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

Spectral properties of limiting solitons in optical fibers

Open Access Open Access

Abstract

It seems to be self-evident that stable optical pulses cannot be considerably shorter than a single oscillation of the carrier field. From the mathematical point of view the solitary solutions of pulse propagation equations should loose stability or demonstrate some kind of singular behavior. Typically, an unphysical cusp develops at the soliton top, preventing the soliton from being too short. Consequently, the power spectrum of the limiting solution has a special behavior: the standard exponential decay is replaced by an algebraic one. We derive the shortest soliton and explicitly calculate its spectrum for the so-called short pulse equation. The latter applies to ultra-short solitons in transparent materials like fused silica that are relevant for optical fibers.

© 2014 Optical Society of America

1. Introduction

Description of short optical pulses is usually based on the generalized nonlinear Schrödinger equation (GNLSE) [1], which yields dynamics of single-cycle [2] and even sub-cycle [3] pulses. The validity of the GNLSE is because of two reasons, (i) the standard slowly varying envelope approximation (SVEA) is relaxed by considering an unidirectional regime of pulse propagation, and (ii) the complex envelope is defined by omitting the negative-frequency components from the real field without any reference to the SVEA. On the other hand, exact analytical solutions of the GNLSE are rare and typically require either a simplified quadratic dispersion law or artificial relations between equation parameters, like in the case of Sasa-Satsuma equation.

An attractive alternative to the GNLSE is to use a special non-envelope propagation model, one of the so-called short pulse equations (SPEs). The latter are derived using a simplified dispersion law. For instance, Schäfer and Wayne [4] noticed that the relative permittivity of bulk fused silica can be precisely approximated by an extremely simple expression

ε(ω)=ns2(1ωp2/ω2),
for all frequencies that are relevant for optical solitons. As opposed by the Drude dispersion law, ns and ωp are just fit parameters. For instance, selecting them as ns = 1.4538 and ωp/(2π) = 21.78THz, one gains the relative accuracy of ≲ 0.15% between 0.9 μm and 3.6 μm (from 83THz to 333THz), i.e., for two octaves in fused silica. From the physical side ns is just a typical value of the refractive index, whereas ωp indicates at which frequency the imaginary part of the refractive index cannot be neglected. The rational approximation (1) naturally applies to the transparency region with ωωp.

In contrast to the Taylor expansion of the dispersion relation in the GNLSE, i.e., a local approach that is very accurate but only within the convergence radius at the carrier frequency [5], Eq. (1) pretends to be a global approximation. First, it is bounded for ω → ∞, as opposed by the unphysical behavior of the GNLSE dispersion. Second, one can apply the Kramers Kronig relations to demonstrate that Eq. (1) is a kind of universal representation of the anomalous dispersion domain for any transparent material [6]. Third, simplified propagation equations yield useful analytic solutions for short optical solitons, in some cases even a full solution of the SPE is available [7]. These solutions can be used to test more general models.

In what follows we use the dispersion relation (1), to address the following fundamental question: what happens with an optical soliton when its duration approaches a single oscillation of the carrier field? The known analytical [6, 8, 9] and numerical solutions [10] suggest the following feature: a singular cusp develops at the soliton top, preventing the soliton from being too short. Such peaked solitons have originally been found outside optics, for the shallow water waves [11].

In the frequency domain the limiting shortest soliton is characterized by the following property: an exponential decay of the spectral power is replaced by a rational one. This behavior is of special interest because pulse spectrum is easier to measure than the actual value of the electric field inside the pulse. In the following we first revisit the derivation of the SPE and its limiting soliton [6, 12] and then explicitly calculate corresponding spectra.

2. Complex short pulse equation

In this section we outline the derivation of the real [4] and complex [6, 12] SPEs and demonstrate that the transition to a more simple complex SPE is dictated by the dispersion law (1).

Propagation of a short pulse in a one-dimensional setting, e.g., in a single-mode polarization-preserving optical fiber, is approximated by a nonlinear wave equation for the electric field

t2(ε^E+χ(3)E3)c2z2E=0,whereε^[ωEωeiωt]=ωε(ω)Eωeiωt.
The nonlinear susceptibility of the third order χ(3) is taken constant. Using Eq. (1) one derives the following self-consistent equation
ns2(t2E+ωp2E)c2z2E+χ(3)t2(E3)=0,
with the following dispersion law β(ω)=(ns/c)(ω2ωp2)1/2 for a linear ei(βzωt) wave.

