Файл: Staliunas K., Sanchez-Morcillo V.J. (eds.) Transverse Patterns in Nonlinear Optical Resonators(ST.pdf

ВУЗ: Не указан

Категория: Не указан

Дисциплина: Не указана

Добавлен: 28.06.2024

Просмотров: 570

Скачиваний: 0

ВНИМАНИЕ! Если данный файл нарушает Ваши авторские права, то обязательно сообщите нам.

References 29

184.A. Dreischuh, S. Chervenkov, D. Neshev, G.G. Paulus and H. Walther, Generation of lattice structures of optical vortices, J. Opt. Soc. Am. B 19, 550 (2002). 11

185.V. Basistiy, V.Yu. Bazhenov, N.S. Soskin and M.V. Vasnetsov, Optics of light beams with screw dislocations, Opt. Commun. 103, 422 (1993). 11

186.G.A. Swartzlander and C.T. Law, Optical vortex solitons observed in Kerr nonlinear media, Phys. Rev. Lett. 69, 2503 (1992). 11

187.G.A. Swartzlander and C.T. Law, The optical vortex soliton, Opt. Photon. News 4, 10 (1993). 11

188.G.A. Swartzlander and C.T. Law, Polarized optical vortex solitons: Instabilities and dynamics in Kerr nonlinear media, Chaos, Solitons Fractals 4, 1759 (1994). 11

189.K.T. Gahagan and G.A. Swartzlander, Optical vortex trapping of particles, Opt. Lett. 21, 827 (1996). 11

190.G.A. Swartzlander, C.T Law and D. Rozas, Propagation dynamics of optical vortices, J. Opt. Soc. Am. B 14, 3054 (1997). 11

191.G.A. Swartzlander, D. Rozas and Z.S. Sacks, Experimental observation of fluidlike motion of optical vortices, Phys. Rev. Lett. 79, 3399 (1997). 11

192.G.A. Swartzlander, Z.S. Sacks and D. Rozas, Holographic formation of optical-vortex filaments, J. Opt. Soc. Am. B 15, 2226 (1998). 11

193.G.A. Swartzlander, Nonlinear optical catastrophe from a smooth initial beam, Opt. Lett. 24, 1224 (1999). 11

194.D. Rozas and G.A Swartzlander, Observed rotational enhancement of nonlinear optical vortices, Opt. Lett. 25, 126 (2000). 11

195.Y.S. Kivshar and E.A. Ostrovskaya, Optical vortices, Opt. Photon. News 12, 27 (2001). 11

196.B. Luther-Davies, J. Christou, V. Tikhonenko and Y.S. Kivshar, Optical vortex solitons: experiment versus theory, J. Opt. Soc. Am. B 14, 3045 (1997). 11

197.W.J. Firth and C. Pare, Transverse modulational instabilities for counterpropagating beams in Kerr media, Opt. Lett. 13, 1096 (1988). 11

198.G. Grynberg, E. Le Bihan, P. Verkerk, P. Simoneau, J.R.R. Leite, D. Bloch, S. Le Boiteaux and M. Ducloy, Observation of instabilities due to mirrorless four-wave mixing oscillation in sodium, Opt. Commun. 67, 363 (1988). 11

199.G. Grynberg and J. Paye, Spatial instability for a standing wave in a nonlinear medium, Europhys. Lett. 8, 29 (1989). 11

200.M. Haelterman, A.P. Sheppard and A.W. Snyder, Bimodal counterpropagating spatial solitary-waves, Opt. Commun. 103, 145 (1993). 11

201.M. Sa man, A.A. Zozulya and D.Z. Anderson, Transverse instability of energy-exchanging counterpropagating waves in photorefractive media, J. Opt. Soc. Am. B 11, 1409 (1994). 11

202.J.B. Geddes, R.A. Indik, J.V. Moloney and W.J. Firth, Hexagons and squares in a passive nonlinear optical system, Phys. Rev. A 50, 3471 (1994). 11

203.P.M. Lushnikov and A.V. Mamaev, Spontaneous hexagon formation in photorefractive crystal with a single pump wave, Opt. Lett. 24, 1511 (1999). 11, 12

204.M. Schwab, C. Denz and M. Sa man, Transverse modulational instability in counterpropagating two-wave mixing with frequency detuning of the pump beams, J. Opt. Soc. Am. B 18, 628 (2001). 11

205.S. Pitois, G. Millot and S. Wabnitz, Polarization domain wall solitons with counterpropagating laser beams, Phys. Rev. Lett. 81, 1409 (1998). 11


