You are on page 1of 10

F. Poletti and P.

Horak

Vol. 25, No. 10 / October 2008 / J. Opt. Soc. Am. B

1645

Description of ultrashort pulse propagation in multimode optical bers


Francesco Poletti and Peter Horak*
Optoelectronics Research Centre, University of Southampton, Southampton SO17 1BJ, United Kingdom *Corresponding author: peh@orc.soton.ac.uk Received May 14, 2008; accepted July 23, 2008; posted August 5, 2008 (Doc. ID 96105); published September 18, 2008 The guided, single-mode propagation of ultrashort optical pulses is commonly described by a well studied and understood generalized nonlinear Schrdinger equation. Here we present and discuss an extended version for multimode optical bers and waveguides including polarization effects, high-order dispersion, Kerr and Raman nonlinearities, self-steepening effects, as well as wavelength-dependent mode coupling and nonlinear coefcients. We then investigate the symmetry properties of the nonlinear coupling coefcients for the cases of step-index and circularly symmetric conventional bers and for microstructured bers with hexagonal symmetry. Finally, we study the computational complexity of the proposed algorithm. 2008 Optical Society of America OCIS codes: 190.4370, 190.5530, 190.7110, 320.6629.

1. INTRODUCTION
Spectral broadening of short pulses of laser light propagating through a nonlinear optical material was noted as early as 1970 [1]. However, it was only with the introduction of microstructured optical bers (MOFs) in the late 1990s that extreme nonlinear spectral broadening, referred to as supercontinuum generation, could be routinely observed using laser sources commonly available in most optical laboratories [2]. The endlessly single-mode operation, high nonlinearity, and unprecedented dispersion control achievable with these bers [3] have subsequently been employed to generate supercontinuum light with a broad range of spectral and temporal properties [4]. Similar results were recently obtained in tapered photonic nanowires [5]. Supercontinuum light sources have already been used in a variety of applications, most notably in the generation of frequency combs for optical frequency metrology applications [6]. Most of these results were achieved in single-mode or nearly single-mode bers. However, as the available laser power continues to grow there is increasing interest in the nonlinear propagation of ultrashort pulses in optical bers with a large mode area supporting multiple modes. Moreover, several recent experiments have demonstrated that there is also signicant potential for new spatial as well as spectral nonlinear effects in multimode bers [711]. The theoretical description of short-pulse propagation in single-mode bers is commonly based on the nonlinear Schrdinger equation (NLSE), modeling the evolution of the electromagnetic pulse envelope [12,13]. Generalized versions of the NLSE (GNLSE), including higher-order dispersion, and Raman and self-steepening terms, have been successfully employed in the description of octave
0740-3224/08/101645-10/$15.00

spanning supercontinua and for pulse durations approaching the single-cycle regime [4]. It was recently demonstrated that no bandwidth constraint is associated with the nonlinear modeling of the envelope evolution, allowing the accurate study of pulses down to the subcycle regime [14]. For multimode bers, nonlinear Schrdinger equations were rst developed for the investigation of soliton effects [1517] and were limited to Kerr nonlinearities and lowest order dispersion. More recent theories of nonlinear pulse propagation in homogeneous media [18,19] serve to describe spatial effects in bulk media such as nonlinear lamentation. However, ber- or waveguide-specic modeling of multimode pulse propagation by a GNLSE has so far been restricted to two birefringent modes [2024] or two spatially distinct modes [2527]. Alternative approaches only included a limited range of intermodal nonlinear effects, e.g., cross-phase modulation [28] or phasematched four-wave mixing [29], or modeled supercontinuum generation in individual modes independent from each other, thereby neglecting mode-to-mode coupling [30]. Here we present a generic set of equations to describe the propagation of ultrashort pulses in multimode optical bers or waveguides including polarization, high-order dispersion, Kerr and Raman nonlinearities, selfsteepening and shock formation effects, and wavelengthdependent mode coupling and nonlinear coefcients. We derive this multimode GNLSE (MM-GNLSE) in Section 2 and discuss several important features in Section 3. In Section 4 we investigate simplications of the mode coupling terms based on ber symmetries for the important cases of step-index, circularly symmetric, and hexagonal microstructured bers. In Section 5 we study the compu 2008 Optical Society of America

1646

J. Opt. Soc. Am. B / Vol. 25, No. 10 / October 2008

F. Poletti and P. Horak

tational complexity of the proposed algorithm based on a split-step Fourier method and, nally, we conclude our discussions in Section 6.

with envelopes Anz , is thus given by

Ex, t =

2. DERIVATION OF THE MULTIMODE NONLINEAR SCHRDINGER EQUATION


In this section we will derive the generalized multimode NLSE following the approach of Kolesik and Moloney [19]. The starting point is Maxwells equations Ex, t = 0

1 F n x , y , 2 N n

einzAnz, eit + c.c. . 4

Hx, t ,

With these denitions and upon Fourier transformation of Eq. (1) the following propagation equation for the envelope of mode p is obtained [19]:

H x , t = 0n x , y 2

A p z ,
Ex, t 1

i 4 N p

ipz * P dxdyFp . NLe

PNLx, t .

Here x = x , y , z with x , y denoting the transverse coordinates and z the longitudinal coordinate, and we adopt the convention that all vectorial quantities are written in boldface. PNLx , t is the nonlinear polarization of the ber induced by the intrinsic cubic nonlinearity 3 of the glass with a time response function Rt accounting for Kerr as well as Raman nonlinearities, PNLx, t = 03Ex, t

dRt Ex, 2 .

