You are on page 1of 8

Research Article Vol. 35, No.

2 / February 2018 / Journal of the Optical Society of America B 323

Supercontinuum generation in organic


liquid-liquid core-cladding photonic crystal
fiber in visible and near-infrared regions
RASOUL RAEI,1 MAJID EBNALI-HEIDARI,1,* AND HAMED SAGHAEI2
1
Faculty of Engineering, Shahrekord University, Shahrekord, Iran
2
Department of Electrical Engineering, Shahrekord Branch, Islamic Azad University, Shahrekord, Iran
*Corresponding author: ebnali-m@eng.sku.ac.ir

Received 31 August 2017; revised 26 November 2017; accepted 15 December 2017; posted 15 December 2017 (Doc. ID 305293);
published 17 January 2018

In this paper, we propose a liquid core-cladding photonic crystal fiber (PCF), which is engineered with different
available organic optofluidics, to generate supercontinuum in the visible and near-infrared regimes by using the
symmetrized split-step Fourier method. Simulations reveal that in response to launching 50 fs input pulses of 10 kW
peak power, centered about λ0  1032 and 1560 nm, into a 10 mm long liquid core-cladding PCF, maximum
2 μm supercontinua from 500 to 2500 nm can be achieved. Our numerical study is important for the new field
of visible and near-IR supercontinuum generation in liquid-core optical fibers. © 2018 Optical Society of America
OCIS codes: (320.6629) Supercontinuum generation; (060.5295) Photonic crystal fibers; (260.2030) Dispersion; (160.4890) Organic
materials; (190.4710) Optical nonlinearities in organic materials; (300.6550) Spectroscopy, visible.

https://doi.org/10.1364/JOSAB.35.000323

1. INTRODUCTION relatively high nonlinear refractive indices when compared to


Supercontinuum generation (SG) in a photonic crystal fiber solids (up to 100 times larger compared to fused silica [7]).
(PCF) has been the topic of wide studies over the past decade. Broad mid-infrared supercontinuum sources are fascinating
For the input pulses in the anomalous dispersion regime of a and sought after in research because they can excite vibrational
PCF, the main dynamics of SG include an initial fission of the absorption bands in the molecular fingerprint region (2.5–
N th-order soliton, self-steepening effect, self-phase modula- 25 μm), which makes it possible to identify molecules and
functional groups [8] correctly. In addition, for the generation
tion, intrapulse Raman scattering, loss and higher-order
of nonlinear optical responses, very high intensities are needed.
dispersion (HOD) [1,2]. The only method to enhance the non-
To combine the benefits of high nonlinearities of optofluidics
linear coefficient of a PCF with silica core is to reduce the ef-
fective core area. Optical fibers, with their small cross-section and high peak powers in fibers, selectively fluidic-filled PCFs
area, cannot deliver the necessary intensities over a long inter- [9–11] and integrated fluidic-core optical fibers [12–14] have
action length. Also, it will lead to coupling difficulty. However, come into attention. Also, in [15,16] supercontinuum with
because of the small nonlinear refractive index of fused silica, 320 nm and 330 nm broadening using toluene and CS2 , respec-
special nonsilica fibers, such as ZBLAN [3], SF6 [4], TF10 [5], tively, was generated. The benefit of this combination is that the
and chalcogenide-based fibers [6], have been developed exper- dispersion properties can be tailored accordingly, e.g., by chang-
imentally with a larger transmission window and enhanced ing the core diameter, using combinations of different fluids, or
nonlinear optical properties, but they are limited in function- changing their temperature and hence nT  [17]. An additional
ality and flexibility. This means that due to the highly nonlinear advantage is that some highly nonlinear liquids have a wide
refractive indices of these materials, nearly an order of magni- transparent range in the mid-infrared region [18], which is es-
tude higher than that of silica, supercontinuum (SC) is ex- sential for spanning supercontinua far into this spectral region.
pected to generate on significantly shorter propagation The highly nonlinear properties of the fluids allow for a decrease
scales. Therefore, the higher the nonlinearity of the PCF, in the required powers significantly, which makes liquid-filled
the broader and smoother supercontinuum will be generated fibers also available for cheap, compact pump lasers.
with the same laser parameters. Highly nonlinear optofluidics Also, the characteristics of self-phase modulation can be
offer striking possibilities in the field of optics because they studied in such liquid-filled capillary devices [19], which makes
are especially suited for making supercontinuum due to their it possible to measure nonlinear refractive index data of the core