Recall, that typical carrier frequencies of optical solitons are considerably larger than the fit parameter ωp from Eq. (1), as discussed in the Introduction. Therefore β(ω)(ns/c)[ωωp2/(2ω)] and one concludes that the third harmonics generation (THG) process is a non-resonant one for our system. Indeed, the corresponding resonance conditions ω1 + ω2 + ω3 = ω and ω11+ω21+ω31=ω1 are incompatible for any four positive frequencies.

The SPE results from the asymptotic expansion of Eq. (3) with respect to μ = ωp0 ≪ 1, e.g., μ = 0.1 for a standard carrier wave at 1.4 μm. Actually, we will keep O(μ2) terms and neglect O(μ4) ones. To derive the SPE we first consider a reference carrier wave (the SVEA is not assumed) ei(β0zω0t) and calculate the phase

ω0tβ0z=ω0[tnscz(1μ2)1/2]=ω0(tnscz)+μ2nsω02cz+O(μ4).
Motivated by this expression we introduce the normalized field F(ζ, τ) and assume the following scaling of the new variables
τ=ω0(tnscz),ζ=μ2nsω0cz,(3χ(3)8)1/2E(z,t)=μnsF(ζ,τ).
A straightforward calculation yields that Eq. (3) takes the following form
2ζτF+F+83τ2(F3)=μ2ζ2F.
After neglecting the last term one obtains the real SPE as derived in [4]. One should mention that a more general SPE with an additional term accounting for transition to the normal dispersion domain first appeared in [13]. To derive the complex SPE we introduce a complex-valued electric field (ζ, τ) such that by construction
ζτ+12+13τ2(3||2+3)=0.
It is easy to see that F = ( + *)/2 is a valid solution of the real SPE. The clear advantage of the complex representation is that the self-phase modulation (SPM) term and the THG term are now separated. Neglecting the non-resonant THG term we obtain the complex SPE
ζτ+12+τ2(||2)=0.
Another physical situation that directly leads to the complex SPE is for a wave with circular polarization. Two components of the real electric field are just combined into a singe complex field and one obtains the complex SPE (8) without referring to the real equation [6, 12].

To conclude this section we note that strictly speaking Eq. (8) presupposes that contains mainly positive (and * mainly negative) frequencies. That is, the initial condition (0, τ) is chosen to be exactly the positive frequency part of F(0, τ) in accord with the most general definition of the complex envelope [1,14]. In the course of pulse propagation, the complex field (z, τ) accrues some small negative-frequency part generated by the SPM term, however, this process is non-resonant such that the negative frequencies are not further amplified.

3. Solitary solutions of the complex SPE

For a linear ∼ exp[i(ℵζντ)] wave, in which the dimensionless frequency ν corresponds to the physical frequency νω0 and the reference wave corresponds to ν = 1, Eq. (8) yields = −1/(2ν). Therefore the phase and group velocities read Vph = −2ν2 and Vgr = 2ν2, in particular, Vph + Vgr = 0. Although both velocities slightly change for a solitary wave, they still have equal magnitudes. Therefore we introduce a positive parameter γ such that

Vphasenonlinear=γVph=2γν2andVgroupnonlinear=γVgr=2γν2,
and assume that soliton’s phase and amplitude depend on τζ/(γVph) and τζ/(γVgr) respectively. In accord with Eq. (9) we then try the following substitution
(ζ,τ)=f(ντζ2γν)exp[i(ντ+ζ2γν)],
where f (ξ) is the complex shape function that depends also on γ and ν. Substituting (ζ, τ) into Eq. (8) we obtain the following ordinary differential equation for f (ξ)
(f2γν2|f|2f)(γ1)f+4iγν2(|f|2f)+2γν2|f|2f=0,
where prime denotes derivative with respect to ξ. At the soliton tails, where f (ξ) → 0, we have f″ = (γ − 1) f, such that localized solitary solutions require γ > 1 and the soliton duration is proportional to (γ − 1)−1/2.

