Gaussian beam propagation and nonlinear process. For example, in a typical experimental case where θ~30°, t~400 µm, w 0~100 µm, Δl ≈3. We can thus expect in such cases that around 10-20 modes should be sufficient to describe the injected waves, as confirmed by the numerical decomposition (Fig. 7(c)). Moreover, the smaller the Gaussian beam waist w0, the higher the diffraction effects and the number of modes necessary to describe the beam propagation.

Appendix C : Overlap integral in an infinitely confining waveguide

As far as the overlap integrals are concerned, the phase between the waves are less important and the infinitely confining waveguide approximation is fair enough. The eigenmode electric field distributions for TE waves are given by :

El,yω(x)=2ωμ0βltp0sc(lπtx).

where sc is the sine function if l is even (odd modes) and cosine if l is odd (even modes). The overlap integral for a SHG process is thus given by:

Sl,l',m=82βlωβl'ωβm2ω(ωμ0tp0)32t2t2sc(lπtx)sc(l'πtx)sc(mπtx)dx.

A careful examination of Eq. (C2) shows that the overlap integral is non zero only if l+l’-m=Δ is an odd integer. In that case, for large values of mode number l and m, the overlap integral may be approximated by:

Sl,l',m(1)m(1)l+l'(m+1)2l+l'm82π2tβlωβl'ωβm2ω(ωμ0p0)32.

Appendix D : Phase mismatch in a highly confining waveguide

Equation (15) shows that the energy flows from the mode l to the mode m proportionally to the phase mismatch (PM) term |sinc(δβ l,l’,m L/2)2. Let us evaluate the wave-vector mismatch δβ l,l’,m for a highly confining waveguide (HCW). We will use the Snyder-De La Rue approximations which are particularly valid in a HCW situation [14].

The field distributions display a sin(αlx) or cos(αlx) functional form inside the waveguide, depending on whether the modes are odd or even, with αI related to the propagation constants βl by α 2 l+β 2 l=nω 2 kω 2. The even (respectively odd) modes correspond to odd (respectively even) values of integer number l or m. For instance, the TE0 correspond to l=1 and the field distribution is a cos one. In the Snyder-De La Rue approximation, the coefficients αωI are given by:

αlωlπtexp[1(KlVω)],

with

Vω=kωt2nω21

and

Kl={1forTEpolarizationnω2forTMpolarization

The propagation constants for the(ω,l) and (2ω,m) modes are then respectively:

βlω=(nωkω)2(lπt)2exp(2KlVω)

and

βm2ω=(2n2ωkω)2(mπt)2exp(2KmV2ω)

Noting that V is very large in these highly multimodal waveguides (t large), expression (D4) can be expended to:

βlωβlω+1KlVωβlω(lπt)2

where βlω=(nωkω)2(lπt)2 is the propagation constant in an infinitely confining waveguide (V=∞). We introduce now the mode mismatch number Δ=l+l’-m (Δ≪l,l’) and δn=n 2ω-n ω. Considerations on the overlap integrals show that Δ is an odd number (see Appendix C). Expression D5 can be expended to:

βm2ω=2βlω+2βlω[nωδnkω2+l2lm2(πt)2+1KmV2ω(lπt)2].

We may now neglect the dispersion term in Vω (i.e. V 2ω≈2 Vω), δ n being a second order term. We can now evaluate the mode phase mismatch:

δβl,l',m=βlω+βl'ωβm2ω(πt)2βlω[l2Vω(1Kl+1Kl'1Km)lΔ2nωδn(tkωπ)2].

Reintroducing the definitions αωl/tnωkωcosθl and Λc=π21δnkω, Eq. (D8) may be cast in the following form:

δβl,l',m=nωkωcosθltβlω[2nωcosθlnω21(1Kl+1Kl'1Km)ΔπtcosθlΛcπ].

A good conversion efficiency for the sets of modes (l,l’,m) is obtained once modal phase matching is obtained i.e.

2nωcosθlnω21(1Kl+1Kl'1Km)ΔπtcosθlΛcπ0.