30 1 Introduction

206.W.J. Firth, Spatial instabilities in a Kerr medium with a single feedback mirror, J. Mod. Opt. 37, 151 (1991). 12

207.G.P. D’Alessandro and W.J. Firth, Spontaneous hexagon formation in a nonlinear optical medium with feedback mirror, Phys. Rev. Lett. 66, 2597 (1991). 12, 13

208.G. D’Alessandro, E. Pampaloni, P.L. Ramazza, S. Residori and F.T. Arecchi, Amplitude equations and pattern formation in a liquid crystal light valve experiment, Phys. Rev. A 52, 4176 (1995). 12

209.A.J. Scroggie and W.J. Firth, Pattern formation in an alkali-metal vapour with a feedback mirror, Phys. Rev. A 53, 2752 (1996). 12

210.M.A. Vorontsov and A.Yu. Karpov, Pattern formation due to interballoon spatial mode coupling, J. Opt. Soc. Am. B 14, 34 (1997). 12

211.G. Giusfredi, J.F. Valley, R. Pon, G. Khitrova and H.N. Gibbs, Optical instabilities in sodium vapor, J. Opt. Soc. Am. B 5, 1181 (1988). 12

212.W. Lange, Yu.A. Logvin and T. Ackemann, Spontaneous optical patterns in an atomic vapor: observation and simulation, Physica D 96, 230 (1996). 12

213.M.A. Vorontsov and A.Y. Karpov, Stimulated optical patterns in a Kerr slice-feedback mirror-type experiment, J. Mod. Opt. 44, 439 (1997). 12

214.C. Denz, M. Schwab, M. Sedlatschek, T. Tschudi and T. Honda, Pattern dynamics and competition in a photorefractive feedback system, J. Opt. Soc. Am. B 15, 2057 (1998). 12

215.M. Kreuzer, A. Schreiber and B. Th¨uring, Evolution and switching dynamics of solitary spots in nonlinear optical feedback system, Mol. Cryst. Liq. Cryst. 282, 91 (1996). 12

216.A. Schreiber, B. Thuring, M. Kreuzer and T. Tschudi, Experimental investigation of solitary structures in a nonlinear optical feedback system, Opt. Commun. 136, 415 (1997). 12

217.M. Kreuzer, B. Th¨uring and T. Tschudi, Creation, dynamics and stability of localized states in a nonlinear optical feedback system, Asian J. Phys. 7, 678 (1998). 12

218.Yu.A. Logvin, B. Sch¨apers and T. Ackemann, Stationary and drifting localized structures near a multiple bifurcation point, Phys. Rev E 61, 4622 (2000). 12

219.B. Sch¨apers, M. Feldmann, T. Ackemann and W. Lange, Interaction of localized structures in an optical pattern-forming system, Phys. Rev. Lett. 85, 748 (2000). 12, 13

220.S.A. Akhmanov, M.A. Vorontsov, V.Yu. Ivanov, A.V. Larichev and N.I. Zheleznik, Controlling transverse wave interactions in nonlinear optics: generation and interaction of spatiotemporal structures, J. Opt. Soc. Am. B 9, 78 (1992). 14

221.M.A. Vorontsov, N.G. Iroshnikov and R.L. Abernathy, Di ractive patterns in a nonlinear optical system with spatial interaction, Chaos, Solitons Fractals 4, 1701 (1994). 14

222.N.I. Zheleznik, M. Le Berre, F. Ressayre and A. Tallet, Rotating spiral waves in a nonlinear optical system with spatial interaction, Chaos, Solitons Fractals 4, 1717 (1994). 14

223.M.A. Vorontsov and W.J. Firth, Pattern formation and competition in nonlinear optical systems with two-dimensional feedback, Phys. Rev. A 49, 2891 (1994). 14


References 31

224.M.A. Vorontsov and B.A. Samson, Nonlinear dynamics in an optical system with controlled two-dimensional feedback: black-eye patterns and related phenomena, Phys. Rev. A 57, 3040 (1998). 13, 14

225.B.A. Samson and M.A. Vorontsov, Localized states in a nonlinear optical system with a binary-phase slice and a feedback mirror, Phys. Rev. A 56, 1621 (1997). 14

226.E. Pampaloni, S. Residori and F.T. Arecchi, Roll–hexagon transition in a Kerr-like experiment, Europhys. Lett. 24, 647 (1993). 14

227.E. Pampaloni, P.L. Ramazza, S. Residori and F.T. Arecchi, Experimental evidence of boundary-induced symmetries in an optical system with a Kerr-like nonlinearity, Europhys. Lett. 25, 587 (1994). 14