0740-3224/18/020323-08 Journal © 2018 Optical Society of America


Corrected 31 May 2018
324 Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B Research Article

liquids precisely. In the previous works, only broadly scattered B. Linear Parameters
values are reported, which are measured with sub-nanosecond The refractive indices of used materials as a function of wave-
pulses in the near-infrared spectral region [20,21]. Today, com- length are given by the Sellmeier equation as
pact sub-femtosecond lasers have become more abundant [22].
This paper is focused on the SC generation through a short nλ  f1  B i λ2 λ2 − C i −1 g0.5 ; (1)
piece of PCF with an optofluidic approach in the visible and where constants B i and C i μm  (i  1; 2; 3) in the Sellmeier
2

near-infrared region (NIR). Infiltration of the PCFs’ core- equation for these materials are shown in Table 1.
cladding by a combination of high nonlinear optical fluids is The real and imaginary parts of refractive indices of the men-
used to control and tailor the PCFs’ dispersion profiles, as de- tioned materials used in this paper versus wavelength are shown
sired. Other considerations, such as loss, dispersion, effective in Figs. 2(a) and 2(b), respectively. As can be observed in
area, and nonlinear coefficient, are taken into account. Fig. 2(a), the refractive index of C7 H8 is more than the other
The rest of this paper is organized as follows. In Section 2, mentioned liquids. Therefore, the guiding mechanism is pro-
profiles for the wavelength dependence of the linear, and exper- vided by the total internal reflection along the liquid core PCF.
imental data for nonlinear parameters for infiltrated liquids The dispersion of the PCF known as the group velocity
used in the proposed PCF are plotted and compared, and dispersion, having both the waveguide and the material disper-
the generalized nonlinear Schrödinger equation (GNLSE) gov- sions, seen by the FM of wavelength λ, is determined by
erning the optical pulse propagation along the optical fiber is λ d2 2πc
presented. Using numerical values for the linear and nonlinear Dλ  − 2 Reneff λ  − 2 β 2 ; (2)
c dλ λ
parameters obtained in Section 2, the GNLSE is rewritten and
numerically solved employing the symmetrized split-step where c, β2 , and neff λ are the light velocity in free space, the
Fourier method (S-SSFM) across the PCF. In Section 3, the second-order dispersion, and the fiber’s effective index, respec-
spectral distributions of the PCF outputs versus the center tively. The higher-order dispersions become important when
wavelengths of the input optical pulses in the visible and the input pulse center wavelength shifts toward the zero
near-infrared regimes are illustrated and compared. Finally, dispersion wavelength (ZDW) that is given by

the paper is closed by the conclusion in Section 4. d m β 
βm  : (3)
d ωm ωω0

2. PROPOSED PCF SPECIFICATIONS In which, ω0 is the light center frequency. To calculate neff λ
for the PCF, we employ a full vector modal solver based on
A. Structure
the finite-difference time-domain numerical method. The
In this section, we consider a conventional PCF whose back- dispersion (D) of the FM is calculated as a function of the wave-
ground materials are assumed to be made of silica (FK51A), the length in the range of 0.5–2 μm in Fig. 2(c), for the simulated
core material is made of C7 H8 , and the cladding material con- PCF, infiltrated with the mentioned organic optofluidics hav-
sists of CCl4 , CHCl3 , C2 H5 OH, and H2 O. The perspective ing well-known losses. Figure 2(c) also shows when optofluidics
and cross-sectional views of the PCF are depicted in Fig. 1. infiltrate the air holes, ZDW is shifted toward longer wave-
The cross-sectional view of the PCF, depicted in Fig. 1(a), lengths, and the dispersion slope is decreased. The ZDWs
is assumed to have five rings of air holes of the same diameters of the PCF FM are 0.9, 1.2, 1.3, 1.52, and 2 μm, for unin-
(d  2 μm) arranged in a triangular lattice of constant filtrated and infiltrated with CCl4 , CHCl3 , H2 O, and
Λ  2.2 μm, encompassing the PCF fluidic core of d c  C2 H5 OH, respectively. Low anomalous dispersion near the
2 μm with 10 mm long. The number of rings of air holes ZDW provides the generation of a broad supercontinuum with
in this PCF is large enough so that the change in the PCF’s strong confinement to the core. By shifting the ZDW using the
dispersion would be negligible if another ring is added on optofluidic approach as desired, we can use available lasers with
the outer side of the cladding. femtosecond optical pulses as input sources. Table 2 compares