This latter expression is particularly illuminating. It shows the different dispersion mechanisms which are at stake in the wafer: the first term represents the Fresnel phase lag between the TE and the TM mode (ΔϕF in expression (2) or (6)), the second one is the modal dispersion while the last one is the optical dispersion contribution (ΔkLb in expressions (2) or (6)). As it can be seen, optical dispersion may be compensated by the modal and the Fresnel ones. Equation (D10) is the main result of this appendix which leads to the following conclusions:

One can notice that the mode mismatch N 0=|Δ| represents what was called the phase matching order in the plane wave analysis.

One notes that there is a correction to the phase matching condition t=ΔΛccosθl due to the Fresnel phase shift, that is more important in the TE-TM configuration than in a TE-TE or TM-TM configuration.

Finally, one notes that the higher the phase matching order N 0, the higher Δ, which leads to a decrease of the overlap integral (C3) and to a decrease of the linear conversion in the waveguide, which is not a surprise.

References and links

1. A. Fiore, V. Berger, E. Rosencher, P. Bravetti, and J. Nagle, “Phase matching using an isotropic nonlinear optical material,” Nature 391, 463–466 (1998). [CrossRef]  

2. A. Yariv, “Coupled-mode Theory for Guided-Wave Optics,” IEEE J. of Quant. Electron. QE-9, 919–933 (1973). [CrossRef]  

3. K.L. Vodopyanov, O. Levi, P. S. Kuo, T. J. Pinguet, J. S. Harris, M. M. Fejer, B. Gerard, L. Becouarn, and E. Lallier, “Optical parametric oscillation in quasi-phase-matched GaAs,” Opt. Lett. 29, 1912–1914 (2004). [CrossRef]   [PubMed]  

4. J.A. Armstrong, N. Bloembergen, J. Ducuing, and P.S. Pershan, “Interactions between light waves in a nonlinear dielectric,” Phys. Rev. 127, 1918–1939 (1962). [CrossRef]  

5. G. Boyd and C. Patel, “Enhancement of optical second-harmonic generation by reflection phase matching in ZnSe and GaAs,” Appl. Phys. Lett. 8, 456–459 (1966). [CrossRef]  

6. H. Komine, W.H. Long, J.W. Tully, and E.A. Stappaerts, “Quasi-phase matched second-harmonic generation by use of a total internal-reflection phase shift in gallium arsenide and zinc selenide plates,” Opt. Lett. 23, 661–663 (1998). [CrossRef]  

7. R. Haidar, N. Forget, P. Kupececk, and E. Rosencher, “Fresnel phase matching for three-wave mixing in isotropic semiconductors,” J. Opt. Soc. Am. B 21, 1522 (2004). [CrossRef]  

8. N. Bloembergen and P.S. Pershan, “Light waves at the boundary of nonlinear media,” Phys. Rev. 128, 606–622 (1962). [CrossRef]  

9. M. Born and E. Wolf, in Principles of Optics, Pergamon Press (1964).

10. M. Raybaut, A. Godard, A. Toulouse, C. Lubin, and E. Rosencher, “Non linear effects on Fresnel phase matching,” Appl. Phys. Lett. 92, 121112 (2008). [CrossRef]  

11. M.M. Fejer, G.A. Magel, D.H. Jundt, and R.L. Byer, “Quasi-Phase-Matched Second Harmonic Generation: Tuning and Tolerances,” IEEE J. Quantum Electron. 28, 2631–3654 (1992). [CrossRef]  

12. A.V. Smith, D.J. Armstrong, and W.J. Alford, “Increased acceptance bandwidths in optical frequency conversion by use of multiple walk-off-compensating nonlinear crystals,” J. Opt. Soc. Am. B 15, 122 (1998). [CrossRef]  

13. D. Marcuse, in Theory of dielectric optical waveguides (Academic press, London, 1974).

14. A. Fiore, V. Berger, E. Rosencher, P. Bravetti, N. Laurent, and J. Nagle, “Phase matched mid-infrared difference frequency generation in GaAs-based waveguides,” Appl. Phys. Lett. 71, 3622–3624 (1997). [CrossRef]  