228.F.T. Arecchi, A.V. Larichev, P.L. Ramazza, S. Residori, J.C. Ricklin and M.A. Vorontsov, Experimental observation of space–time chaos in a nonlinear optical system with 2D feedback, Opt. Commun. 117, 492 (1995). 14

229.P.L. Ramazza, S. Residori, E. Pampaloni and A.V. Larichev, Transition to space–time chaos in a nonlinear optical system with two-dimensional feedback, Phys. Rev. A 53, 400 (1996). 14

230.P.L. Ramazza, S. Ducci and F.T. Arecchi, Optical di raction-free patterns induced by a discrete translational transport, Phys. Rev. Lett. 81, 4128 (1998). 14

231.E. Pampaloni, P.L. Ramazza, S. Residori and F.T. Arecchi, Two-dimensional crystals and quasicrystals in nonlinear optics, Phys. Rev. Lett. 74, 258 (1995). 14, 15

232.S. Residori, P.L. Ramazza, E. Pampaloni, S. Boccaletti and F.T. Arecchi, Domain coexistence in two dimensional optical patterns, Phys. Rev. Lett. 76, 1063 (1996). 14

233.E. Pampaloni, S. Residori, S. Soria and F.T. Arecchi, Phase locking in nonlinear optical patterns, Phys. Rev. Lett. 78, 1042 (1997). 14

234.P.L. Ramazza, S. Boccaletti, A. Giaquinta, E. Pampaloni, S. Soria and F.T. Arecchi, Optical pattern selection by lateral wavefront shift, Phys. Rev. A 54, 3472 (1996). 14

235.P.L. Ramazza, S. Boccaletti and F.T. Arecchi, Transport induced patterns in an optical system with focussing nonlinearity, Opt. Commun. 136, 267 (1997). 14

236.P.L. Ramazza, E. Benkler, U. Bortolozzo, S. Boccaletti, S. Ducci and F.T. Arecchi, Tailoring the profile and interactions of optical localized structures, Phys. Rev. E 65, 066204 (2002). 15


2 Order Parameter Equations for Lasers

Order parameter equations are the simplest (minimal) equations that describe, in the leading order, the dynamics of the field in broad-aperture lasers and other nonlinear optical systems. This chapter is devoted to the derivation of OPEs for lasers.

The OPEs have a twofold significance in studying the formation and evolution of transverse patterns in nonlinear optics. First, they usually allow a simplification of the analytical and numerical treatment, since the OPEs are structurally simpler than the initial (microscopic) equations. The OPEs are often obtained in the form of the complex Ginzburg–Landau or the complex Swift–Hohenberg equation. These equations have been intensively studied in recent years outside optics, and their properties are well, although not completely, known. Thus, a reduction to these equations solves a part of the problem [1].

Second, the OPEs allow one to consider the patterns in a particular system (in our case, a laser) from a general point of view. As shown below, the reduced laser equations in some limits are similar to the hydrodynamic (Navier–Stokes) equation. In the other limit (a class B laser), the reduced equations are similar to those derived for oscillatory chemical and biological systems. Thus the derivation of OPEs allows one to demonstrate an analogy between nonlinear optics and hydrodynamics in one limit, and between nonlinear optics and oscillatory chemical systems in the other limit. A knowledge of the existence of vortices in superfluids allows one to predict vortices in nonlinear optics thanks to the optics–hydrodynamics analogy, and similarly a knowledge of “self-sustained meandering” of spiral waves in oscillatory or excitable systems provides a motivation to look for the “restless vortex” in class B lasers.

We start from a semiclassical model of a laser, the Maxwell–Bloch (MB) system of equations, which includes transverse degrees of freedom. We then reduce the MB system using two alternative methods: adiabatic elimination (Sect. 2.3.1) and the multiscale expansion technique (Sect. 2.3.2). We deal in this chapter with class A and class C lasers, since these cases lead to a single OPE in the form of the CGL or CSH equation. The study of a class B laser (a laser with relatively slow population inversion) is postponed to

K. Stali¯unas and V. J. S´anchez-Morcillo (Eds.):

Transverse Patterns in Nonlinear Optical Resonators, STMP 183, 33–49 (2003)c Springer-Verlag Berlin Heidelberg 2003

34 2 Order Parameter Equations for Lasers

Chap. 7, since this class of laser leads to more complicated OPEs, and to richer dynamics than those of class A and class C lasers.

2.1 Model of a Laser

Our starting point is the semiclassical Maxwell–Bloch (MB) equation system, which describes many types of lasers with transverse degrees of freedom:

∂E

= κ

(1 + iω) E + P + id

 

2E ,

(2.1a)

∂P

 

 

 

 

 

∂t

 

 

 

 

 

 

 

 

= −γ (P − ED) ,

 

 

 

 

(2.1b)

∂t

2 (EP + E P ) .