Fig. 1. (a) Cross-sectional and (b) perspective view of liquid-liquid core-cladding PCF infiltrated with organic optofluidics.
Research Article Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B 325

Table 1. Sellmeier Coefficients of Liquids Used in This simulation window and to evaluate the confinement loss of
Paper the presented PCF, we assume anisotropic perfectly matched
Sellmeier Coefficients
layers (absorbing boundaries) to be placed outside the outer-
most ring of the air holes.
Materials B1 C1 B2 C2 B3 C3 Note that material loss is neglected due to the weak absorp-
C7 H8 1.1747 0.0182 — — — — tion coefficients of the liquids and the small fiber lengths used
Silica 0.9712 0.0047 0.2169 0.0153 0.9046 168.6813 in our numerical study [10,23].
(FK51A) Also, the scattering loss in our proposed PCF in the wave-
CCl4 1.0921 0.0118 — — — — length range of interest (i.e., the NIR region) is negligible.
CHCl3 1.0464 0.0104 0.0034 0.152 — —
C2 H5 OH 0.8318 0.0093 −0.1558 −49.452 — — C. Nonlinear Parameter
H2 O 0.7583 0.01 0.0849 8.9137 — —
Now, consider the fiber nonlinear parameter defined as [24]

γW · m−1  2πn2I ∕λ0 Aeff λ; (4)


the values of higher-order dispersions from β2 to β8 calculated
where λ0 is the center wavelength and n2I is the nonlinear Kerr
for infiltrated PCF with the mentioned fluids for an input op-
index of the fiber’s background material, which can be
tical pulse with a center wavelength of 1032 and 1560 nm.
expressed as follows:
Another linear parameter that needs to be examined is the
fiber’s total loss, L  Lm  Lc with Lm and LC λ  n2I  ΦNL;max λAeff T FWHM f rep ∕1.76πP avg L: (5)
8.686k0 Im neff λ being the material loss of the mentioned
chemical compositions and the confinement (leakage) loss that The attained results from the measurements at both pump laser
depends on the fiber’s structural parameters, with k0  2π∕λ sources at a wavelength of 1032 nm and 1560 nm according to
being the free-space wavenumber. Moreover, to reduce the [25] are summarized in Table 3.

Fig. 2. Comparison of the (a) real part and (b) imaginary part of the used material refractive indices, (c) dispersion profiles, and (d) confinement
loss of the uninfiltrated clad with those infiltrated with optical fluids of various indices.
326 Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B Research Article

Table 2. Values of the Dispersion Orders of Proposed PCF Infiltrated by Various Organic Optofluidics at 1032 and 1560
nm Pump Wavelength
Infiltrated Liquids λ (nm) β2 (ps2 ∕m) β3 (ps3 ∕m) β4 (ps4 ∕m) β5 (ps5 ∕m) β6 (ps6 ∕m) β7 (ps7 ∕m) β8 (ps8 ∕m)
CCl4 0.027 9.51e − 5 −8.96e − 8 2.63e − 10 −6.7e − 13 1.94e − 15 −3.94e − 18
CHCl3 0.041 9.04e − 5 −1.05e − 7 2.74e − 10 −5.96e − 13 1.65e − 15 −3.39e − 18
C2 H5 OH 1032 0.061 3.66e − 6 5.81e − 8 3e − 10 −2.01e − 12 4.69e − 15 7.58e − 18
H2 O 0.045 5.29e − 5 −2.61e − 8 −1.08e − 9 4.94e − 12 4.32e − 14 −4.89e − 16
Uninfiltrated clad −0.036 0.0001 −2.43e − 8 3.08e − 10 1e − 12 −4.45e − 14 3.35e − 16
CCl4 −0.056 0.0002 −4.43e − 7 1.13e − 9 −2.59e − 12 4.83e − 15 −5.94e − 18
CHCl3 −0.042 0.0002 −4.4e − 7 1.03e − 9 −2.22e − 12 4.07e − 15 −4.81e − 18
C2 H5 OH 1560 0.047 0.0001 −5.35e − 7 1.62e − 9 −6.02e − 13 −1.54e − 14 6.79e − 17
H2 O 0.001 0.0003 −3.59e − 6 3.03e − 8 −1.79e − 10 7.4e − 13 −2.05e − 15
Uninfiltrated clad −0.063 −0.0003 3.81e − 6 −2.54e − 8 1.29e − 10 −4.87e − 13 1.26e − 15