Solving Eqs. (1) without the nonlinear polarization term gives the discrete ber modes n = 1 , 2 , 3 , with propagation constant n and electric and magnetic eld mode functions Fnx , y , expinz and H n x , y , expinz, respectively. These modes fulll the normalization and orthogonality relation 1 4

* H + F H* e dxdyFm n n z m

= mnNn2 ,

where mn is the Kronecker delta and we have omitted the spatial arguments for clarity. The modal decomposition of the electric eld into a sum of mode functions Fnx , y ,

Note that this rst-order envelope equation was derived by separating forward and backward propagating elds and neglecting any nonlinear interactions between them, but without any restriction on the bandwidth over which the model is valid [14,18]. Consequently, only positive frequency components are considered in the following. Following Mamyshev and Chernikov [31] we now proceed to specialize Eq. (5) to the case of glass optical bers with a third-order nonlinearity given by Eq. (2), where PNL involves convolutions of three mode elds FnAn and of the time response function R. In a rst step, Eq. (5) is further simplied by neglecting the frequency dependence of the transverse mode functions and of the corresponding normalization constants, which allows us to separate them from the convolution integrals of PNL, but instead assuming a frequency dependence of the resulting transverse overlap integrals. This frequency dependence is nally truncated to the linear term around a central frequency 0, which represents a free parameter of the model but is often chosen to coincide with the carrier frequency of the initial pulse. As has been emphasized recently [32] this method of accounting for the frequency dependence of transverse mode functions cannot be justied a priori from the form of Eq. (5), but it yields improved results for a range of phenomena [33]. After some algebra we obtain the following MMGNLSE for mode p:

A p z , t z

p p = i0 0 A p z , t 1 1

A p z , t t

+i

1 1 + iplmn

1 Qplmn 0 2 A l z , t

n2

p n

n!

A p z , t + i

n 2 0 c

d R A m z , t A n z , t e 2i0 = D p z , t + N p z , t .

* z , t + 1 + i 2 d R A m z , t A n plmn

l,m,n

2 * z , t Qplmn 0 A l

F. Poletti and P. Horak

Vol. 25, No. 10 / October 2008 / J. Opt. Soc. Am. B

1647

Here, Dpz , t and Npz , t refer to the dispersive and the nonlinear part of the MM-GNLSE for mode p, respectively. The nonlinearity Npz , t couples the mode p to every combination of modes l, m, and n, and thus the propagation of a general ultrashort multimode pulse is described by a set of coupled nonlinear partial differential equations for p = 1 , 2 , 3 , . . .. The mode dispersion Dpz , t appears in Eq. (6) from the inclusion of the frequency dependence of n in the Fourier transform of Eq. (5). Note that mode and frequency dependent propagation losses can be considered by allowing for complex values of the propagation constants p, and that in deriving Eq. (6) we have dropped the non-phase-matched terms at 30 responsible for third-harmonic generation, which are only relevant for single or subcycle propagation dynamics [14]. The overlap integrals are rigorously dened as
1 Qplmn =
2 2 2 dxdyF* F F F* 0 n 0c l m p n

type nonlinearity. Moreover, for a single linearly polarized ber mode with a real-valued mode function F1 = Fex, we have
2 N1

neff0c 2

dxdyF2 ,

10

1 2 Q1111 = Q1111
2 n0

dxdyF4

2 3neff dxdyF22 2 n0 2 3neff Aeff

11

1 2 1 = 1111 = 1111

12

N p N l N m N n

1 =

2 lnneff Aeff

,
0

12

2 Qplmn =

2 2 2 dxdyF* F*F F 0 n 0c m n p l

12

N p N l N m N n

, 7

where Aeff and neff are the usual effective mode area and effective refractive index, respectively. Then the nonlinear term of Eq. (6) becomes N 1 z , t = i n 2 0 cAeff0

and the shock time constants as


1,2 plmn =

1,2 lnQplmn

.
0

1 + i1

A 1 z , t 13

dRA1z, t 2 ,

1 2 = Qplmn and For real-valued mode functions we have Qplmn 1 2 plmn = plmn. In Eq. (6) the constants 0 and 1 / 1 are free parameters corresponding to an overall phase factor and the velocity of a reference frame, respectively, which can be conveniently chosen to represent the propagation constant and group velocity of the fundamental mode at frequency 0. The nonlinear time response function can be written as

with Rt = 1 fRt + fRht, in agreement with the widely used single-mode GNLSE. B. Two-Mode GNLSE Similarly, the MM-GNLSE can be specialized to the case of two birefringent modes of a ber. As an example, we assume two circularly polarized modes, F1 = Fe+ and F2 = Fe, where e = ex iey / 2 are unit vectors of circular polarization and F is again assumed to be real valued. The nonlinear term of Eq. (6) for mode 1 then reads N 1 z , t = i n 2 0 cAeff0

3 R t = 1 f R t + f Rh t , 2

where fR is the fractional contribution of the Raman response to the total nonlinearity and ht is the delayed Raman response function. For example, for silica bers we have fR = 0.18, and ht can either be obtained from an analytic approximation [13] or from a measured curve [34]. The set of coupled nonlinear equations (6) is the main result of this paper and in the following we will proceed to discuss a number of features and consequences of this MM-GNLSE.

1 + i1

1 f R A 1 z , t

4 A1z, t2 + A2z, t2 + fRA1z, t 3 3 2

A1z, t 2 + A2z, t 2 ,

dh 14

3. DISCUSSION
A. Equivalence with Single-Mode GNLSE The integral of the last line of Eq. (6) contains a fast rotating term at twice the optical frequency. It is thus customary to neglect the time delayed part of the response function here and only retain the instantaneous Kerr-

in agreement with [23]. An interesting feature of Eq. (14) is that it is energy conserving for each mode individually as it lacks an intermodal four-wave mixing term. Thus, if a pulse is launched exclusively into one circular polarization mode, no light is coupled into the orthogonal polarization during pulse propagation and any numerical simulation can in fact be restricted to a single-mode GNLSE. This is owing to the symmetry of the modes, which will be discussed in more detail later.

