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

Spatiotemporal modulation instability as off-axis parametric amplification: insights from the phase

Open Access Open Access

Abstract

The canonical linear-stability analysis of spatiotemporal modulation instability (MI), which treats MIs as instabilities of the steady-state solution of the field evolution equation with respect to weak complex harmonic perturbations, is shown to be physically equivalent to a coupled-wave analysis of off-axis parametric amplification of the Stokes and anti-Stokes fields by an intense pump wave. With an appropriate space-evolution transfer matrix, a complex harmonic trial function used in the standard MI model translates into a pair of coupled off-axis waves, of which one is exponentially growing, while the other is exponentially decreasing, thus recovering a two-wave field structure that is inherent in the fields undergoing parametric amplification through four-wave mixing. Analysis of the phase of these fields offers useful physical insights into high-order dispersion effects in spatiotemporal MIs, suggests a physically transparent and accurate method to include high-order dispersion in MI gain calculations, and reveals new spatiotemporal MI effects induced by high-order dispersion.

© 2016 Optical Society of America

1. Introduction

Optical nonlinearity sets fundamental limits on the peak power that an intense light field can accommodate without losing its connectedness in space and time through the exponential growth of modulation instabilities (MIs) across the beam and along the pulse. In the time domain, nonlinearity and dispersion make a high-peak-power light pulse unstable with respect to small modulations of its temporal envelope, which tend to split the pulse into multiple peaks and give rise to spectral sideband generation [1]. As their manifestation in spatial nonlinear dynamics, MIs cause a light beam with a peak power above the critical power of self-focusing to develop a complex transverse mode structure [2], eventually breaking up into multiple filaments [3,4].

In optical fibers, temporal MIs play a central role in parametric frequency conversion and soliton pulse transformation scenarios [1]. Since the peak powers needed for temporal MIs to take off in optical fibers are typically much lower than the critical power of self-focusing, manifested as nonlinearity-induced mode coupling, such MIs can often be treated separately from their spatial counterpart. Spatial MIs, on the other hand, are currently the focus of much interest in the context of multiple filamentation [3,4], which sets limits on the peak powers, and, hence, radiation energies, that can be transmitted or, with properly chosen parameters, compressed in a single filament, that is, without compromising the output beam quality. Although solutions for spatial MIs coupled to temporal MIs have been known for decades [5], spatial MIs of self-focusing beams and laser filaments are often examined separately from temporal MIs. This tendency owes much to the physically transparent picture of spatial MIs and simple, closed-form solutions for the MI gain provided in the seminal work by Bespalov and Talanov [2]. As filamentation-based approaches find growing applications in remote sensing [6] and high-power pulse compression [7], expanding into new wavelength ranges [8], helping generate high-power subcycle pulses [9], and offering new solutions for standoff detection using filamentation-driven lasing [10], a more detailed understanding of multiple filamentation regimes is needed [11] with a special emphasis on effects that go beyond the standard treatment of MIs, based on the slowly varying envelope approximation (SVEA).

Here, we show that useful insights into the properties of spatiotemporal MIs can be gained from the analysis of noncollinear four-wave mixing (FWM) leading to a coupled parametric amplification of MIs in space and time. We demonstrate that, with the MI gain and sideband frequencies expressed through the wave-vector mismatch for such an FWM process, the main results of the canonical SVEA treatment of spatiotemporal MIs can be fully recovered. Moreover, such analysis will be shown to offer important physical insights into non-SVEA effects in spatiotemporal MIs.