15. A. W. Snyder and R. De La Rue, “Asymptotic solution of eigenvalue equations for surface waveguide structure,” IEEE Trans. Microwave Theory Tech. 18, 650–651 (1970). [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 (14)

Fig. 1.
Fig. 1. Scheme of the Fresnel phase-matching configuration for SHG and DFG: geometry of the Fresnel phase-matched plate: (a) side view for geometrical description; and (b) view from above for angle definition.
Fig. 2.
Fig. 2. Resonant Fresnel phase-matching: the distance Lb between two successive bounces is optimized (Lb =(2p+1) Λ c , where p is an integer and Λ c the coherence length of the nonlinear interaction).
Fig. 3.
Fig. 3. Experimental set-up: (a) nonlinear material samples geometry, the only varying parameter is the length L, (b) SHG experimental setup, (c) single coherence length intensity calibration experiment.
Fig. 4.
Fig. 4. Experimental results of 4→2 µm FPM SHG experiment: SH intensity as a function of the internal angle (a), SHG intensity normalized to one coherence length intensity as a function of the sample length (b).
Fig. 5.
Fig. 5. Spatial profiles of the pump beam (using an infrared Pyrocam camera) (a) and SHG beam (using a SWIR camera) (b), at the output of the GaAs crystal. Note that the profiles are only given here for a qualitative appreciation of the beam quality, not to give an actual value of the beams sizes.
Fig. 6.
Fig. 6. Nonlinear total internal reflection law
Fig. 7.
Fig. 7. Unfolding of the zig-zag motion of the waves at total internal reflection (a). Because of the nonlinear law of reflection, the waves suffer from a non collinear propagation due to the periodicity of the structure (b).
Fig. 8.
Fig. 8. Recombination of the pump and SHG waves after a few bounces
Fig. 9.
Fig. 9. Fresnel phase-matching waveguide geometry.
Fig. 10.
Fig. 10. Calculation of the pump wave intensity, (a) along the longitudinal direction of the guide, (b) zigzag propagation, (c) modal distribution (intensity) of the incoming pump for the zigzag case geometry.
Fig. 11.
Fig. 11. Calculation of the SHG TE (a) and TM (b) waves generated with a TM pump wave propagating along a 300 µm guide axis (n.b. the TM intensity is 7 orders of magnitude lower than the TE one).
Fig. 12.
Fig. 12. Calculation of a TE SHG wave for a TM pump wave injected at a 32.5° angle inside a 400 µm thick 4 mm long guide.
Fig. 13.
Fig. 13. Calculation of a TE SHG wave for a TM pump wave injected at a 32.5° angle inside a 400 µm thick 30 mm long guide.
Fig. 14.
Fig. 14. Comparison of the measured and calculated SHG intensity as a function of the crystal length. Note that the point dispersion on the waveguide calculation line is an aliasing effect.

Equations (64)