1648

J. Opt. Soc. Am. B / Vol. 25, No. 10 / October 2008

F. Poletti and P. Horak

C. Mode Normalization and Shock Term In the derivation presented in Section 2 we have chosen to normalize the modes with respect to Nn rather than to dxdyFn2, which is more common for the singlemode GNLSE. As a consequence we nd that in the absence of the delayed Raman response the MM-GNLSE conserves energy in the form

dtAnz, t2 = const,

15

semianalytically. The well-known results [13] lead to mode functions that can be written in cylindrical coordinates as Fr , = Frexpim, where m 0 is an integer number. These modes are commonly classied as hybrid electromagnetic modes HEmn and EHmn for m 1, and as transverse electric modes TE0n and transverse magnetic modes TM0n for m = 0; see also Table 1. Here, n = 1 , 2 , . . . indicates the nth mode in each symmetry class. Using the notation of Section 2, we can thus write all modes in the form Fpr, = Fprexpimp , 17

such that Anz , t2 gives the pulse power in units of watts. Including the Raman effects leads to propagation losses, since energy is transferred to phonons in the ber material. 1,2 The shock terms proportional to plmn , also frequently referred to as self-steepening terms, take into account the frequency dependence of the nonlinear coupling and of the effective refractive index, and lead to reshaping of ultrashort pulses during propagation. In the MM-GNLSE these coefcients differ for every set of four mode indices. In numerical simulations based on the widely employed split-step Fourier method (SSFM), the shock terms are most efciently computed by fast Fourier transforms (FFTs) and simple multiplications in the frequency space. Hence, the large number of coefcients will signicantly increase computational times. In practice, it may thus be necessary to neglect the differences between the coef1,2 cients and use the approximation plmn 1 / 0. In this p case, the time derivatives of N z , t can be taken out of the summation over all modes and only a single FFT is required. While this simplication may lead to some inaccuracies of the nal results [33], no qualitative changes in the pulse propagation are expected.

where p is the mode number and mp is an integer number. From the angular part of the overlap integrals, Eq. (7), we obtain immediately the following selection rules:
1 =0 Qplmn 2 Qplmn =0

for mp + ml + mm mn for mp ml + mm + mn

0, 0. 18

4. SYMMETRY CONSIDERATIONS
In principle, the number of nonlinear coupling coefcients 1 2 Qplmn and Qplmn increases with the fourth power of the number of ber modes. However, because of the specic form of Eq. (7) many of these coefcients are identical,
1 1 * =Q * , = Qnmlp = Qlpnm Qplmn mnpl 2 2 2 2 * =Q * = Qlpmn = Qplnm = Qlpnm = Qmnpl Qplmn mnlp 2 2 1 1

In the weakly guiding limit a larger number of modes become degenerate. The mode functions of these approximate modes, usually denoted as LP modes, can then be simplied considerably [35]. In particular, the longitudinal components become negligible and all mode functions become polarized in the plane orthogonal to the propagation direction. More specically, all HE and EH modes with rotational symmetry [Eq. (17)] become either e+ or e circularly polarized at every point in space, as specied in Table 1. The TE and TM modes become linearly polarized but, because they are degenerate in this limit, they can be combined to form new, circularly polarized modes with a symmetry of m = 0. Thus, in the weakly guiding limit all modes can be rewritten as purely circularly polarized modes. Therefore, the scalar product of two mode functions, as required in the evaluation of the nonlinear coupling coefcients [Eq. (7)], will vanish at every point for nonmatching circular polarization modes. If we denote by p = 1 the circular polarization state of mode p in the weakly guiding limit, we obtain the following polarization selection rules in addition to the spatial selection rules (18):
1 Qplmn =0 2 =0 Qplmn

for p

l or m

n ,
19

for p = l or m = n .

2* Qnmpl

2* Qnmlp .

16

Moreover, a large number of coefcients vanish exactly or approximately because of mode symmetries, thus reducing the number of convolution integrals and vector multiplications and improving the algorithms efciency, as will be discussed in Section 5. Since it is often difcult to establish numerically whether the result of a calculation is small or it differs from zero only because of numerical errors, in the following we derive analytical rules identify1,2 ing the identically or approximately zero Qplmn for different types of optical bers. A. Step-Index Fibers In the case of cylindrically symmetric optical bers with a step-index prole, the mode functions can be calculated

In particular, we nd that in this case all terms responsible for four-wave mixing between e+ and e polarization states vanish, that is, if the pump is exclusively in a single circular polarization, no light is coupled into modes of opposite polarization and thus these modes can be exTable 1. Symmetry Properties of Step-Index Fiber Modes
Exact Mode Functions Mode HEmn EHmn TE0n, TM0n Spatial Symmetry expim expim 1 Weakly Guiding Limit Mode LPm1,n LPm+1,n LP1,n Polarization e e e