2. Linear stability analysis of modulation instabilities

Within the framework of the canonical linear-stability analysis [1,12], the gain and sideband frequencies of spatiotemporal MIs are derived by considering a weak plane-wave perturbation [5]

ε=[ε1cos(KrΩτ)+iε2sin(KrΩτ)]exp(iγI0z)
of the steady-state field E(r,t)=(1/2)A(r,t)exp[i(k0zω0t)]+c.c. with an envelope A(r,t)=A0exp(iγI0z) found from the SVEA field evolution equation
i(Az+1uAt)+12k0(2Ax2+2Ay2)k222At2+γ|A|2A=0
Here, I0=A02, k0=ω0n0/c, n0 = n(ω0) is the refractive index at the frequency ω0, n2 is the nonlinear refractive index, γ=ω0n2/c, k2 is the coefficient in the power-series expansion of the wave number k(ω)m=0(km/m!)(ωω0)mabout the frequency ω0, u = k11 is the group velocity, x and y are the transverse coordinates, z is the longitudinal coordinate, r = {x, y, z} is the position vector, τ=tz/u, and t is the time.

The perturbed field can then be represented as

A˜(r,t)=A(r,t)+ε=(I01/2+ε˜)exp(iγI0z)
where ε˜=ε1cos(KrΩτ)+iε2sin(KrΩτ).

Substituting Eq. (3) into Eq. (2) and linearing in ε˜, we find

i(ε˜z+1uε˜t)+12k0(2ε˜x2+2ε˜y2)k222ε˜t2+γ(ε˜+ε˜)=0

Linearized equations for small perturbation amplitudes ε1 and ε2 lead to the dispersion relation 4k02Kz2=(K2k0k2Ω2)(K2k0k2Ω24k0γI0), where K=(Kx2+Ky2)1/2 and Kz are the transverse and longitudinal components of K = {Kx, Ky, Kz}. When Kz has a negative imaginary part, the steady-state solution to Eq. (2) becomes unstable with respect to plane-wave perturbations (1). The gain of this instability is found from the imaginary part of Kz as [5]

Γ(Ω,K)=12(K2k0k2Ω2)1/2(4γI0+k2Ω2K2k0)1/2

As can be seen from Eq. (3), in a striking contrast to purely temporal MIs [1], spatiotemporal MIs are not restricted to the region of anomalous dispersion, but exist for group-velocity dispersion k2 and Kerr nonlinearity n2 of any sign except when both k2 and n2 are negative. When the spatial part of MI is suppressed in Eq. (5), K = 0, the properties of purely temporal MIs are recovered. In this regime, in media with n2 > 0, such as standard materials for optical, MIs can only occur when dispersion is anomalous (k2 < 0), with the maximum gain achieved with Ω02=2γI0/k2 [1, 5, 12].

3. Modulation instabilities as parametric amplification via four-wave mixing

We now examine an alternative perspective of spatiotemporal MIs by considering a noncollinear FWM 2ω0 = (ω0 + Ω) + (ω0 − Ω), which couples the pump field with frequency ω0 and wave vector kp to its Stokes and anti-Stokes sidebands with frequencies ω0 ± Ω and wave vectors k ± that are allowed to be generated in the off-axial direction. With kp = k0 + γI0 and k±k0±Ω/u+k2Ω2/2+2γI0, we find that the wave-vector mismatch for this FWM process can be found as Δk=|k1+k22kp|, where k1,2=k±(1K2/k±2)1/2. With k0>>|Ω/u|,|k2|Ω2,|γ|I,K, we find Δkk2Ω2+2γI0K2/k0==Δk0+2γI0, where Δk0 is the intensity-independent part of Δk, determined only by dispersion and wave-mixing geometry.

With K = 0 for on-axis MIs, phase matching Δk = 0 is then achieved at Ω02=2γI0/k2, which exactly recovers the SVEA expression for the sideband frequencies in purely temporal MI [1]. In this regime phase matching and, hence, sideband generation is possible only when the nonlinearity n2 and the second-order dispersion coefficient k2 have opposite signs. In most of the nonlinear materials, including silica, the material most commonly used in optical fibers, n2 > 0, and the ω0 ± Ω sidebands are generated only in the region of anomalous dispersion.