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

Δ ϕ = Δ k L b + Δ ϕ F + δ ϕ .
I 3 out = Z 0 ω 3 2 2 c 2 ( N d eff ) 2 n 1 n 2 n 3 ( sin ( Δ k L b 2 ) sin ( Δ k 2 ) ) 2 ( sin ( N Δ ϕ 2 ) N sin ( Δ ϕ 2 ) ) 2 I 1 in I 2 in ,
Δ ϕ ( θ ) = Δ k L b ( θ ) + Δ ϕ F ( θ ) + δ ϕ = 0 [ 2 π ] .
n ω sin θ ω = n 2 ω sin θ 2 ω .
δ θ ( δ n n ) tan θ ω ,
k 1 + K F = k 2 + k 3
K F = Δ ϕ F t ,
N rec 2 tan θ ω δ θ 2 n δ n .
E y ω ( x , z , t ) = l A l ω ( z ) E l ω ( x ) e i ( ω t β l ω z ) e y + c c ,
E y 2 ω ( x , z , t ) = m A m 2 ω ( z ) E m 2 ω ( x ) e i ( 2 ω t β m 2 ω z ) e y + c c .
tan ( α l ω t ) = 2 κ l ω α l ω ( ( α l ω ) 2 ( κ l ω ) 2 ) ,
d d z A m , y 2 ω ( z ) = i ω ε 0 p 0 l ' l χ y , y , y ( 2 ) S l , l ' , m ( A l , y ω A l ' , y ω ) e i δ β l , l ' , m z ;
= 1 μ 0 + E . B d x = β l ω 2 ω μ 0 + E l , y ω 2 d x = p 0
S l , l ' , m = + E l , y ω ( x ) E l ' , y ω ( x ) E m , y 2 ω ( x ) d x ,
δ β l , l ' , m = β l ω + β l ' ω β m 2 ω .
A m , x 2 ω ( z ) = ω ε 0 p 0 χ y y y ( 2 ) l ' l S l , l ' , m ( A l , y ω A l ' , y ω ) L ( ( exp ( i δ β l , l ' , m L ) 1 ) δ β l , l ' , m L ) .
P 2 ω ( z ) = m A m , y 2 ω ( z ) 2 .
δ β l , l ' , m = β l ω + β l ' ω β m 2 ω 2 k ω ( n ω sin θ l ω n 2 ω sin θ m 2 ω ) 0 ,
S l , l ' , m ( 1 ) m ( 1 ) l + l ' ( m + 1 ) 2 Δ 8 2 π 2 t β 1 ω β l ' ω β m 2 ω ( ω μ 0 p 0 ) 3 2 ,
A l ω = 2 i I ω μ 0 β l t p 0 π 2 w sin θ e F ( l ) 2 [ Erfi ( F ( l ) i sin θ 2 t w 0 ) Erfi ( F ( l ) + i sin θ 2 t w 0 ) ] ,
F ( l ) = 1 2 w 0 sin θ ( k ω cos θ l π t ) .
δ β l , l ' , m = n ω k ω cos θ t β l ω [ 2 n ω cos θ l n ω 2 1 ( 1 K l + 1 K l ' 1 K m ) Δ π t cos θ l Λ c π ] ,
β l ω = ( n ω k ω ) 2 ( l π t ) 2 ,
K l = { 1 for TE polarization n ω 2 for TM polarization
δ β l , l ' , m n ω k ω cos θ t β l ω ( 2 n ω cos θ l n ω 2 1 Δ π t π Λ c cos θ l ) .
2 n ω cos θ l n ω 2 1 t π Λ c cos θ l Δ π ,
E TE ω ( x , z , t ) = l A l TE , ω ( z ) E l ω ( x ) e i ( ω t β l TE , ω z ) e y + c c
H TM ω ( x , z , t ) = l K l ω ( z ) H l ω ( x ) e i ( ω t β l TM , ω z ) e y + c c ,
E TM ω ( x , z , t ) = l ( A l , x TM , ω ( z ) E l , x TM , ω ( x ) e x + A l , z TM , ω ( z ) E l , z TM , ω ( x ) e z ) e i ( ω t β l TM , ω z ) + c c
d d z A m , y TE , 2 ω ( z ) = i ω 4 p 0 P x NL ( ω , x , z ) E m TE , 2 ω ( x ) d x
d d z K p , y TM , 2 ω ( z ) = 1 4 ε 0 n 2 ω 2 [ d d z P x NL ( ω , x , z ) d d z P z NL ( ω , x , z ) ] H p TM , 2 ω ( x ) d x
d E 3 d z = i ω 3 n 3 c χ eff 2 E 1 E 2 * e i ( Δ k z + Δ k x ) ,
