You are on page 1of 11

Effect of initial frequency chirp on Airy pulse

propagation in an optical fiber


Lifu Zhang,1, 2, * Kun Liu,1 Haizhe Zhong,1 Jinggui Zhang,1, 3 Ying Li,1
and Dianyuan Fan1
1
SZU-NUS Collaborative Innovation Center for Optoelectronic Science & Technology, Key Laboratory of
Optoelectronic Devices and Systems of Ministry of Education and Guangdong Province, College of Optoelectronic
Engineering, Shenzhen University, Shenzhen 518060, China
2
Department of Electronics and Communication Engineering, Changsha University, Changsha 410003, China
3
School of Information Technology, Hunan First Normal College, Changsha 410205, China
*
zhanglifu68@hotmail.com

Abstract: We study both analytically and numerically the propagation


dynamics of an initially chirped Airy pulse in an optical fiber. It is found
that the linear propagation of an initially chirped Airy pulse depends
considerably on whether the second-order dispersion parameter 2 and
chirp C have the same or opposite signs. For 2 C < 0 , the chirped Airy
pulse first undergoes an initial compression phase, then reaches a breakup
area as depending on the values of C , and then experiences a lossy
inversion transformation such that it continues to propagate with an
opposite acceleration. The chirped Airy pulse is always dispersed during
propagation in the case of 2 C > 0 . The impact of truncation coefficient
and Kerr nonlinearity on the chirped Airy pulse propagation is also
disclosed separately.
2015 Optical Society of America
OCIS codes: (060.4370) Nonlinear optics, fibers; (190.0190) Nonlinear optics; (190.5530)
Pulse propagation and temporal solitons.

References and links


1. M. V. Berry and N. L. Balazs, Nonspreading wave packets, Am. J. Phys. 47(3), 264267 (1979).
2. G. A. Siviloglou and D. N. Christodoulides, Accelerating finite energy Airy beams, Opt. Lett. 32(8), 979981
(2007).
3. G. A. Siviloglou, J. Broky, A. Dogariu, and D. N. Christodoulides, Observation of accelerating Airy beams,
Phys. Rev. Lett. 99(21), 213901 (2007).
4. P. Polynkin, M. Kolesik, J. Moloney, G. Siviloglou, and D. N. Christodoulides, Extreme nonlinear optics with
ultra-intense self-bending Airy beams, Opt. and Photon. News 21(9), 3843 (2010).
5. Y. Hu, G. A. Siviloglou, P. Zhang, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, Self-Accelerating Airy
Beams: Generation, Control, and Applications, in Nonlinear Photonics and Novel Optical Phenomena, Z. Chen
and R. Morandotti, eds, (Springer, New York, 2012), 146.
6. C. M. A. Bandres, I. Kaminer, M. Mills, B. M. Rodrguez-Lara, E. Greenfield, M. Segev, and D. N.
Christodoulides, Accelerating Optical Beams, Opt. and Photon. News 24(6), 3037 (2013).
7. J. Broky, G. A. Siviloglou, A. Dogariu, and D. N. Christodoulides, Self-healing properties of optical Airy
beams, Opt. Express 16(17), 1288012891 (2008).
8. P. Polynkin, M. Kolesik, J. V. Moloney, G. A. Siviloglou, and D. N. Christodoulides, Curved plasma channel
generation using ultraintense Airy beams, Science 324(5924), 229232 (2009).
9. P. Polynkin, M. Kolesik, and J. Moloney, Filamentation of femtosecond laser Airy beams in water, Phys. Rev.
Lett. 103(12), 123902 (2009).
10. J. Baumgartl, M. Mazilu, and K. Dholakia, Optically mediated particle clearing using Airy wavepackets, Nat.
Photonics 2(11), 675678 (2008).
11. P. Zhang, J. Prakash, Z. Zhang, M. S. Mills, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, Trapping and
guiding microparticles with morphing autofocusing Airy beams, Opt. Lett. 36(15), 28832885 (2011).
12. T. Vettenburg, H. I. Dalgarno, J. Nylk, C. Coll-Llad, D. E. Ferrier, T. imr, F. J. Gunn-Moore, and K.
Dholakia, Light-sheet microscopy using an Airy beam, Nat. Methods 11(5), 541544 (2014).
13. S. Jia, J. C. Vaughan, and X. Zhuang, Isotropic three-dimensional super-resolution imaging with a self-bending
point spread function, Nat. Photonics 8(4), 302306 (2014).