∂t

= −γ

D − D0 +

(2.1c)

∂D

 

 

1

 

 

 

 

The complex fields E(r, t) and P (r, t) are the envelopes of the electromagnetic (optical) field and of the polarization, and D(r, t) is the real valued field of the population inversion, which in the absence of stimulated radiation is equal to its unsaturated value D0(r). κ is the relaxation rate of the optical field in the resonator due to the (small) transmittivity of the mirrors and to linear losses in the resonator; γ and γ are the decay rates of the polarization and population inversion, respectively. Finally, ω is the resonator detuning (the detuning of the resonance frequency of the corresponding longitudinal mode with respect to the center of the gain line).

It is assumed that the optical field E(r, t) is linearly polarized, and the gain line is homogeneously broadened. It is also assumed that only one longitudinal mode family is excited. Otherwise, there would be a dependence on the longitudinal coordinate z, and not only a dependence on the time t and on the transverse coordinates r = (x, y) as in (2.1).

The system (2.1) describes a laser with multiple transverse modes but a single longitudinal mode. The evolution in time is assumed to occur on a timescale much slower than the round-trip time of the light in the cavity; otherwise, the assumption of a single longitudinal mode would be invalid.

The di raction term is related to the spatial degrees of freedom. This term, being nonlocal, couples the field throughout the cross section of the laser, and is responsible for the collective behavior of the laser radiation.

The simplest limit of the Maxwell–Bloch system (2.1) is the class A laser, in which the polarization and population inversion are fast compared with the optical field in the resonator. In this limit, sometimes called the “good cavity limit”, the fast material variables can be adiabatically eliminated, and a relatively simple order parameter equation can be obtained. A straightforward adiabatic elimination has been performed in this case in [2]. Although many pattern-forming properties of lasers are lost in this adiabatic elimination, let us start from this procedure.


2.1 Model of a Laser

35

It is assumed that the material variables decay fast, i.e. κ/γ = O(ε) and κ/γ = O(ε), where ε is a smallness parameter, and that the temporal derivatives of all variables have finite values, i.e. ∂E/∂t ∂P/∂t ∂D/∂t = O(1). Multiplying both sides of (2.1b) and (2.1c) by κ/γ and κ/γ , respectively, we obtain the result that the left-hand sides of both equations are of order O(ε). Keeping only the terms of zero order O(1), we can eliminate the material variables from (2.1b) and (2.1c), and obtain

D =

D0

 

,

P =

D0E

 

.

(2.2)

1 + |E|

2

1 + |E|

2

 

 

 

 

 

 

 

The expressions (2.2) imply that the material variables P and D follow instantaneously, or adiabatically, the changes of the field variable E. Inserting (2.2) into (2.1a) we obtain a single equation for the field,

∂E

 

D E

 

 

 

= −E +

0

+ i a 2 − ω E ,

(2.3)

∂τ

1 + |E|2

where τ = κt is a slow time. Close to the emission threshold p = (D0 1) 1, the emitted fields are relatively weak, |E|2 1, which allows a cubic approximation for the nonlinear term in (2.3),

∂E

= pE + i a 2 − ω E − E |E|2 ;

(2.4)

∂t

this is the complex Ginzburg–Landau equation. In (2.4), p is the balance between the gain and loss of the laser, and is a criticality parameter of the CGL.

The CGL equation is a crude approximation for a laser. For instance, the selection of transverse wavenumber (transverse mode) is not accounted for by (2.4). A linear stability analysis of the zero (nonlasing) solution of (2.4) leads to equal growth exponents of all the components of the spatial spectrum, which can be easily checked by inserting a test solution in the form of a tilted wave, E(r, t) = e exp(ik · r + λt), with a small amplitude e, linearizing it with respect to e, and calculating the exponent λ. Here k = k is the transverse wavenumber, or, in other words, the transverse component of the wavevector tilted with respect to the optical axis of the resonator. The value of λ is independent of the transverse wavenumber k = |k| and is equal to λ = p = D0 1.

It is well known, however, that lasers emit particular transverse modes (transverse wavenumbers) that depend on the length of the resonator. When the maximum of the gain line coincides with a particular transverse mode family (Fig. 2.1), the corresponding transverse mode is favored, and grows the fastest. This tunability property of the laser is lost in the derivation of (2.4). It is not di cult to understand why spatial-frequency selection is absent in (2.4): the derivation assumes, among other things, the condition γ /κ → ∞,