We now split the amplitude and the phase by setting γ1/2νf (ξ) = a(ξ)e(ξ) and obtain two real differential equations for a(ξ) and ϕ(ξ). The phase equation can be integrated and yields

ϕ=(34a2)a2(12a2)2.
The amplitude equation A″ − (γ − 1)aϕ′2A − 4ϕ′a3 + 2a3 = 0 with Aa − 2a3 can be multiplied by A′ and then integrated as well. Restricting ourselves to the localized solutions we obtain a “mechanical” equation for some effective potential U(a)
a2+U(a)=0,U(a)=(γ1)(13a2)a2(16a2)2+(17a2+12a4)a4(12a2)2(16a2)2.

An exemplary shape of U(a) for γ slightly smaller than 9/8 is shown in Fig. 1(a). The target soliton starts from the equilibrium position a = 0 for ξ → −∞ and returns to it for ξ → +∞. Physically acceptable solutions require U(a) ≤ 0. The solution of the latter inequality that is “connected” to the a = 0 state is given by

0a[4γ3(98γ)1/28γ]1/2,
the right limit is labelled by the red point in Fig. 1(a). The target solitary solution exists for 1 ≤ γ ≤ 9/8. Classical solitons appear for γ = 1 + ε2 with ε ≪ 1, then a = O(ε) and in the leading term Eq. (13) yields that a(ξ) = ε/cosh(εξ). As γ increases the soliton becomes shorter, the final limiting soliton is obtained for γ → 9/8 and is yielded by the equations
a2=a2(13a2)8(12a2)2witha(0)=16aeB(a)=Λe|ξ|22,
B(a)[23(13a2)1/2ln[1+(13a2)1/2]]0a,ΛeB(16)60.3935.

The phase dependence of the limiting soliton is calculated from Eq. (12). Without loss of generality we take ϕ(0) = 0, then ϕ(−ξ) = −ϕ(ξ). The shape of ϕ(ξ) is shown in Fig. 2(a). It is of interest that the limiting values ϕ(±∞) = ±ϕ can be calculated analytically from (12) and (15)

ϕ=0ϕ(ξ)dξ=0(34a2)a2(12a2)2(dξdada)=4(21)3arcsin13,
the quantity ϕ ≈ −0.8921 appears later in the equation for the soliton spectrum.

 figure: Fig. 1

Fig. 1 (a) An exemplary potential U(a) from Eq. (13) for γ = 9/8 − δ with δ = 10−3. The red point labels the upper value of a(ξ) from the inequality (14). As δ → 0, an infinite wall is formed at a=1/6, resulting in cusp formation at the top of the soliton. (b) Shape of the shortest soliton calculated from Eq. (15).

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 (a) Phase of the limiting soliton and (b) its power spectrum. Blue line: numerical solution, red line: expression (22). The two auxiliary thin lines show the spectrum of the fundamental soliton (dashed), and the Ω−4 power law (solid).

Download Full Size | PDF

4. The limiting spectrum

In this section we explicitly calculate the spectrum of the shortest soliton. At first glance, the key Eq. (15) looks too complicated and numerical treatment of the soliton spectrum seems to be the only choice. Actually, this is not the case. Our analytic approach is based on the observation that for all relevant amplitudes, 0a1/6, the factor eB(a) from Eq. (15) is a slow function that is close to 1. This happens just because B(0) = 0 and B(1/6)0.037. Actually a(ξ)Λe|ξ|/(22) is already a very good approximation to the full solution of Eq. (15) that is shown in Fig. 1(b). As a natural generalization we consider ξ > 0 and expand the amplitude a(ξ) and phase ϕ(ξ) in a power series with respect to X=Λeξ/(22).

With the help of any computer algebra system it is easy to obtain that

X=aeB(a)=aa34+a58+49a7192+a=X+X34+X51661X7192+
Inserting a(X) in Eq. (12) and performing direct integration we get the full solution for ξ > 0
ϕ=ϕ+32X2+19X422+457X6242+,
aeiϕ=eiϕ(X+1+12i24X3+143+88i216X5+),
where X=Λeξ/(22), Λ and ϕ are yielded by Eq. (16) and Eq. (17). Recall that the solution for ξ < 0 follows from the fact that a(ξ) and ϕ(ξ) are even and odd functions respectively.