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2566
14. P. Rose, F. Diebel, M. Boguslawski, and C. Denz, Airy beam induced optical routing, Appl. Phys. Lett.
102(10), 101101 (2013).
15. D. Abdollahpour, S. Suntsov, D. G. Papazoglou, and S. Tzortzakis, Spatiotemporal airy light bullets in the
linear and nonlinear regimes, Phys. Rev. Lett. 105(25), 253901 (2010).
16. A. Chong, W. Renninger, D. N. Christodoulides, and F. W. Wise, Airy-Bessel wave packets as versatile linear
light bullets, Nat. Photonics 4(2), 103106 (2010).
17. P. Panagiotopoulos, D. G. Papazoglou, A. Couairon, and S. Tzortzakis, Sharply autofocused ring-Airy beams
transforming into non-linear intense light bullets, Nat Commun 4, 2622 (2013).
18. I. Kaminer, Y. Lumer, M. Segev, and D. N. Christodoulides, Causality effects on accelerating light pulses,
Opt. Express 19(23), 2313223139 (2011).
19. P. Panagiotopoulos, D. Abdollahpour, A. Lotti, A. Couairon, D. Faccio, D. G. Papazoglou, and S. Tzortzakis,
Nonlinear propagation dynamics of finite-energy Airy beams, Phys. Rev. A 86(1), 013842 (2012).
20. A. Lotti, D. Faccio, A. Couairon, D. G. Papazoglou, P. Panagiotopoulos, D. Abdollahpour, and S. Tzortzakis,
Stationary nonlinear Airy beams, Phys. Rev. A 84(2), 021807 (2011).
21. R. Driben, Y. Hu, Z. Chen, B. A. Malomed, and R. Morandotti, Inversion and tight focusing of Airy pulses
under the action of third-order dispersion, Opt. Lett. 38(14), 24992501 (2013).
22. S. Wang, D. Fan, X. Bai, and X. Zeng, Propagation dynamics of Airy pulses in optical fibers with periodic
dispersion modulation, Phys. Rev. A 89(2), 023802 (2014).
23. I. M. Besieris and A. M. Shaarawi, Accelerating airy wave packets in the presence of quadratic and cubic
dispersion, Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 78(4), 046605 (2008).
24. Y. Fattal, A. Rudnick, and D. M. Marom, Soliton shedding from Airy pulses in Kerr media, Opt. Express
19(18), 1729817307 (2011).
25. L. Zhang, J. Zhang, Y. Chen, A. Liu, and G. Liu, Dynamic propagation of finite-energy Airy pulses in the
presence of higher-order effects, J. Opt. Soc. Am. B 31(4), 889897 (2014).
26. Y. Hu, M. Li, D. Bongiovanni, M. Clerici, J. Yao, Z. Chen, J. Azaa, and R. Morandotti, Spectrum to distance
mapping via nonlinear Airy pulses, Opt. Lett. 38(3), 380382 (2013).
27. L. Zhang, H. Zhong, Y. Li, and D. Fan, Manipulation of Raman-induced frequency shift by use of asymmetric
self-accelerating Airy pulse, Opt. Express 22(19), 2259822607 (2014).
28. C. Ament, P. Polynkin, and J. V. Moloney, Supercontinuum generation with femtosecond self-healing Airy
pulses, Phys. Rev. Lett. 107(24), 243901 (2011).
29. Y. Zhang, M. Beli, Z. Wu, H. Zheng, K. Lu, Y. Li, and Y. Zhang, Soliton pair generation in the interactions of
Airy and nonlinear accelerating beams, Opt. Lett. 38(22), 45854588 (2013).
30. L. Zhang and H. Zhong, Modulation instability of finite energy Airy pulse in optical fiber, Opt. Express
22(14), 1710717115 (2014).
31. X. D. Cao, G. P. Agrawal, and C. J. McKinstrie, Self-focusing of chirped optical pulses in nonlinear dispersive
media, Phys. Rev. A 49(5), 40854092 (1994).
32. L. Berg, J. J. Rasmussen, E. A. Kuznetsov, E. G. Shapiro, and S. K. Turitsyn, Self-focusing of chirped optical
pulses in media with normal dispersion, J. Opt. Soc. Am. B 13(9), 18791891 (1996).
33. L. Zhang, X. Fu, J. Deng, J. Zhang, S. Wen, and D. Fan, Role of chirp in spatiotemporal instability of
broadband pulsed laser, J. Opt. 13(1), 015203 (2011).
34. J. Kasparian, R. Sauerbrey, D. Mondelain, S. Niedermeier, J. Yu, J.-P. Wolf, Y.-B. Andr, M. Franco, B. Prade,
S. Tzortzakis, A. Mysyrowicz, M. Rodriguez, H. Wille, and L. Wste, Infrared extension of the super
continuum generated by femtosecond terawatt laser pulses propagating in the atmosphere, Opt. Lett. 25(18),
13971399 (2000).
35. V. Kartazaev and R. R. Alfano, Supercontinuum generated in calcite with chirped femtosecond pulses, Opt.
Lett. 32(22), 32933295 (2007).
36. R. Nuter, S. Skupin, and L. Berg, Chirp-induced dynamics of femtosecond filaments in air, Opt. Lett. 30(8),
917919 (2005).
37. A. Couairon and A. Mysyrowicz, Femtosecond filamentation in transparent media, Phys. Rep. 441(24), 47
189 (2007).
38. J. Park, J. H. Lee and C. H. Nam, Laser chirp effect on femtosecond laser filamentation generated for pulse
compression, Opt. Express 16(7), 44654470 (2008).
39. C. Milin, A. Jarnac, Y. Brelet, V. Jukna, A. Houard, A. Mysyrowicz, and A. Couairon, Effect of input pulse
chirp on nonlinear energy deposition and plasma excitation in water, J. Opt. Soc. Am. B 31(11), 28292837
(2014).
40. J. D. McMullen, Chirped-pulse compression in strongly dispersive media, J. Opt. Soc. Am. 67(11), 15751578
(1977).
41. M. Miyagi and S. Nishida, Pulse spreading in a single-mode fiber due to third-order dispersion, Appl. Opt.
18(5), 678682 (1979).
42. G. P. Agrawal, Nonlinear Fiber Optics, 4ed. (Academic Press, 2007).