With second-order dispersion effects included in Δk, the buildup of the Stokes and anti-Stokes sidebands occurring as a part of FWM parametric amplification can be described [13] with coupled SVEA equations for the sideband amplitudes A ± at ω0 ± Ω: dA/dz=iγ[2I0A+I0A+exp(iqz)] and dA+/dz=iγ[2I0A++I0Aexp(iqz)], where q=Δk03γI0. The solution to these equations is written as [13]

A(z)=[A1exp(gz)+A2exp(gz)]exp(iΔkz/2)
A+(z)=[A1+exp(gz)+A2+exp(gz)]exp(iΔkz/2)
where the coefficients A1,2± are defined by the initial conditions (e.g., the input noise at ω0 ± Ω) and

g=[(γI0)2(Δk/2)2]1/2

Representing the MI gain Γ given by Eq. (5) as Γ(Ω,K)=(γI0Δk/2)1/2(γI0+Δk/2)1/2, we find that g = Im(Γ). Thus, the gain of ω0 ± Ω sidebands in FWM parametric amplification is equal to the gain predicted by the field evolution equation [Eq. (2)] for the instabilities of the form of Eq. (1). As a straightforward consequence, all the properties of the ω0 ± Ω sidebands in FWM parametric amplification, including their central frequencies, gain bands, and directions of maximum growth rates, are identical to the properties of the spatiotemporal MIs [Eq. (1)] of the steady-state solutions of Eq. (2). Specifically, the gain of spatiotemporal MIs peaks at Δk = 0. Physically, this implies that the maximum MI gain is achieved when the nonlinear phase shifts compensates for the material dispersion, Δk0 = 2γI0,. For anomalously dispersive self-focusing materials (k2 < 0, n2 > 0), spatiotemporal MIs are confined to a region within an ellipse Ω2+Ωs2<Ωc2 in the ΩΩs plane, Ωs=(k0|k2|)1/2Kand Ωc=2(|γ|I0/|k2|)1/2, with the maximum MI gain achieved along the ellipse Ω2+Ωs2=Ωc2/2. In the normal dispersion region, spatiotemporal MIs exist when the conditions Ωs2Ωc2Ω2Ωs2 and Ωs2Ω2Ωs2+Ωc2 are satisfied for materials with positive and negative n2, respectively [5].

The physics behind the equivalence of ω0 ± Ω sidebands in FWM parametric amplification and spatiotemporal MIs of the form of Eq. (1) is, however, anything but clear. Indeed, there is a clear gap between the canonical treatment of FWM parametric amplification based on coupled equations for A ± and the standard linear stability analysis of MIs, which treats MIs as an amplification of weak, yet complex harmonic perturbations [Eq. (1)] of the steady-state solution of Eq. (2).

4. Physical equivalence between two pictures of modulation instabilities

To close this gap and to shed light on the physics behind the equivalence of ω0 ± Ω sidebands in FWM parametric amplification and complex spatiotemporal MIs of the field evolution equation, we represent the perturbed solution to the generic nonlinear field evolution equation as a sum of a steady-state, background field A and a small perturbation [cf. Equation (3)]: A˜=A+41π2ε(K,z)exp(iKr)d2K. With the perturbation taken in the form of a complex harmonic field, ε(K,z)=uK(z)+ivK(z), exactly as prescribed by Eq. (1), the field-evolution equation yields a typical coupled-wave solution [14]: uK(z)=cosh(κz)uK(0)+k0/(4κ)sinh(κz)vK(0), vK(z)=4κ/k0sinh(κz)uK(0)+cosh(κz)vK(0), where κ=(K/2)(2γI0K2/k0)1/2. Such a solution is typical of a vast variety of wave-mixing processes [13, 15], where the fields are coupled by the nonlinear polarization induced in a medium by an intense pump field and where properly phase-matched weak fields can experience amplification, gaining energy from the pump.