Finally Eq. (20), which can be extended to any power, is used to obtain the spectrum

(aeiϕ)Ω=a(ξ)eiϕ(ξ)eiΩξdξ=0a(ξ)eiϕ(ξ)eiΩξdξ+c.c.,
where Ω refers to the normalized frequency. The result reads
(aeiϕ)Ω=ΛR1+S1Ω1+8Ω2+Λ3R3+S3Ω9+8Ω2+Λ5R5+S5Ω25+8Ω2+,
where R1=42C, S1 = −16S, R3=32C72S, S3=482C4S, R5=(715/4)2C220S, S5=882C+143S, and we set C = cosϕ and S = sinϕ for the sake of brevity. For Ω ≫ 1 Eq. (22) should be replaced by the power law, (ae)Ω ∼ Ω−2, which should be observed for any generic cusp solution [Fig. 2(b)]. Actually, such frequencies are beyond validity of Eq. (1) and SPE, such that Eq. (22) is the only reasonable result for the soliton spectrum.

5. Conclusions

Our main result is Eq. (22) for the spectrum of the shortest soliton. As expected, the spectral power has an algebraic decay, as opposed by the exponential one for ordinary solitons. The spectral shape is non-trivial, such a function is difficult to reconstruct from a purely numerical solution. Figure 2(b) shows predictions of Eq. (22) and compares them with the numerical results for the soliton spectrum. The agreement is reasonably good for all frequencies of interest and can be easily improved by calculating further terms in Eq. (22).

Acknowledgments

Sh.A. gratefully acknowledges support by The Einstein Center for Mathematics Berlin.

References and links

1. R. W. Boyd, Nonlinear Optics (Academic, New York, 2008), 3rd ed.

2. T. Brabec and F. Krausz, “Nonlinear optical pulse propagation in the single-cycle regime,” Phys. Rev. Lett. 78, 3282–3285 (1997). [CrossRef]  

3. G. Genty, P. Kinsler, B. Kibler, and J. M. Dudley, “Nonlinear envelope equation modeling of sub-cycle dynamics and harmonic generation in nonlinear waveguides,” Opt. Express 15, 5382–5387 (2007). [CrossRef]   [PubMed]  

4. T. Schäfer and C. E. Wayne, “Propagation of ultra-short optical pulses in cubic nonlinear media,” Physica D 196, 90–105 (2004). [CrossRef]  

5. K. E. Oughstun and H. Xiao, “Failure of the quasimonochromatic approximation for ultrashort pulse propagation in a dispersive, attenuative medium,” Phys. Rev. Lett. 78, 642–645 (1997). [CrossRef]  

6. S. A. Skobelev, D. V. Kartashov, and A. V. Kim, “Few-optical-cycle solitons and pulse self-compression in a Kerr medium,” Phys. Rev. Lett. 99, 203902 (2007). [CrossRef]  

7. A. Sakovich and S. Sakovich, “The short pulse equation is integrable,” J. Phys. Soc. Jpn. 74, 239–241 (2005). [CrossRef]  

8. A. Sakovich and S. Sakovich, “Solitary wave solutions of the short pulse equation,” J. Phys. A 39, L361–L367 (2006). [CrossRef]  

9. S. Amiranashvili, U. Bandelow, and N. Akhmediev, “Dispersion of nonlinear group velocity determines shortest envelope solitons,” Phys. Rev. A 84, 043834 (2011). [CrossRef]  

10. S. Amiranashvili, U. Bandelow, and N. Akhmediev, “Few-cycle optical solitary waves in nonlinear dispersive media,” Phys. Rev. A 87, 013805 (2013). [CrossRef]  

11. R. Camassa and D. Holm, “An integrable shallow water equation with peaked solitons,” Phys. Rev. Lett. 71, 1661–1664 (1993). [CrossRef]   [PubMed]  

12. S. Amiranashvili, A. G. Vladimirov, and U. Bandelow, “Solitary-wave solutions for few-cycle optical pulses,” Phys. Rev. A 77, 063821 (2008). [CrossRef]  

13. S. A. Kozlov and S. V. Sazonov, “Nonlinear propagation of optical pulses of a few oscillations duration in dielectric media,” JETP 84, 221–228 (1997). [CrossRef]  

14. D. Gabor, “Theory of communication,” Journal of the Institute of Electrical Engineers 93, 429–457 (1946).

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 (2)

Fig. 1
Fig. 1 (a) An exemplary potential U(a) from Eq. (13) for γ = 9/8 − δ with δ = 10−3. The red point labels the upper value of a(ξ) from the inequality (14). As δ → 0, an infinite wall is formed at a = 1 / 6, resulting in cusp formation at the top of the soliton. (b) Shape of the shortest soliton calculated from Eq. (15).
Fig. 2
Fig. 2 (a) Phase of the limiting soliton and (b) its power spectrum. Blue line: numerical solution, red line: expression (22). The two auxiliary thin lines show the spectrum of the fundamental soliton (dashed), and the Ω−4 power law (solid).