1. Introduction
The original Airy wave packet was first found in the context of quantum mechanics by Berry
and Balazs more than 30 years ago, as a solution to the Schrdinger equation for a free

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2567
particle [1]. The Airy wave packet was not experimentally realized in the optical domain until
Christodoulides et al. made a great breakthrough in 2007. They first introduced the concept of
finite-energy Airy beams [2], and then performed the first experiments [3]. Since then, Airy
beams, a kind of nonspreading beams that exhibit not only propagation invariance but also
transverse acceleration with respect to the propagation direction, have attracted a great deal of
interest in recent years [46]. The Airy beams not only have the characteristics of quasi-
diffraction-free and self-reconstruction but also have an additional interesting feature of
transverse acceleration, ie., its trajectory follows a parabolic curve during propagation [2, 3,
7]. Those features have been employed for various applications, such as curved filament
generation [8, 9], optical trapping particle [10, 11], high resolution microscopy [12, 13], all-
optical routing [14], light bullets generation [1517].
Recent works introduced further pulsed version of Airy beams to optics by exploiting the
analogy between diffraction in space and dispersion in time. These so-called Airy pulses
have attracted growing attention, owing to their remarkable properties of quasi-nondispersive,
self-healing and self-bending their dominant intensity peaks, similar to the unique properties
of Airy beam. However, there is an important physical difference between spatial and
temporal accelerations: although both have the Airy shape, a spatial accelerating beam bends
its trajectory in space, whereas only the acceleration of Airy pulse corresponds to a change in
the velocity of the intensity peak of the pulse that manifests as self-acceleration or self-
deceleration depending on its tails behind or in front of the main peak [18]. In addition, the
nonlinear propagation dynamics of spatial Airy beams was investigated and shown some
novel behaviors, such as the existence of an additional class of stationary accelerating Airy
wave forms [19, 20]. This has stimulated great interest in the problem of Airy pulse
propagation from linear to nonlinear regimes [1517, 2129]. Under the action of Kerr
nonlinearity, soliton will be shed during the process of the Airy pulse propagation [24]. The
impact of high-order linear and nonlinear effects on Airy pulse propagation has also been
reported [25]. Interestingly, it has been used to experimentally realize the linear light bullets,
showing that they are capable of healing the dispersion and nonlinearity induced distortion
[15, 16]. Airy pulses are also exploited to not only manipulate the Raman-induced frequency
effects [27] but also to control super-continuum generation in highly nonlinear fibers [28].
Soliton pair was generated and engineered through the interaction of Airy pulse [29]. Zhang
et al. investigated the impact of truncation coefficient on the modulation instability of Airy
pulse [30].
The previous works on Airy pulse propagation did not consider the effects of initial
frequency chirping. In practice, the laser system based on the chirped pulse amplification
technology involves chirp in pulse generation, propagation and amplification. Consequently,