To connect these solutions to the solutions for parametrically amplified ω0 ± Ω sidebands in FWM, as described by Eqs. (6) and (7), we represent the evolution of uK(z) and vK(z) in the matrix form [14], wK(z)=TK(z)wK(0), where wK=(uK,vK) is a two-component vector and T(z) is a 2x2 matrix whose four components are defined by the four coefficients in front of uK(0) and vK(0) in the coupled-wave solution governing the evolution of uK(z) and vK(z). Expanding w(z) in the eigenvectors of T(z), wK±=[1+(4κ/k0)2]1/2(1,±4κ/k0), and expressing the result in terms of the respective eigenvalues, θ±=exp(±κz), we find wK(z)=bK+(0)exp(κz)wK++bK(0)exp(κz)wK, where bK±(0) are defined by the initial conditions (e.g., the input noise).

It is straightforward to see from these equations that a weak complex harmonic perturbation [Eq. (1)] gives rise to two waves, wK±, of which one is exponentially growing and the other one is exponentially decreasing. This is exactly what the two parametrically amplified waves A1± and A2± in Eqs. (6) and (7) do. Thus, although the complex harmonic wave of Eq. (1) appears in the standard MI treatment as a purely nominal trial function, which helps isolate MIs in space and time, allowing their dispersion relation to be defined through the pertinent characteristic equation for Kz, such a two-wave field structure is, in fact, more than a purely formal substitution. Rather, it is inherent in secondary waves generated in a broad class of FWM processes as a part of nonlinear spatiotemporal field evolution. Since this two-wave structure of MI-type fields arising as a result of FWM processes reflects the universal properties of nonlinear field evolution, it is manifested in MIs of different nature and is independent of beam geometry.

5. High-order dispersion in spatiotemporal modulation instabilities

We are going to show now that the parametric amplification perspective of spatiotemporal MIs offers a convenient framework to understand the role of high-order dispersion in spatiotemporal MIs and suggests a physically transparent method to calculate corrections to the MI gain due to these effects. Such an extension of MI analysis beyond the standard SVEA procedures would offer valuable physical insights into spatial instabilities of few-cycle and subcycle field waveforms [16] and help understand multiple filamentation of high-power ultrashort laser pulses [3, 4].

Effects related to high-order dispersion can be examined by using Eq. (8) for the parametric gain g with Δk including dispersion terms beyond k2, Δk2m=2j(km/m!)(ωω0)m+2γI0K2/k0, j = 1, 2, …. The MI gain is thus given by

Γ(Ω,K)=12(K2k02m=2j(km/m!)Ωm)1/2,(4γI0+2m=2j(km/m!)ΩmK2k0)1/2.

As can be seen from this expression, only even-order dispersion terms contribute to spatiotemporal MIs. In a particular case of purely temporal MIs, Eq. (9) with K = 0 recovers the results of standard linear stability analysis of Eq. (2) without the diffraction term using the complex harmonic field ansatz of Eq. (1) [17–19], which shows that the third-order dispersion term does not appear in the gain of purely temporal MI. Equation (9) extends the analysis to MIs with K ≠ 0 and allows effects related to the dispersion of any order to be quantified.

It is straightforward to see from Eq. (9) that, with higher order dispersion included, the maximum gain is still g0 = γI0, and it is still achieved when phase matching is satisfied, Δk = 0. However, as can also be seen from Eq. (9), high-order dispersion gives rise to new effects in spatiotemporal MIs, never observed when only the k2 term is included. In Figs. 1(a)–1(l), these new effects are illustrated for the parameters of dispersion and nonlinearity similar to those of YAG (n0 ≈1.76, n2 ≈4 10−16 cm2/W, zero group-velocity dispersion (GVD) wavelength is λz ≈1605 nm) – material that enables efficient filamentation-assisted compression of high-peak-power laser pulses [20]. Deep in the anomalous dispersion region [λ0 = 2π/ω0 = 1.8 μm, Figs. 1(j)–1(l)], the second-order dispersion plays the dominant role, with higher order dispersion only slightly distorting the MI gain bands, giving rise to small corrections to the Ω2+Ωs2=Ωc2/2 phase-matching condition, needed for the maximum MI gain.