F. Poletti and P. Horak

Vol. 25, No. 10 / October 2008 / J. Opt. Soc. Am. B

1649

cluded entirely from the simulation of nonlinear pulse propagation. This is a generalization of the symmetry already observed in Eq. (14). Besides halving the number of 2 equations, this approximation also leads to Qplmn = 0 for any combination of the remaining modes. As an example, Fig. 1 shows the distribution of the 1,2 magnitude of all coupling coefcients Qplmn for the rst ten modes of a multimode step-index ber (6 m core radius and numerical aperture NA = 0.17 at 1.5 m). Using the exact, complex-valued mode functions with symmetries [Eq. (17)], 17872 out of the total 2 104 coupling coefcients are found to vanish identically because of the selection rule (18). From the distribution of the remaining coefcients (dashed curve) a plateau is clearly evident, indicating that a further 936 coefcients are below 1% of the maximum coupling coefcient. These small but nonzero terms would be identically zero under the weakly guiding approximation that leads to the selection rule (19) and for practical purposes can be neglected, leaving a total of 1192 signicant coupling coefcients. A further signicant reduction in the number of nonzero terms can be achieved by restricting the modes to those that correspond to a single circular polarization in the weakly guiding limit (dashed-dotted line), as outlined above. In this case only 85 signicant coupling constants remain. Finally, note that instead of using complex mode functions with an expimp azimuthal dependence as in Eq. (17), we could consider the more frequently employed realvalued eigenmodes with a sine or cosine azimuthal dependence, which correspond to linearly polarized modes in the weakly guiding limit [36]. In this case the number of coupling coefcients approximately doubles (solid curve), 1 which is only partly compensated by the fact that Qplmn 2 = Qplmn, and therefore some simplications in Eq. (6) are possible. B. Circular Fibers With Arbitrary Index Prole In the Subsection 4.A we discussed the symmetry properties of step-index optical bers because they are the best

understood class of bers. However, many bers of practical interest present an index prole with a more complicated radial dependence. In the following we briey outline a generalization of the previous results, demonstrating that the modes of any cylindrically symmetric ber with an arbitrary refractive index prole n , r = nr are subject to the same selection rules we have discussed in Subsection 4.A. The transverse electric eld of mode p of a cylindrically symmetric ber fullls the following coupled wave equations in cylindrical coordinates [37]
2 t F r,p

2 F ,p r
2

d ln n2 F r,p dr

F r,p r2

+ n 2k 2 2 F r,p

= 0, 1 d ln n2 r dr +

2 F ,p + t

2 F r,p r

F ,p r2

+ n2k2 2F,p = 0, 20

where Fr,p and F,p are the radial and the azimuthal components of the mode function, respectively. It can be shown that any general modal solution is of the form Fr,pr , = Fr,prexpimp, F,pr , = F,prexpimp with an integer value of mp [37]. Thus the modes obey the same symmetry properties as the step-index ber modes, Eq. (17), and the selection rule (18) equally applies. Furthermore, in the weakly guiding (scalar) limit, the mode functions are given by [37] Fpr, = epFprexpimp , 21

where p = . Thus, also the approximate polarization selection rule (19) holds for general cylindrical, weakly guiding bers. C. Microstructured Fibers With Hexagonal Symmetry Microstructured optical bers allow for improved control of the nonlinear and dispersive properties compared to conventional bers. In recent years, various different MOF designs have therefore been employed in nonlinear optics experiments: silica solid-core microstructured bers [2,4], soft-glass bers [38], and even hollow-core designs [39]. No analytical solution exists for the mode functions of these bers, and thus Maxwells eigenproblem must be solved numerically. This prevents the derivation of simple selection rules such as selection rule (18), based on the analytical form of the eigenmodes. A different set of rules, however, can still be derived through symmetry analysis. It is well-known that the most common MOFs fabricated through a stack-and-draw procedure present a triangular lattice of air holes, which exhibits a sixfold rotation symmetry with six planes of reection and therefore belongs to the C6v symmetry group. Exploiting symmetry analysis, McIsaac showed that the eigenmodes of longitudinally uniform waveguides with C6v symmetry can be divided into eight distinct classes, four of which contain nondegenerate modes (and are indicated by the class index P = 1 , 2 , 7 , 8), and four of which contain modes that are degenerate with corresponding modes in a comple-

4000 Number of Q(1,2)> max{Q(1,2)}

3000

2000

1000

0 8 10

10

10 Threshold

10

10

Fig. 1. (Color online) Distribution of the magnitude of coupling 1 2 and Qplmn for the rst ten modes of a step-index coefcients Qplmn ber with a 6 m core radius, a core refractive index of 1.45, and a cladding refractive index of 1.44 at 1.5 m wavelength. Dashed curve: exact mode functions with rotational symmetry [Eq. (17)]; solid curve: exact real-valued mode functions; dashed-dotted curve: ve approximate e+ polarized mode functions only. See text for details.

1650

J. Opt. Soc. Am. B / Vol. 25, No. 10 / October 2008


1,2 Qplmn =0

F. Poletti and P. Horak

mentary mode class (class P = 3 is degenerate with P = 4 and P = 5 with P = 6) [40]. Each mode class presents a different azimuthal symmetry for the electromagnetic elds. McIsaac also suggested that the azimuthal dependence of the longitudinal components Fz,pr , of the electric elds can be expressed in terms of Fourier series in the azimuthal angle [40]. From these results the corresponding Fourier series form of the radial and azimuthal components of the electric eld, Fr,pr , and F,pr , , can be derived, which we summarize in Table 2. These symmetry properties lead to an orthogonality relation between modes in different classes,

for Pp + Pl + Pm + Pn = 2M + 1,

M 23

= 2,3,4, . . . .