the pulses emitted from laser sources are often chirped. Frequency chirp can also be imposed
externally and is expected to engineer the laser pulse propagation. It is shown that the
frequency chirp can influence laser self-focusing significantly [3133]. Initial requency chirp
have also been used to control supercontinuum generation [34, 35], filamentation [36, 37] and
pulse compression [38]. More recently, Milin et al. reported the electron density and energy
deposition in filament channel can be tuned by changing input chirp of laser pulse [39].
Moreover, we should also consider the effect of initial frequency chirp on Airy pulse
propagation. This subject, to the best of our knowledge, remains however unexplored. We are
still missing a clear understanding of the role of a chirped phase in Airy pulse propagation. In
this paper we are devoted to the study of propagation properties of Airy pulse imposed an
initial frequency chirp, and disclosed the novel characteristics.
2. Theoretical model
The propagation of optical pulses in an optical fiber can be described by the well-known
nonlinear Schrodinger equation (NLSE). To simplify the model and broaden the applicability
of the results, we normalize all the variables including the field that is normalized so that its

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2568
peak input value is unity. The coordinates are normalized as follows: temporal coordinate T
is normalized to the incident pulse width T0 , propagation distance Z is measured in units of
the dispersion length LD = T02 2 , where 2 is the group velocity dispersion (GVD)
parameter. The normalized NLSE then takes the form [40]
U 1 2U 2
i + s + N 2 U U = 0. (1)
Z 2 T 2

Here the parameter N = T02 P0 2 , represents the strength of the Kerr nonlinearity, where
P0 and are the input peak power and the nonlinear coefficient respectively. s = +1 ( s = 1)
denotes anomalous (normal) GVD. It must be pointed out that, for Airy pulse with multipeak
structure, the width of the main lobe of Airy pulse T0 is usually used as a temporal scale [15].

3. Analytical results of the linear propagation


The linear propagation of Airy pulse is studied by setting N = 0 in Eq. (1). U ( Z , T ) satisfies
the following linear partial differential equation:
U 1 2U
i + s = 0. (2)
Z 2 T 2
Equation (2) is readily solved by use of the Fourier-transform method. The general solution of
Eq. (2) is give by

( 0, ) exp i s Z exp ( iT ) d ,
2
1
U ( Z ,T ) = U (3)
2 2

where U ( 0, ) is the Fourier transform of the incident field at Z = 0 and is obtained by using

U ( 0, ) = U ( 0, T ) exp ( iT ) dT . (4)

The unchirped Airy pulse, U (T , Z = 0 ) = Ai (T ) exp ( aT ) , is well documented in Ref [2]. Its
Fourier spectrum is given by
1 i
U ( , Z = 0 ) = exp ( a 2 ) exp ( 3 3a 2 ia 3 ) . (5)
2 3
Equation (5) apparently shows that temporal Airy pulse can be obtained by imposing a cubic
frequency modulation on a pulse spectrum with Gaussian distribution. Substituting Eq. (5)
into Eq. (3), we find

Z2 aZ 2 i s 3Z 3
U ( Z , T ) = Ai T + isaZ exp aT sa 2 Z sTZ (6)
4 2 2 6
In fact, the Airy pulse has been reported as early as the 1970s as a result of the propagation of
Gaussian pulse through a fiber with third-order dispersion [40, 41]. When s = 1 , Eq. (6) is
consist with the expression for spatial Airy beam [2].
For chirped Airy pulse U (T , Z = 0 ) = Ai (T ) exp ( aT ) exp ( iCT 2 ) , its Fourier spectrum
is given by