 figure: Fig. 1

Fig. 1 The MI gain g(Ω, K) for dispersion and nonlinearity of YAG for a pump field with I0 = 0.5 TW/cm2 and λ0 = 1200 nm (a, b, c), 1400 nm (d, e, f), 1600 nm (g, h, i), 1800 nm (j, k, l). Dispersion is included up to the second (a, d, g, j), fourth (b, e, h, k), and twelfth (c, f, i, l) order.

Download Full Size | PDF

The scenery, however, changes drastically near λz, where |k2| is small. Here, the k2-only approximation fails both to the left and to the right of λz,, while Eq. (9) predicts smooth and continuous transformation of elliptical MI gain bands, typical of the k2 < 0 regime, toward more complicated profiles, with the curvature of these bands first decreasing and then reversing its sign around Ω ≈0 [λ0 = 1.6 μm, Figs. 1(g)–1(i)]. Deeper into the normal dispersion region [1.2 to 1.4 μm, Figs. 1(a)–1(f)], high-order dispersion effects are striking, as Eq. (9) reveals the existence of new MI gain bands, in addition to the signature hyperbolic MI gain bands [5], predicted by the k2-only approximation. As a result, within a finite range of K, two pairs of MI sidebands experience an exponential buildup through phase-matched FWM parametric amplification.

To gain deeper insights into new MI features induced by high-order dispersion, we focus now on a practically significant case when all the k2j effects with j ≥ 3 are negligible, that is, dispersion up to the fourth-order needs to be included. In this case, the Δk = 0 phase-matching condition allows instructive analytical solutions for the MI sideband frequencies. In particular, near the zero-GVD wavelength, where the k2 term in Δk is small, phase matching, Δk = 0, is achieved at Ω02[12(K2/k02γI0)/k4]1/26k2/k4. As can be seen from Figs. 1(b), 1(e), 1(h), 1(k), this approximation is reasonably accurate in reproducing all the main tendencies in the behavior of the MI gain. Keeping only the leading term in this expression and setting K = 0, we recover the results of the earlier studies for purely temporal MIs [17, 19], Ω0±(24γI0/|k4|)1/4. MIs of this type can only occur when k4 is negative. With the spatial part of MI included, however, the MIs, due to the K term in Δk, are no longer limited to the region of negative k4 [Figs. 1(b), 1(e), 1(h), 1(k)].

In the general case, when both k2 and k4 effects are nonnegligible, the fourth-order dispersion may dramatically change the entire picture of spatiotemporal MI [Figs. 1(b), 1(h)]. As perhaps its most striking MI manifestation, the k4 term in Eq. (9) can give rise to the second MI gain band [Fig. 1(b)]. Indeed, solving the equation Δk=2γI0+k2Ω2+k4Ω4/12K2/k0=0, we find

Ω±2=6k2k4{1±[113k4k22(2γI0K2k0)]1/2}

When the k4 term is small, Eq. (10) gives two solutions, Ω2(K2/k02γI0)/k2and Ω+212k2/k4Ω2, corresponding to two pairs of MI sidebands. This result is fully consistent with MI gain calculations with all the dispersion orders included [Figs. 1(c), 1(f)], providing important insights into the physics behind this effect. It is straightforward to see now that the first of these sideband pairs represents the ordinary MI gain band of k2-only MI analysis. This band has a shape of an ellipse for k2 < 0 [Figs. 1(j), 1(k)] and a hyperbola for k2 > 0 [Figs. 1(a), 1(b)].