All the remaining nonzero overlap integrals present an angular part that can be written as the sum of cosine functions of integrated between 0 and 2. These terms vanish identically whenever the cumulative sums of the phases of the modes involved are different from zero. The second rule therefore can be expressed as
1,2 Qplmn =0

for Kp Kl Km Kn

6N,

N = 0, 1, 2, 24

*F= dFp l

* F + F* F * F dFr + Fz ,p r,l ,p ,l ,p z,l

=0

for Pp

Pl ,

22

where Pi indicates the symmetry class to which mode i belongs. Inserting the series representation for the transverse mode functions shown in Table 2 into the overlap integrals [Eq. (7)] and expanding the sine and cosine products, two selection rules can be derived. According to the 1,2 rst rule, the overlap integrals Qplmn are identically zero for those four-mode combinations in which an odd number of modes presents an odd class number P, because in this case Eq. (7) can be written as a sum of sine functions of , integrated between 0 and 2. This rule, which halves the number of nonzero coefcients, can be written as

where Ki is the coefcient shown in Table 2, corresponding to the class Pi of mode i. The selection rules (23) and (24) reduce the total number of mode class combinations that lead to nonzero nonlinear coupling coefcients from 84 = 4096 to 768. As an example, Fig. 2 shows the distribution of the magnitude 1 2 of coupling coefcients Qplmn and Qplmn for the rst 7, 10, and 12 modes of a multimode MOF. The rst seven modes are distributed in the rst seven classes, and as a result of 1,2 selection rules (23) and (24) most of the Qplmn vanish identically (dashed-dotted curve). For ten modes all the eight classes contain one mode each, with the exception of classes 3 and 4, which contain two modes. Therefore in all the allowed class combinations involving classes 3 or 4 the number of nonzero coefcients is doubled, resulting in a total of about 4000 nonzero coefcients (dashed curve), similarly to the step-index ber discussed above. Since modes 11 and 12 are also in classes 3 and 4, respectively,

Table 2. Fourier Series Representation for the Radial and Azimuthal Components of the Electric Field for the Eight Symmetry Classes of a C6v Waveguidea
Mode Class P 1 F r,q r , F ,q r , K 0

=0

1qrcos6

=0

1qrsin6

=0

2qrsin6
3qrcos6 + + 3qrcos6 4qrsin6 + + 4qrsin6 5qrcos6 + 2+ 5qrcos6 2
6qrsin6 + 2+ 6qrsin6 2

=0

2qrcos6

=0

=0

1 3qrsin6 + + 3qrsin6 1 4qrcos6 + + 4qrcos6 2 5qrsin6 + 2+ 5qrsin6 2 2


6qrcos6 + 2+ 6qrcos6 2

=0

=0

=0

=0


=0


=0

=0

7qrcos6 + 3

=0

7qrsin6 + 3

=0
a

8qrsin6 + 3

=0

8qrcos6 + 3

The longitudinal components exhibit the same symmetry as the radial components 40.

F. Poletti and P. Horak

Vol. 25, No. 10 / October 2008 / J. Opt. Soc. Am. B

1651

10000 Number of Q(1,2)> max{Q(1,2)} 8000 6000 4000 2000 0 6 10

12 modes 10 modes 7 modes

(a)

10

10 Threshold

10

40 modes of the ber in Fig. 2(b) can be converted into nearly circularly polarized modes Fc. In all cases the degree of circular polarization of the new modes, given by * 2 / dxdyF 2, was found to be greater than dxdyFc e+ c 95%, and in many cases 99%. For these approximately circularly polarized modes the polarization selection rules (19) still apply, with the crucial consequence already observed that, since one half of the modes cannot couple to the other half, the number of equations in Eq. (6) can be halved and the number of nonzero coefcients is reduced by a factor of 16. Because of the simultaneous presence of sine and cosine terms in the Fourier expansions of the new modes, the selection rule (23) does not hold for circularly polarized modes. On the other hand, each new e+ or e polarized mode obtained by combining linearly polarized modes of classes 1 and 2, 3 and 4, 5 and 6, and 7 and 8 retains the same value of K as the modes it is originated from. Choosing modes of a single circular polarization, selection rule (24) becomes
1 Qplmn =0 2 =0 Qplmn

for Kp + Kl + Km Kn for all modes p, l, m, n .

6N,

N = 0, 1, 25

Fig. 2. (Color online) (a) Distribution of the magnitude of cou1 2 and Qplmn of the rst 7, 10, and 12 modes pling coefcients Qplmn of the MOF shown in (b). The ber presents = 6 m, d / = 0.92, D = 6 m and a glass refractive index of 1.444 at a wavelength of 1.5 m.

This can be shown using the mode expansions of Table 2 and the fact that for degenerate modes in classes 3 and 4 (5 and 6) the coefcients can be written as 4q = 3q, 4q = 3q, 4q = 3q, 4q = 3q (6q = 5q, 6q = 5q, 6q = 5q, 6q = 5q). The total number of nonzero coupling coefcients for the ve e+ polarized modes resulting from the rst ten modes of the ber in Fig. 2(b) is 121, and only 67 of them are larger than 1% of the maximum coupling coefcient and therefore are likely to produce signicant nonlinear effects during pulse propagation.