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2569
+ ia 1 a
U ( 0, ) = Ai exp 2 ( 4C 1)
iC 2C 16C 2 8C
(7)
2 1 a 2
exp i + 2
4C 8C 96C 3 4C

Equation (7) shows the Fourier spectrum of chirped Airy pulse exhibits Airy distribution as
well. Figure 1 depicts the amplitude of spectrum U of such a pulse with two different values
of chirp at Z = 0 . The analytical results (red dashdot lines) are well agreement with
numerical results (black solid lines). Owning to the linear propagation, the spectral shape
does not change during whole propagation. Substituting Eq. (7) into Eq. (3), we obtain the
analytical solution

1 T Z2 saZ aT aZ 2
U ( Z ,T ) = Ai + i exp 2
, (8)
42 2
with = 1 s 2CZ .

Fig. 1. Comparison of input spectra for chirped Airy pulse between analytical (black solid
lines) and numerical (red dash lines) results for two values of chirp parameter (a) C = 0.2
and (b) C = 0.2 .

Figure 2 shows the positively (left column) and negatively (right column) chirped Airy
pulse shapes at different propagation distances by direct calculation according to analytical
expression (Eq. (8)) in the anomalous dispersion regime. The linear propagation of positively
chirped Airy pulse shown in Fig. 2(a)-2(g) displays an interesting process. It first undergoes

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2570
slight compression and reaches a focal point; then its Airy pattern breaks down. After passing
through such breakdown area, however, the pulses reconstruct a new Airy pattern with
opposite acceleration. While it can be clearly seen from Fig. 2(a1)-2(g1) that the negatively
chirped Airy pulses are always dispersed with an increasing propagation distance. These
features can be understood from Eq. (8). It shows the chirped Airy pulse evolution keeps the
airy structure only if is far from zero. In the anomalous-dispersion regime ( s = 1 ), the
value of = 1 2CZ changes from positive to negative for C > 0 as the propagation distance
increases while the value of is always positive for C < 0 . The opposite occurs in the
normal-dispersion regime ( s = 1 , = 1 + 2CZ ). As a result, chirped Airy pulses can switch
direction between acceleration and deceleration for sC > 0 , while it always dispersed for
sC < 0 . Therefore, the linear propagation dynamics of chirped Airy pulse depends strongly
on the signs of dispersion and chirp. It can switch between acceleration and deceleration for
sC > 0 , while it always dispersed for sC < 0 .

Fig. 2. Airy pulse shapes with positive (left column) and negative (right column) chirps for six
different propagation distances in the anomalous-dispersion regime.

3. Numerical results
To confirm the propagation dynamics of chirped Airy pulse obtained from the analytical
analysis, we model Airy pulse propagation with the Eq. (1) in an optical fiber by using the
well-known split-step Fourier method [42]. To get a better visualization of the low-intensity
parts of the Airy pulse propagation, hereafter, we plot the evolution of temporal and spectral
absolute amplitude U ( Z , T ) and U ( Z , ) and instead of the corresponding intensity
2
U and U as a function of propagation distance. For N = 0 , Eq. (1) is only capable of
2

modeling the effects of GVD on optical pulses propagating in a linear dispersive medium.
Figure 3(a) demonstrates the linear propagation of an initially unchirped Airy pulse, showing
that it maintains its all remarkable properties of the ideal self-decelerating Airy pulses over an

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2571
finite propagation distance. The shape of self-decelerating unchirped Airy pulse remains
quasi-invariant over several dispersion lengths while again the intensity features tend to
freely decelerate until dispersion dominates. It is consistent with results in Ref [2].
However, this behaviors change drastically if the Airy pulse has an initial frequency chirp.
Figures 3(b) and 3(c) display the dynamics of the Airy pulse with positive and negative
frequency chirp in the case of anomalous dispersion regime respectively. Clearly, the linear
dynamics of chirped Airy pulses are quite different when compared to that associated with
unchirped Airy pulse, as depicted in Fig. 3(a). It is evident that the sign of chirp parameter C
plays a critical role.