As long as k4 is small, the second MI band never shows up, lost at Ω → ± ∞. With k2 < 0 and n2 > 0, spatiotemporal MIs in this regime can only exist within the region Ω2[1k4Ω2/(12|k2|)]+Ωs2<Ωc2, whose boundary is slightly distorted relative to the ellipse Ω2+Ωs2=Ωc2, which bounds the region of ordinary, k2-only MIs. However, when the k4 term in Δk becomes comparable to the k2 term, it can manifest itself in a very prominent way, dramatically distorting MI gain bands and giving rise to the second pair of MI sidebands [Fig. 1(b)], in agreement with the full MI gain analysis including all the dispersion orders [Fig. 1(c)]. Clearly, the approximation that neglects all the k2j effects for j ≥ 3 fails within a narrow region near λz where k2 and k4 have the same sign [Figs. 1(d)–1(f)]. Higher order dispersion effects have to be included for an accurate description of spatiotemporal MIs in this regime.

As shown in the earlier work [21], transverse MIs in counterpropagating optical waves can be understood in terms of FWM processes that couple counterpropagating pump waves and their Kerr-nonlinearity-induced sidebands. The case of transverse MIs in counterpropagating waves is, however, drastically different from single-beam spatiotemporal MIs considered in this work in its physics and mathematical framework, as well as in its place in the context of ultrafast optical science and its current challenges. Physically, transverse MIs in counterpropagating waves are induced through the FWM of sidebands generated by the counterpropagating fields due to the Kerr nonlinearity. The sidebands induced as a part of single-beam spatiotemporal MIs considered in this work are all generated in the direction of the only input laser field involved in the problem. The generation of these sidebands is not possible without group-velocity dispersion of the medium. In a striking contrast, the GVD does not even enter the analysis of transverse MIs in counterpropagating waves. Mathematically, analysis of transverse MIs in counterpropagating waves is based on coupled-wave equations for the slowly varying envelopes of the counterpropagating fields that do not include the higher-order time derivative terms [21]. For single-beam spatiotemporal MIs examined in this work, these terms are, on the contrary, are of crucial significance. As a result, the question of how important high-order dispersion effects are, which is one of the key questions of concern in ultrafast photonics of high-peak-power ultrafast optical waveforms (see, e.g., [3, 4, 10, 11, 20]), as well as in this study, is not even relevant in the context of transverse MIs in counterpropagating waves. Finally, the role that the MIs of these two types play in the context of optical science and technologies is very different. While transverse MIs in counterpropagating waves play an important role in phase-conjugation applications, single-beam spatiotemporal MIs considered in this work is one of the key factors in a predominantly unidirectional evolution of high-peak-power ultrashort laser pulses. Laser-induced filamentation, where single-beam spatiotemporal MIs examined in this work lead to beam breakup into multiple filaments (e.g., [3, 4, 10, 11, 20]), is, perhaps, one of the most prominent examples, explaining the motivation behind this study.

6. Conclusion

To summarize, the canonical linear-stability analysis of MIs, which treats MIs as instabilities of the steady-state solution of the field evolution equation with respect to weak complex harmonic perturbations has been shown to be physically equivalent to a coupled-wave analysis of off-axis parametric amplification of the Stokes and anti-Stokes fields by an intense pump wave. The physics behind this equivalence has been revealed via a space-evolution transfer matrix, which transforms a complex harmonic trial function used in linear-stability analysis of MIs into a pair of coupled off-axis waves, thus fully recovering a two-wave field structure of the waves undergoing parametric amplification as a part of FWM. This perspective of spatiotemporal MIs has been shown to offer a convenient framework to understand the role of high-order dispersion in spatiotemporal MIs and to suggest a physically transparent method to calculate corrections to the MI gain due to high-order dispersion in a broad class of problems, including multiple filamentation of high-power ultrashort laser pulses. Within a vast parameter space, high-order dispersion has been shown to drastically change the entire picture of spatiotemporal MIs, substantially modifying MI gain bands and giving rise to new pairs of exponentially growing MI sidebands.

Funding

Russian Foundation for Basic Research (RFBR) (project nos. 14-29-07182 and 16-02-00843); Welch Foundation (grant no. A-1801); Russian Science Foundation (RSF) (14-12-00772); ONR (Award No. 00014-16-1-2578).