5. COMPUTATIONAL COMPLEXITY
Among the various numerical techniques for solving GNLSEs, the symmetrized SSFM is probably the most widely employed because of its favorable combination of efciency, accuracy, and coding simplicity [12,13]. In the classic single-mode case, for each longitudinal step the most efcient implementation of the algorithm requires four FFTs to propagate the linear operator D1z , t in the spectral domain. Every evaluation of the delayed Raman term and of the time derivative within N1z , t responsible for self-steepening and shock formation effects [see Eq. (13)] also requires four FFTs. However, the propagation of the nonlinear term within the SSFM requires several evaluations of N1z , t, for example, four evaluations for a fourth-order RungeKutta integration such that a full SSFM step requires a total of 20 FFTs. For other implementations the exact number of FFTs may be different, but overall the computational cost associated with the FFTs is far larger than that of the remaining vector multiplications and data manipulations. Therefore the algorithm complexity is ON log2 N for a temporal grid of N points (N = 2k for integer k). For multimode bers or waveguides, Eq. (6) shows that when the number of grid points N is held constant the number of FFTs required to propagate the linear opera-

the total number of nonzero coefcients for the case of 12 modes (solid curve) is more than twice the number for ten modes. Figure 2 also shows a clear plateau in the distribution of coupling coefcients when each mode class contains at most one mode. In this case, the coupling coefcients can be quite naturally divided into signicant and marginal terms. However, if more modes per symmetry class are considered, this feature rapidly disappears. In this case one may have to choose an articial threshold in order to reduce the computational effort of numerically integrating the MM-GNLSE. In analogy to the procedure suggested for step-index bers, a further reduction of the nonzero terms can occur by selecting suitable combinations of linearly polarized modes in order to create new modes of approximately e+ or e circular polarization at every point in space. Such modes are also true solutions of Maxwells equations if created from degenerate mode pairs belonging to classes 3 and 4 (or 5 and 6), while they are only approximate solutions if created from corresponding modes in classes 1 and 2 (or 7 and 8). Despite this approximation, we have veried that by selecting appropriate mode pairs, all the rst

1652

J. Opt. Soc. Am. B / Vol. 25, No. 10 / October 2008


10
3

F. Poletti and P. Horak

tors Dp is linearly proportional to the total number of modes M and thus the complexity of this step is OMN log2 N. Similarly, also the self-steepening term is OMN log2 N under the assumption, as previously dis1,2 1 / 0. In contrast, the number of FFTs cussed, of plmn term required to evaluate the convolution of the AmA* n with the delayed Raman response ht scales quadratically with M leading to a complexity of OM2N log2 N. Apart from these FFT-computable terms, however, the full evaluation of the nonlinear terms requires the multiplication of combinations of three distinct eld envelopes [or of Al and the convolution of ht and Al, Am, and A* n ] and nally a multiplication with the coupling coefA mA * n 1,2 for any given p. This term scales as OM4N cients Qplmn if no constraints on the coupling coefcients are imposed. We therefore expect this term to dominate the overall complexity of the algorithm already for relatively modest values of M. We have already shown in Section 4 that the presence of symmetries in the ber cross section leads to an in1,2 that are either identically or apcreased number of Qplmn proximately equal to zero through selection rules (18) and (19), or (23) and (24). This reduces the number of threeterm multiplications and may therefore save a considerable amount of computational time. However, the simplication is not large enough to prevent the three-term multiplication from being the dominant factor in the overall computational cost for a sufciently large number of modes. We nally observe that since for large M the algorithm is dominated by the cost of vector multiplications, its efciency cannot be improved, as in the single-mode case, by recurring to a more efcient choice of basis functions (e.g., wavelets instead of harmonic functions [41]), which can only reduce the asymptotical complexity of the Fourier transforms. To illustrate these considerations, we have measured the time requirement for each of the main functional blocks of the multimode GNLSE algorithm in the case of the multimode MOF shown in Fig. 2. We have run SSFM simulations of the MM-GNLSE in Eq. (6), with the sim1,2 1 / 0, for 1 to 6 linearly polarized plication of plmn modes, while keeping a constant propagation distance, longitudinal step size, and grid size (213 points). The nonlinear term is propagated using a fourth-order Runge Kutta algorithm. Simulations were performed using the software package MATLAB 7.5 (The MathWorks Inc.). The results are shown in Fig. 3(a). As expected, the execution time for the linear operator and for the calculation of the self-steepening term (i.e., the time derivative within Np) is approximately linear with the number of modes. The total time required to calculate the Raman terms is approximately quadratic in M and is the dominant term for up to three modes. The computational time required by the three-term multiplications, however, scales like M3.43 and becomes the dominant contribution for M 3. Note that the ber symmetries and related selection rules (23) and (24) were accurately accounted for, which reduces the scaling of this term signicantly below the scaling like M4 that would be obtained in the absence of symmetries.

AlAmA* term n y = 1.40*x3.43 Raman term y = 4.86*x2.22 Selfsteepening term y = 0.59+6.11*x Linear operator y = 0.11+2.34*x

Execution time [sec]

10

10

10

(a)

Number of modes M

Total execution time/N [sec]

10

5 modes y = 0.00226*log2(N) y = 0.00207*N 3 modes y = 0.000663*log2(N) 1 mode y = 0.000086*log2(N)

10

(b)

10

10

Number of grid points N

Fig. 3. (Color online) (a) Execution time of the main functional blocks of numerical simulations of the MM-GNLSE as a function of the number of modes M. The markers indicate the measured times while the lines are the best t, corresponding to the equation in the legend. (b) Total execution time as a function of the grid size N for 1, 3, and 5 modes. The markers indicate the measured times and the lines the best logarithmic t, as shown in the legend.

Figure 3(b) shows how the total execution time of our MM-GNLSE implementation scales with the number of grid points N. For one and three modes, where most of the computational time is spent on FFT-based terms, the total time scales like N log2 N, as expected. In the case of ve modes the cost of three-term multiplications, scaling linearly with N, becomes predominant, but the FFT-based terms cannot be neglected yet. As a result, the scaling of the total execution time is bound between N and N log2 N. As the number of modes is further increased, the importance of the FFT-based terms becomes more and more

F. Poletti and P. Horak

Vol. 25, No. 10 / October 2008 / J. Opt. Soc. Am. B

1653

negligible, and for large values of M the numerical cost should be proportional to N. We also observed that the grid size N and the particular numerical algorithm chosen to propagate the nonlinear operator Np have some effect on the value of M at which the Raman contribution and the three-term multiplications contribution are similar: for larger grid sizes and lower accuracy algorithms the crossover point happens at larger values of M. The variation, however, is rather small and for all the different cases we have considered we have always found the three-term multiplication part becoming dominant for M 5.