Fig. 3. Evolution of (left column) pulse U ( Z , T ) and (right column) spectra U ( Z , )


shapes over a distance of 10 dispersion lengths for an initially unchirped and chirped Airy
pulses in the anomalous dispersion regime.

A comparison of Figs. 3(b) and 3(c) shows that, the effect of frequency chirp on the
propagation of Airy pulse depends on whether C is positive or negative. When the incident
Airy pulse was positively chirped ( C > 0 ), the propagation of Airy pulse shown in Fig. 3(b)
can be separated to three regimes of interest: it undergoes compression in the first 2.2
propagation distance; then the pulse pattern breaks up with further increasing of the
propagation distance; Finally, a new Airy pattern with the rotational symmetry distribution
with respect to the input pulse was regenerated. The amplitude of new Airy pulse is smaller
than that of original Airy pulse. However when the incident Airy pulse was imposed a
negative chirp ( C < 0 ), the Airy pulse was always dispersed during the propagation. Their
lobes experience an apparent acceleration, exactly opposite to its original deceleration. The
numerical simulations confirm the novel propagation dynamics of chirped Airy pulses
predicted by analytical analysis. In addition, Fig. 4 displays a comparison of pulse shapes
between analytical and numerical results at some representative propagation distances. All
solid redlines overlap with red dash lines. Once again, we get very good qualitative agreement
between analytical results and numerical simulation of the NLSE. Although, this process is

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2572
similar to that reported in [21], where the reason is the dominant contribution of third-order
dispersion parameter, the physical reason behind such propagation dynamic is different.
To have a better understanding of the propagation dynamics of chirped Airy pulse, let us
look into the corresponding spectra. The spectral evolution of chirped Airy pulses have
dramatic difference compared to that of unchirped Airy pulse as shown in the right column of
Fig. 3. The spectrum of unchirped Airy pulse shown in Fig. 3(d) always keeps the Gaussian
distribution without any change along the propagation direction. While the spectral evolution
of chirped Airy pulse has oscillatory tail behind or in front of its main peak, depending on
whether chirp ( C ) is negative or positive. This agrees well with analytical results (Fig. 1).
Furthermore, its shape is also not changed during propagation. We also performed the linear
propagation of a chirped Airy pulse in a fiber with normal dispersion (data not shown),
indicating that the opposite occurs. The features of chirped Airy pulse propagation seen in
Fig. 3 can be understood qualitatively as follows. The dispersion-induced chirp is positive or
negative, depending on whether dispersion is normal or anomalous. When 2 C > 0 , the
dispersion-induced chirp adds to initial input chirp because the two contributions have the
same sign. In the case of 2 C < 0 , the situation changes dramatically, the contribution of the
dispersion-induced chirp is of a kind opposite to that of the initial input chirp. Naturally, the
same propagation dynamic occurs, if the incident chirped Airy pulses are launched with a
reversed acceleration.

Fig. 4. Comparison of pulse shapes obtained from numerical (black solid lines) and analytical
(red dash lines) methods at Z = 1 and Z = 7 for positively and negatively chirped Airy
pulse.

Figure 5 shows the temporal evolution of chirped Airy pulse as a function of propagation
distance for different values of C ranging from 0.1 to 2.0. It illustrates how much an initial
chirp can modify the propagation behavior of Airy pulse. For C = 0.1 , the chirped Airy pulse
just experiences an initial compression, and then breaks up in the 20 propagation distances.
There is no new Airy pattern formation. With the increasing the chirp, new Airy pattern with
reversal acceleration can be clearly seen. The propagation distance required for new Airy
formation decreases with increasing chirp. It can be clearly seen from Fig. 6(a) that the size of
breakup area ( LB , measured from one amplitude maximum to another and marked in Fig.

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2573
4(a1)) becomes smaller and smaller with an increasing C . For a small chirp ( C 0.1 ), LB
becomes very large such that the size diverges. In addition, the new formed Airy pulse carries
more energy for larger chirp. But the Airy pulse is also distorted in shape.

Fig. 5. Temporal evolutions of a chirped Airy pulse with truncation coefficient a = 0.1 and
different initial chirp as a function of propagation distance.

Fig. 6. (a) the size of breakup area ( LB ) depends on the chirp parameter for three different
truncation factor; (b) the LB as a function of the truncation coefficients for three different
chirps.