observed from this table, the nonlinear parameter for the fiber
Table 3. Nonlinear Refractive Indices at the Two Pump
decreases as the wavelength increases.
Wavelengths of 1032 nm and 1560 nm
Chemical Formula Wavelength (nm) n2I m2 ∕W D. Generalized Nonlinear Schrödinger Equation

C7 H8 1032 0.62e − 18 The evolution of an optical pulse, having a slowly varying elec-
1560 0.59e − 18 tric field amplitude represented by envelope function Az; t,
propagating along the fiber shown in Fig. 1 can be described by
the solution to the GNLSE [24]:
Moreover, Aeff λ shows the wavelength dependence of the X8
∂A α β ∂n A
fiber effective cross-sectional area defined as  A i n−1 n n
∂z 2 n1
n! ∂t
 2  
RR ∞ ∂
−∞ jF x; yj dxdy
2
 i γω0   iγ 1
∂t
Aeff  R ∞ ; (6) Z t
−∞ jF x; yj4 dxdy
× Az; t Rt 0 jAz; t − t 0 j2 dt 0 ; (7)
−∞
where F x; y is the optical field distribution across the fiber
where z and t are the spatial coordinates along the fiber and the
cross section. The numerical values of the effective cross section
time variable, α is the total fiber loss, and βq represent the qth
calculated for the PCF are depicted in Fig. 3. A comparison
order dispersion parameter. Rt in the integrand on the right-
shows that the PCF experiences larger effective cross-sectional
hand side of the equation is the response function including the
area when infiltrated by chemical composition with higher
Raman and Kerr nonlinearities. Assuming that the electronic
refractive index.
The nonlinear parameters related to the mentioned PCF cal- contribution is nearly instantaneous, the functional form of
culated by Eq. (4) are also illustrated in Table 4. As can be Rt can be written as
Rt  1 − f R δt  f R hR t; (8)
where δt is the Dirac delta function, and f R represents the
fractional contribution of the delayed reorientational response
function hR t, which is measured in 1032 and 1560 nm for
C7 H8 and depicted in Table 5 [25].

Table 4. γ and Aeff in 1032 and 1560 nm for Uninfiltrated


and Infiltrated Clad by Various Optofluidics
Infiltrated Liquids Wavelength (nm) Aeff μm2  γ (1/W/m)
CCl4 4.721 0.79957
CHCl3 4.08 0.92519
C2 H5 OH 1032 2.524 1.49556
H2 O 2.296 1.64407
Uninfiltrated clad 1.622 2.32724
CCl4 6.25 0.38021
CHCl3 5.75 0.41328
C2 H5 OH 1560 4 0.59408
H2 O 3.5 0.67895
Fig. 3. Effective area versus wavelength for uninfiltrated and infil-
Uninfiltrated clad 2 1.18817
trated clad by various optofluidics.
Research Article Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B 327

Table 5. Fractional Contributions at the Two Pump use practically the minimum length of the PCF. Figure 4 shows
Wavelengths of 1032 and 1560 nm for C7 H8 the transmittance curves as a function of wavelength of 10 mm
Infiltrated Liquid f R at 1032 nm f R at 1560 nm
thickness in 20°C, for the used fluidics in this paper.
C7 H8 0.88 0.70