ACKNOWLEDGMENTS
This work was supported by the UK Engineering and Physical Sciences Research Council (EPSRC). The authors thank David J. Richardson for valuable discussions.

REFERENCES
1. 2. R. R. Alfano and S. L. Shapiro, Observation of self-phase modulation and small-scale laments in crystals and glasses, Phys. Rev. Lett. 24, 592594 (1970). J. K. Ranka, R. S. Windeler, and A. J. Stentz, Visible continuum generation in airsilica microstructure optical bers with anomalous dispersion at 800 nm, Opt. Lett. 25, 2527 (2000). P. Russell, Photonic crystal bers, Science 299, 358362 (2003). J. M. Dudley, G. Genty, and S. Coen, Supercontinuum generation in photonic crystal ber, Rev. Mod. Phys. 78, 11351184 (2006). M. A. Foster, A. C. Turner, M. Lipson, and A. L. Gaeta, Nonlinear optics in photonic nanowires, Opt. Express 16, 13001320 (2008). T. W. Hnsch, Nobel lecture: passion for precision, Rev. Mod. Phys. 78, 12971309 (2006). J. K. Ranka, R. S. Windeler, and A. J. Stentz, Optical properties of high-delta air silica microstructure optical bers, Opt. Lett. 25, 796798 (2000). J. H. V. Price, T. M. Monro, K. Furusawa, W. Belardi, J. C. Baggett, S. Coyle, C. Netti, J. J. Baumberg, R. Paschotta, and D. J. Richardson, UV generation in a pure-silica holey ber, Appl. Phys. B 77, 291298 (2003). A. Emov, A. J. Taylor, F. G. Omenetto, J. C. Knight, W. J. Wadsworth, and P. St. J. Russell, Nonlinear generation of very high-order UV modes in microstructured bers, Opt. Express 11, 910918 (2003). T. Delmonte, M. A. Watson, E. J. ODriscoll, X. Feng, T. M. Monro, V. Finazzi, P. Petropoulos, J. H. V. Price, J. C. Baggett, W. Loh, D. J. Richardson, and D. P. Hand, Generation of mid-IR continuum using tellurite microstructured ber, in Conference on Lasers and Electro-Optics (2006), paper CTuA4. H.-G. Choi, C.-S. Kee, K.-H. Hong, J. Sung, S. Kim, D.-K. Ko, J. Lee, J.-E. Kim, and H. Y. Park, Dispersion and birefringence of irregularly microstructured ber with an elliptic core, Appl. Opt. 46, 84938498 (2007). K. J. Blow and D. Wood, Theoretical description of transient stimulated Raman scattering in optical bers, IEEE J. Quantum Electron. 25, 26652673 (1989). G. P. Agrawal, Nonlinear Fiber Optics, 3rd ed. (Academic, 2001). G. Genty, P. Kinsler, B. Kibler, and J. M. Dudley, Nonlinear envelope equation modeling of subcycle dynamics and harmonic generation in nonlinear waveguides, Opt. Express 15, 53825387 (2007). A. Hasegawa, Self-connement of multimode optical pulse in a glass ber, Opt. Lett. 5, 416417 (1980). B. Crosignani and P. Di Porto, Soliton propagation in multimode optical bers, Opt. Lett. 6, 329330 (1981). C. R. Menyuk, Stability of solitons in birefringent optical bers. I. Equal propagation amplitudes, Opt. Lett. 12, 614616 (1987). T. Brabec and F. Krausz, Nonlinear optical pulse propagation in the single-cycle regime, Phys. Rev. Lett. 78, 32823285 (1997). M. Kolesik and J. V. Moloney, Nonlinear optical pulse propagation: from Maxwells to unidirectional equations, Phys. Rev. E 70, 036604 (2004). S. Trillo and S. Wabnitz, Parametric and Raman amplication in birefringent bers, J. Opt. Soc. Am. B 9, 10611082 (1992). S. G. Murdoch, R. Leonhardt, and J. D. Harvey, Polarization modulation instability in weakly birefringent bers, Opt. Lett. 20, 866868 (1995).

3.

6. CONCLUSION
We have presented a generalized multimode nonlinear Schrdinger equation (NLSE) describing the propagation of ultrashort pulses in multimode optical bers or waveguides based on a full vectorial mode expansion. The model describes arbitrary orders of dispersion as well as Kerr and Raman nonlinearities, self-steepening effects, and, to lowest order, frequency dependencies of intermodal coupling coefcients. The equations have been shown to be equivalent to the established generalized nonlinear Schrdinger equation (GNLSE) for one or two modes. We have also investigated in detail the reduction of the number of nonlinear coupling terms in the presence of ber symmetries, in particular for step-index and circularly symmetric bers, and for hexagonally stacked microstructured bers. The most signicant simplications were found for circularly polarized modes in the weakly guiding limit, where the number of terms can be reduced by over 2 orders of magnitude and the number of coupled nonlinear equations can be halved. We have veried that when the number of modes M exceeds 4 or 5, the dominant contribution to the overall computational complexity comes from vector multiplications and that for highly multimode bers the algorithms complexity is OMxN, where N is the number of temporal grid points and x 4 if ber symmetries are taken into account (the best scaling we have observed in our simulations so far was x 3). As a result, typical multimode nonlinear simulations can be carried out on a state-of-the-art personal computer with a running time of a few hours if the number of circularly polarized modes does not exceed 810 (which is equivalent to 1620 linearly polarized modes in the weakly guiding limit). Nonlinear simulations of heavily multimode bers may require the use of dedicated multiprocessor clusters. We anticipate that the theory presented here will allow for a thorough investigation of multimode supercontinuum generation, thus explaining recent experimental results [711]. Other areas of applications may include high-power ber lasers, parametric ampliers, optical switches, or large mode area ber transmission systems. In each of these areas a detailed understanding of multimode processes may subsequently lead to the design of novel optical bers with unprecedented nonlinear properties.