When the self-accelerating chirped Airy pulses have a large value of truncation
coefficient, their unique features disappear rapidly. Their unique features therefore strongly
depend on the value of truncation coefficient. Figure 7 shows temporal evolutions of chirped
Airy pulse with different values of truncation coefficient over 2 dispersion lengths. The
process of compression, breakup and regeneration are almost unchanged for different
truncation coefficients. However, the compression factor, defined as the ratio of maximum
peak intensity to input peak intensity, decreases with increasing truncation coefficients. The
size of breakup area slightly increases as the truncation coefficient is increased. The rate of
the size growth depends on the initial chirp [Fig. 6(b)]. The Airy pulses with larger truncation
coefficient will loss more energy through the breakup area.

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2574
Fig. 7. Temporal evolutions of an initially chirped Airy pulse with different truncation
coefficient as a function of propagation distance.

Fig. 8. Temporal evolutions of an initially chirped Airy pulse with a = 0.1 as a function of
propagation distance in the anomalous dispersion regime under the action of different Kerr
nonlinearity ( N ).

Under the action of Kerr nonlinearity, the Airy pulse was distorted in the form of soliton
shedding and dispersion background during propagation [24]. Figures 8(b)-8(f) show the
temporal evolution of Airy pulse for different Kerr nonlinearity. Compared the case of linear
propagation shown in Fig. 8(a), the compression ratio is reduced owing to the presence of
nonlinear self-focusing as nonlinear parameter N is increased. For N 1 , the nonlinear self-
focusing effects play an increasingly more important with increasing N , then become

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2575
comparable to the dispersion effects. The combined effects of GVD and Kerr nonlinearity are
to reduce the compression ratio; the size of breakup area widens but, at the same time, the
resulting energy loss increased. The situation changes dramatically for N > 1 . In this case,
the nonlinear self-focusing effect dominates the dispersion effect, which completely disrupts
the phase evolution of chirped Airy pulse. Under the action of an even stronger nonlinearity,
as shown in Figs. 8(g)-8(i), the high energy Airy pulse enters a soliton shedding regime,
similar to that reported in [24]. The trajectories of the shedding solitons are affected by
successive collisions with the side lobes and dispersion background. The number of shedding
soliton increases with increasing N . As a consequence, the rotational symmetry of
propagation process has been changed dramatically.
4. Conclusion
In summary, we have investigated the propagation dynamics of truncated Airy pulse with an
initial frequency chirp imposed in optical fibers by means of direct simulation and theoretical
analysis. The analytical expression for chirped Airy pulse propagation is obtained. We find
that, when the second-order dispersion parameter 2 and chirp C have the opposite signs,
the chirped Airy pulse reaches the breakup area after experiencing an initial compression,
then undergoes a lossy inversion, and finally continues to travel with an opposite acceleration.
The size of breakup area, measured from one amplitude maximum to another, decreases with
increasing C or decreasing truncation coefficients. While the chirped Airy pulses are always
dispersive in the case of 2 C > 0 . The analytical results are very well agreement with the
numerical simulations. In the weakly nonlinear propagation regime, the compression ratio is
reduced that accompanies severe energy loss as the nonlinear parameter N is increased
ranging from 0 to 1. With further increasing Kerr nonlinearity ( N > 1 ), such stronger Kerr
nonlinearity brings the chirped Airy pulse propagation into the regime of multiple soliton
generation and successive collisions.
Acknowledgments
The authors thank the referees for enlightening comments on the analytical results of chirped
Airy pulse propagation. This work was supported by the Program of Fundamental Research
of Shenzhen Science and Technology Plan (Grant Nos. JCYJ20140828163634005,
JCYJ20140418095735599), the Natural Science Foundation of SZU (Grant Nos. 201449,
201450) and the Hunan Provincial Natural Science Foundation of China (Grant Nos.
13JJ4108, 15JJ2036).

#226391 - $15.00 USD Received 6 Nov 2014; revised 19 Jan 2015; accepted 19 Jan 2015; published 29 Jan 2015
(C) 2015 OSA 9 Feb 2015 | Vol. 23, No. 3 | DOI:10.1364/OE.23.002566 | OPTICS EXPRESS 2576

You might also like