E 3 ( z + 1 ) e i ( + 1 ) ϕ 3 = E 3 ( z ) e i ϕ 3 e i Δ ϕ F z z + 1 i ω 3 n 3 c χ eff 2 E 1 E 2 * e i ( Δ k z + i Δ k x ) d z ,
E 3 ( z + 1 ) exp ( i ( + 1 ) ϕ 3 ) = E 3 ( z ) exp ( i ϕ 3 )
exp ( i Δ ϕ F ) 0 t cos θ i ω 3 n 3 c χ eff 2 E 1 E 2 * exp ( i ( Δ k z ' + ( Δ k Δ k tan θ ) x + Δ k t cos θ ) ) d z .
Δ k tan θ + Δ k = 0 .
E 3 ( z + 1 ) exp ( i ( + 1 ) ϕ 3 ) = E 3 ( z ) exp ( i ϕ 3 )
i ω 3 2 n 3 c t cos θ sin⁡ c ( Δ k t 2 cos θ ) exp ( i Δ k t 2 cos θ ) χ eff 2 E 1 E 2 * exp ( i ( Δ k t cos θ + Δ ϕ F ) ) .
E 3 ( z N ) exp ( i N ϕ 3 ) = i ω 3 2 n 3 c t cos θ sin⁡ c ( Δ k t 2 cos θ ) exp ( i Δ k t 2 cos θ ) χ eff ( 2 ) E 1 E 2 *
× l = 0 N 1 exp ( i ( Δ k t cos θ + Δ ϕ F ) )
E 3 ( N ) 2 = ( ω 3 2 n 3 c t cos θ χ eff ( 2 ) ) 2 E 1 2 E 2 2 sin c 2 ( Δ k t 2 cos θ ) sin 2 [ N 2 ( Δ k t cos θ + Δ ϕ F ) ] sin 2 [ 1 2 ( Δ k t cos θ + Δ ϕ F ) ] .
Δ k + Δ ϕ F t cos θ = 0 .
Δ k + Δ ϕ F t sin θ = 0 .
k 1 + K F = k 2 + k 3 ,
E inc ( x ) = I exp ( ( x w 0 sin θ ) 2 i k ω x cos θ ) ,
A l ω = 2 ω μ 0 β l t p 0 I t 2 t 2 exp ( ( x w 0 sin θ ) 2 i k ω x cos θ ) exp ( i l π t x ) d x .
A l ω = 2 i I ω μ 0 β l t p 0 π 2 w sin θ exp ( F ( l ) 2 )
× [ Erfi ( F ( l ) i sin θ 2 t w 0 ) Erfi ( F ( l ) + i sin θ 2 t w 0 ) ]
F ( l ) = 1 2 w 0 sin θ ( k w cos θ l π t ) ,
F ( l ) = 1 2 w 0 sin θ ( k ω cos θ l π t ) = ± 1 ,
Δ l = 4 π t w 0 sin θ .
E l , y ω ( x ) = 2 ω μ 0 β l t p 0 sc ( l π t x ) .
S l , l ' , m = 8 2 β l ω β l ' ω β m 2 ω ( ω μ 0 t p 0 ) 3 2 t 2 t 2 sc ( l π t x ) sc ( l ' π t x ) sc ( m π t x ) d x .
S l , l ' , m ( 1 ) m ( 1 ) l + l ' ( m + 1 ) 2 l + l ' m 8 2 π 2 t β l ω β l ' ω β m 2 ω ( ω μ 0 p 0 ) 3 2 .
α l ω l π t exp [ 1 ( K l V ω ) ] ,
V ω = k ω t 2 n ω 2 1
K l = { 1 for TE polarization n ω 2 for TM polarization
β l ω = ( n ω k ω ) 2 ( l π t ) 2 exp ( 2 K l V ω )
β m 2 ω = ( 2 n 2 ω k ω ) 2 ( m π t ) 2 exp ( 2 K m V 2 ω )
β l ω β l ω + 1 K l V ω β l ω ( l π t ) 2
β m 2 ω = 2 β l ω + 2 β l ω [ n ω δ n k ω 2 + l 2 l m 2 ( π t ) 2 + 1 K m V 2 ω ( l π t ) 2 ] .
δ β l , l ' , m = β l ω + β l ' ω β m 2 ω ( π t ) 2 β l ω [ l 2 V ω ( 1 K l + 1 K l ' 1 K m ) l Δ 2 n ω δ n ( t k ω π ) 2 ] .
δ β l , l ' , m = n ω k ω cos θ l t β l ω [ 2 n ω cos θ l n ω 2 1 ( 1 K l + 1 K l ' 1 K m ) Δ π t cos θ l Λ c π ] .
2 n ω cos θ l n ω 2 1 ( 1 K l + 1 K l ' 1 K m ) Δ π t cos θ l Λ c π 0 .
Select as filters


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