4. 5. 6. 7. 8.

9.

10.

11.

12. 13. 14.

15. 16. 17. 18. 19. 20. 21.

1654 22.

J. Opt. Soc. Am. B / Vol. 25, No. 10 / October 2008 M. Lehtonen, G. Genty, H. Ludvigsen, and M. Kaivola, Supercontinuum generation in a highly birefringent microstructured ber, Appl. Phys. Lett. 82, 21972199 (2003). S. Coen, A. H. L. Chau, R. Leonhardt, J. D. Harvey, J. C. Knight, W. J. Wadsworth, and P. St. J. Russell, Supercontinuum generation by stimulated Raman scattering and parametric four-wave mixing in photonic crystal bers, J. Opt. Soc. Am. B 19, 753764 (2002). E. R. Martins, D. H. Spadoti, M. A. Romero, and B.-H. V. Borges, Theoretical analysis of supercontinuum generation in a highly birefringent D-spaced microstructured optical ber, Opt. Express 15, 1433514347 (2007). A. Tonello, S. Pitois, S. Wabnitz, G. Millot, T. Martynkien, W. Urbanczyk, J. Wojcik, A. Locatelli, M. Conforti, and C. De Angelis, Frequency tunable polarization and intermodal modulation instability in high birefringence holey ber, Opt. Express 14, 397404 (2005). J. M. Dudley, L. Provino, N. Grossard, H. Maillotte, R. S. Windeler, B. J. Eggleton, and S. Coen, Supercontinuum generation in airsilica microstructured bers with nanosecond and femtosecond pulse pumping, J. Opt. Soc. Am. B 19, 765771 (2002). D. A. Akimov, E. E. Serebryannikov, A. M. Zheltikov, M. Schmitt, R. Maksimenka, W. Kiefer, K. V. Dukelskii, V. S. Shevandin, and Y. N. Kondratev, Efcient anti-Stokes generation through phase-matched four-wave mixing in higher-order modes of a microstructure ber, Opt. Lett. 28, 19481950 (2003). T. Chaipiboonwong, P. Horak, J. D. Mills, and W. S. Brocklesby, Numerical study of nonlinear interactions in a multimode waveguide, Opt. Express 15, 90409047 (2007). P. Dupriez, F. Poletti, P. Horak, M. N. Petrovich, Y. Jeong, J. Nilsson, D. J. Richardson, and D. N. Payne, Efcient white light generation in secondary cores of holey bers, Opt. Express 15, 37293736 (2007). R. Cherif, M. Zghal, L. Tartara, and V. Degiorgio,

F. Poletti and P. Horak Supercontinuum generation by higher-order mode excitation in a photonic crystal ber, Opt. Express 16, 21472152 (2008). P. V. Mamyshev and S. V. Chernikov, Ultrashort-pulse propagation in optical bers, Opt. Lett. 15, 10761078 (1990). J. Lgsgaard, Mode prole dispersion in the generalized nonlinear Schrdinger equation, Opt. Express 15, 1611016123 (2007). B. Kibler, J. M. Dudley, and S. Coen, Supercontinuum generation and nonlinear pulse propagation in photonic crystal ber: inuence of the frequency-dependent effective mode area, Appl. Phys. B 81, 337342 (2005). R. H. Stolen and E. P. Ippen, Raman gain in glass optical waveguides, Appl. Phys. Lett. 22, 276278 (1973). A. W. Snyder, Asymptotic expressions for the eigenfunctions and eigenvalues of a dielectric or optical waveguide, IEEE Trans. Microwave Theory Tech. 17, 11301138 (1969). K. Okamoto, Fundamentals of Optical Waveguides, 2nd ed. (Academic, 2006). A. W. Snyder and J. D. Love, Optical Waveguide Theory (Kluwer, 2000). J. H. V. Price, T. M. Monro, H. Ebendorff-Heidepriem, F. Poletti, P. Horak, V. Finazzi, J. Y. Y. Leong, P. Petropoulos, J. C. Flanagan, G. Brambilla, X. Feng, and D. J. Richardson, Mid-IR supercontinuum generation from nonsilica microstructured bers, IEEE J. Sel. Top. Quantum Electron. 13, 738749 (2007). A. R. Bhagwat and A. L. Gaeta, Nonlinear optics in hollow-core photonic bandgap bers, Opt. Express 16, 50355047 (2008). P. R. McIsaac, Symmetry-induced modal characteristics of uniform waveguidesI: Summary of results, IEEE Trans. Microwave Theory Tech. 23, 421429 (1975). T. Kremp and W. Freude, Fast split-step wavelet collocation method for WDM system parameter optimization, J. Lightwave Technol. 23, 14911502 (2005).

31. 32. 33.

23.

24.

34. 35.

25.

26.

36. 37. 38.

27.

39. 40. 41.

28. 29.

30.

You might also like