Table 6. Coefficients of the Reorientational Response


Function of C7 H8
A1 A2 A3 t diff t rise;1 t int t fast t rise;2
(1/s) (1/s) (1/s) (ps) (ps) (ps) (ps) (ps)
0.1265 0.3106 2.646 5.9 0.15 0.50 0.10 0.15

hR t takes an approximate analytic form of


hR t  A1 e −t∕t diff  1 − e −t∕t rise;1  
 A2 e −t∕t int  1 − e −t∕t rise;1  
 A3 e −t
2 ∕2t 2 
fast sint∕t rise;2 ; (9)
with a diffusive (t diff ), an intermediate (t int ), and a librational
(t fast ) relaxation process. To account for the inertial effects, a
rise time (t rise ) is included, which should be similar for all
of the noninstantaneous nuclear responses [26]. Coefficients
of Eq. (9) are listed in Table 6.
A short temporal optical soliton of width T 0 , peak power
P 0 , and order N is used to have the advantage of a lower
average power and broader bandwidth:
 
pffiffiffiffiffi t
Az  0; t  P 0 sech
T0
  0.5  
N jβ2 j t
 sech : (10)
T0 γ T0
Using the S-SSFM, one can obtain the numerical solution of
Eq. (7) [24].
Increasing the PCF length causes transmittance decreases
and total loss increases, respectively. Therefore, we should

Fig. 5. Spectral distribution along the liquid-liquid core-clad PCF


for uninfiltrated clad (row 1) and infiltrated by C2 H5 OH (row 2),
Fig. 4. Transmittance versus wavelength for 10 mm thickness in CCl4 (row 3), CHCl3 (row 4), and H2 O (row 5) pumped by
20°C. 50 fs laser at 1032 nm (column 1) and 1560 nm (column 2).
328 Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B Research Article

It demonstrates that CCl4 has the maximum and flattest trans- the simulation result in these figures demonstrates about
mission window. However, H2 O has the lowest one. Spectral 800 nm bandwidth of supercontinuum in the case of
broadening in the PCF depends on the transmittance window C2 H5 OH at 1560 nm.
of the used material. For Fig. 5 (11, 12, 32, 42, 52), the wavelengths of the input
pulses have been considered in the anomalous dispersion re-
gime where the effect of SPM and SSFS due to the stimulated
3. SUPERCONTINUUM GENERATION Raman scattering (SRS) are predominant and responsible for
In this section, by extracting both linear parameters containing broadband generated spectra. When pumping in the anoma-
loss and dispersion from the second to eighth orders and non- lous dispersion regime, the input pulse transforms into
linear parameters in our desired wavelength, the GNLSE is re- higher-order solitons that a small perturbation, such as
written for the proposed PCF depicted in Fig. 1. Then via the HOD, SS, and SRS, can break up a given soliton of order
S-SSFM, the GNLSE will be solved in both the time and N into N fundamental solitons [24]. As a result, SSFS shifts
frequency domains for femtosecond input pulses. We assume the center wavelengths of the resulting fundamental solitons
that input pulses with center wavelengths of 1032 and continuously toward the longer wavelengths of the spectrum,
1560 nm, having the same width of T 0  50 fs for peak causing a considerable spectral expansion on the red side. On
powers of P 0  10 kW, are launched into 10 mm length of the other hand, nonsolitonic radiations result in broadening
the PCF. on the blue side of the spectrum. The simulation result in
Figure 5 compares the spectral distributions of the optical these figures demonstrates about the 2 μm bandwidth of
pulses evolving along the infiltrated PCF with various optoflui- supercontinuum in uninfiltrated and CCl4 clad cases at
dics. This figure also reveals the potential of the optofluidic 1560 nm.
technique for SC generation with wide bandwidth, simply As can be seen in Fig. 5 (21–52), depending on the refrac-
by choosing an appropriate fluid and the input signal wave- tive index of the infiltrated fluid, SC with desired intensity and
length. As can be observed in Fig. 5, the maximum spectral bandwidth is achieved. With this technique, we do not need to
broadening occurs for pumping in the anomalous dispersion change the physical parameters of the PCF such as air hole
regime. diameter, core diameter, and lattice constant by too complex
It is clear that by increasing the optical pulse peak power, fabrication processes. The optical fluids used for this purpose
broader SC is achieved. Various linear and nonlinear phenom- must be nontoxic, highly nonlinear, and transparent to the
ena are accountable for the generation of the supercontinua wavelengths of interest. Therefore, available fluids help us gen-
observed in Fig. 5. erate supercontinuum more easily, faster, and more cheaply and
As can be seen from Fig. 5 (21, 22, 31, 41, 51), by pumping changing the infiltrated fluids leads to adjusting the ZDW in
optical pulses in the normal dispersion regime, the pulses the range of available laser sources.
initially undergo strong self-phase modulation (SPM), leading Consequently, by the appropriate choice of the infiltrated
to wave breaking due to self-steepening (SS) and HOD that fluid, input optical pulse, and PCF properties, a broad and
results in a part of the light to be blueshifted [27,28]. coherent spectrum at the fiber output is achieved. Figure 5 also
However, the minimum wavelength of the blueshift is limited demonstrates that a coherence SC as wide as 2 μm is generated
by the total loss. The redshifting part will eventually cross at 10 mm length of the PCF pumped in the anomalous
the ZDW, at which point soliton dynamics, in particular, dispersion regime.
Raman-induced soliton self-frequency shifting (SSFS), domi- Aiming to study the coherence degradation, the analysis of
nates further spectral broadening [28,29]. As expected, the coherence properties of the generated supercontinua from