Equations (22)

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

ε ( ω ) = n s 2 ( 1 ω p 2 / ω 2 ) ,
t 2 ( ε ^ E + χ ( 3 ) E 3 ) c 2 z 2 E = 0 , where ε ^ [ ω E ω e i ω t ] = ω ε ( ω ) E ω e i ω t .
n s 2 ( t 2 E + ω p 2 E ) c 2 z 2 E + χ ( 3 ) t 2 ( E 3 ) = 0 ,
ω 0 t β 0 z = ω 0 [ t n s c z ( 1 μ 2 ) 1 / 2 ] = ω 0 ( t n s c z ) + μ 2 n s ω 0 2 c z + O ( μ 4 ) .
τ = ω 0 ( t n s c z ) , ζ = μ 2 n s ω 0 c z , ( 3 χ ( 3 ) 8 ) 1 / 2 E ( z , t ) = μ n s F ( ζ , τ ) .
2 ζ τ F + F + 8 3 τ 2 ( F 3 ) = μ 2 ζ 2 F .
ζ τ + 1 2 + 1 3 τ 2 ( 3 | | 2 + 3 ) = 0 .
ζ τ + 1 2 + τ 2 ( | | 2 ) = 0 .
V phase nonlinear = γ V ph = 2 γ ν 2 and V group nonlinear = γ V gr = 2 γ ν 2 ,
( ζ , τ ) = f ( ν τ ζ 2 γ ν ) exp [ i ( ν τ + ζ 2 γ ν ) ] ,
( f 2 γ ν 2 | f | 2 f ) ( γ 1 ) f + 4 i γ ν 2 ( | f | 2 f ) + 2 γ ν 2 | f | 2 f = 0 ,
ϕ = ( 3 4 a 2 ) a 2 ( 1 2 a 2 ) 2 .
a 2 + U ( a ) = 0 , U ( a ) = ( γ 1 ) ( 1 3 a 2 ) a 2 ( 1 6 a 2 ) 2 + ( 1 7 a 2 + 12 a 4 ) a 4 ( 1 2 a 2 ) 2 ( 1 6 a 2 ) 2 .
0 a [ 4 γ 3 ( 9 8 γ ) 1 / 2 8 γ ] 1 / 2 ,
a 2 = a 2 ( 1 3 a 2 ) 8 ( 1 2 a 2 ) 2 with a ( 0 ) = 1 6 a e B ( a ) = Λ e | ξ | 2 2 ,
B ( a ) [ 2 3 ( 1 3 a 2 ) 1 / 2 ln [ 1 + ( 1 3 a 2 ) 1 / 2 ] ] 0 a , Λ e B ( 1 6 ) 6 0.3935 .
ϕ = 0 ϕ ( ξ ) d ξ = 0 ( 3 4 a 2 ) a 2 ( 1 2 a 2 ) 2 ( d ξ d a d a ) = 4 ( 2 1 ) 3 arcsin 1 3 ,
X = a e B ( a ) = a a 3 4 + a 5 8 + 49 a 7 192 + a = X + X 3 4 + X 5 16 61 X 7 192 +
ϕ = ϕ + 3 2 X 2 + 19 X 4 2 2 + 457 X 6 24 2 + ,
a e i ϕ = e i ϕ ( X + 1 + 12 i 2 4 X 3 + 143 + 88 i 2 16 X 5 + ) ,
( a e i ϕ ) Ω = a ( ξ ) e i ϕ ( ξ ) e i Ω ξ d ξ = 0 a ( ξ ) e i ϕ ( ξ ) e i Ω ξ d ξ + c . c . ,
( a e i ϕ ) Ω = Λ R 1 + S 1 Ω 1 + 8 Ω 2 + Λ 3 R 3 + S 3 Ω 9 + 8 Ω 2 + Λ 5 R 5 + S 5 Ω 25 + 8 Ω 2 + ,
Select as filters


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