Acknowledgments

Stimulating discussions with A.A. Voronin are gratefully acknowledged.

References and links

1. G. P. Agrawal, Nonlinear Fiber Optics (Academic, 2001).

2. V. I. Bespalov and V. I. Talanov, “Filamentary structure of light beams in nonlinear media,” JETP Lett. 3, 307 (1966).

3. A. Couairon and A. Mysyrowicz, “Femtosecond filamentation in transparent media,” Phys. Rep. 441(2-4), 47–189 (2007). [CrossRef]  

4. L. Berge, S. Skupin, R. Nuter, J. Kasparian, and J.-P. Wolf, “Ultrashort filaments of light in weakly ionized, optically transparent media,” Rep. Prog. Phys. 70(10), 1633–1713 (2007). [CrossRef]  

5. L. W. Liou, X. D. Cao, C. J. McKinstrie, and G. P. Agrawal, “Spatiotemporal instabilities in dispersive nonlinear media,” Phys. Rev. A 46(7), 4202–4208 (1992). [CrossRef]   [PubMed]  

6. J. Kasparian, M. Rodriguez, G. Méjean, J. Yu, E. Salmon, H. Wille, R. Bourayou, S. Frey, Y.-B. André, A. Mysyrowicz, R. Sauerbrey, J.-P. Wolf, and L. Wöste, “White-light filaments for atmospheric analysis,” Science 301(5629), 61–64 (2003). [CrossRef]   [PubMed]  

7. A. Couairon, M. Franco, A. Mysyrowicz, J. Biegert, and U. Keller, “Pulse self-compression to the single-cycle limit by filamentation in a gas with a pressure gradient,” Opt. Lett. 30(19), 2657–2659 (2005). [CrossRef]   [PubMed]  

8. A. V. Mitrofanov, A. A. Voronin, D. A. Sidorov-Biryukov, A. Pugžlys, E. A. Stepanov, G. Andriukaitis, T. Flöry, S. Ališauskas, A. B. Fedotov, A. Baltuška, and A. M. Zheltikov, “Mid-infrared laser filaments in the atmosphere,” Sci. Rep. 5, 8368 (2015). [CrossRef]   [PubMed]  

9. E. A. Stepanov, A. A. Lanin, A. A. Voronin, A. B. Fedotov, and A. M. Zheltikov, “A solid-state source of subcycle pulses in the mid-infrared,” Phys. Rev. Lett. (to be published).

10. D. Kartashov, S. Ališauskas, G. Andriukaitis, A. Pugžlys, M. Shneider, A. Zheltikov, S. L. Chin, and A. Baltuška, “Free-space nitrogen gas laser driven by a femtosecond filament,” Phys. Rev. A 86(3), 033831 (2012). [CrossRef]  

11. L. Bergé, S. Mauger, and S. Skupin, “Multifilamentation of powerful optical pulses in silica,” Phys. Rev. A 81(1), 013817 (2010). [CrossRef]  

12. D. Anderson and M. Lisak, “Modulational instability of coherent optical-fiber transmission signals,” Opt. Lett. 9(10), 468–470 (1984). [CrossRef]   [PubMed]  

13. R. H. Stolen and J. E. Bjorkholm, “Parametric amplification and frequency conversion in optical fibers,” IEEE J. Quantum Electron. 18(7), 1062–1072 (1982). [CrossRef]  

14. J. F. Holzrichter, D. Eimerl, E. V. George, J. B. Trenholme, W. W. Simmons, and J. T. Hunt, “High power pulsed lasers,” J. Fusion Energy 2(1), 5–45 (1982). [CrossRef]  

15. Y. R. Shen, The Principles of Nonlinear Optics (Wiley-Interscience, 1984).

16. T. Balciunas, C. Fourcade-Dutin, G. Fan, T. Witting, A. A. Voronin, A. M. Zheltikov, F. Gerome, G. G. Paulus, A. Baltuska, and F. Benabid, “A strong-field driver in the single-cycle regime based on self-compression in a kagome fibre,” Nat. Commun. 6, 6117 (2015). [CrossRef]   [PubMed]  

17. M. J. Potasek, “Modulation instability in an extended nonlinear Schrödinger equation,” Opt. Lett. 12(11), 921–923 (1987). [CrossRef]   [PubMed]  

18. S. B. Cavalcanti, J. C. Cressoni, A. S. Gouveia-Neto, and A. S. Gouveia-Neto, “Modulation instability in the region of minimum group-velocity dispersion of single-mode optical fibers via an extended nonlinear Schrödinger equation,” Phys. Rev. A 43(11), 6162–6165 (1991). [CrossRef]   [PubMed]  

19. A. Höök and M. Karlsson, “Ultrashort solitons at the minimum-dispersion wavelength: effects of fourth-order dispersion,” Opt. Lett. 18(17), 1388–1390 (1993). [CrossRef]   [PubMed]  

20. V. Shumakova, P. Malevich, S. Alisauskas, A. A. Voronin, A. M. Zheltikov, D. Faccio, D. Kartashov, R. Maksimenka, G. Gitzinger, N. Forget, A. Baltuska, and A. Pugzlys, “250-GW sub-three-cycle multi-millijoule mid-IR pulses self-compressed in a YAG plate,” Proc. CLEO:2015 (New York: Optical Society of America) p FTu4D.1 (2015). [CrossRef]  

21. G. G. Luther and C. J. McKinstrie, “Transverse modulational instability of counterpropagating light waves,” J. Opt. Soc. Am. B 9(7), 1047–1060 (1992). [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 (1)

Fig. 1
Fig. 1 The MI gain g(Ω, K) for dispersion and nonlinearity of YAG for a pump field with I0 = 0.5 TW/cm2 and λ0 = 1200 nm (a, b, c), 1400 nm (d, e, f), 1600 nm (g, h, i), 1800 nm (j, k, l). Dispersion is included up to the second (a, d, g, j), fourth (b, e, h, k), and twelfth (c, f, i, l) order.

Equations (10)

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

ε=[ ε 1 cos( KrΩτ )+i ε 2 sin( KrΩτ ) ]exp( iγ I 0 z )
i( A z + 1 u A t )+ 1 2 k 0 ( 2 A x 2 + 2 A y 2 ) k 2 2 2 A t 2 +γ | A | 2 A=0
A ˜ ( r,t )=A( r,t )+ε=( I 0 1/2 + ε ˜ )exp( iγ I 0 z )
i( ε ˜ z + 1 u ε ˜ t )+ 1 2 k 0 ( 2 ε ˜ x 2 + 2 ε ˜ y 2 ) k 2 2 2 ε ˜ t 2 +γ( ε ˜ + ε ˜ )=0
Γ( Ω, K )= 1 2 ( K 2 k 0 k 2 Ω 2 ) 1/2 ( 4γ I 0 + k 2 Ω 2 K 2 k 0 ) 1/2
A ( z )=[ A 1 exp( gz )+ A 2 exp( gz ) ]exp( iΔkz /2 )
A + ( z )=[ A 1 + exp( gz )+ A 2 + exp( gz ) ]exp( iΔkz /2 )
g= [ ( γ I 0 ) 2 ( Δk /2 ) 2 ] 1/2
Γ( Ω, K )= 1 2 ( K 2 k 0 2 m=2j ( k m / m! ) Ω m ) 1/2 , ( 4γ I 0 +2 m=2j ( k m / m! ) Ω m K 2 k 0 ) 1/2 .
Ω ± 2 = 6 k 2 k 4 { 1± [ 1 1 3 k 4 k 2 2 ( 2γ I 0 K 2 k 0 ) ] 1/2 }
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.