Fig. 6. First-order degree of coherence for (a) 1032 nm and (b) 1560 nm for uninfiltrated and infiltrated with mentioned optofluidics.
Research Article Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B 329

the proposed PCF is performed. It considers the addition of one 6. Y. Yu, X. Gai, T. Wang, P. Ma, R. Wang, Z. Yang, D.-Y. Choi, S.
Madden, and B. Luther-Davies, “Mid-infrared supercontinuum
photon per mode noise with a random phase, and it is evaluated
generation in chalcogenides,” Opt. Mater. Express 3, 1075–1086
within the first-order degree of coherence given by (2013).
7. P. Ho and R. Alfano, “Optical Kerr effect in liquids,” Phys. Rev. A 20,
jhE 1 λ; t 1 E 2 λ; t 2 ij
jg 1
2170–2187 (1979).
12 λ; t 1 − t 2  0j  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi; (11) 8. D. Lin-Vien, N. B. Colthup, W. G. Fateley, and J. G. Grasselli, The
hjEλ; t 1 j2 ihjEλ; t 1 j2 i
Handbook of Infrared and Raman Characteristic Frequencies of
Organic Molecules (Elsevier, 1991).
where E 1 λ and E 2 λ are two electric fields. The angular 9. M. Vieweg, T. Gissibl, S. Pricking, B. Kuhlmey, D. Wu, B. Eggleton,
brackets indicate an ensemble average over independently gen- and H. Giessen, “Ultrafast nonlinear optofluidics in selectively
erated pairs of SC spectra, and t is the time measured at the liquid-filled photonic crystal fibers,” Opt. Express 18, 25232–25240
scale of the temporal resolution of the spectrometer used to (2010).
10. R. Zhang and H. Giessen, “Theoretical design of a liquid-core pho-
resolve these spectra. g 12 is considered at t 1 − t 2  0 to focus tonic crystal fiber for supercontinuum generation,” Opt. Express 14,
on the wavelength dependence of the coherence. It is known 6800–6812 (2006).
that the coherence property of SC light is significantly 11. J. Bethge, A. Husakou, F. Mitschke, F. Noack, U. Griebner, G.
affected by the quantum noise of the pump pulse. From an Steinmeyer, and J. Herrmann, “Two-octave supercontinuum
ensemble average of 50 independent simulations, we find that generation in a water-filled photonic crystal fiber,” Opt. Express 18,
6230–6240 (2010).
the generated supercontinuum in the 10 mm long PCF infil- 12. S. Kedenburg, T. Gissibl, T. Steinle, A. Steinmann, and H. Giessen,
trated by C2 H5 OH with an input peak power of 10 kW is “Towards integration of a liquid-filled fiber capillary for supercontin-
coherent over the entire generated bandwidth that is shown uum generation in the 1.2–2.4 μm range,” Opt. Express 23,
in Fig. 6, in which C2 H5 OH has the flattest coherence among 8281–8289 (2015).
them. 13. K. Kieu, L. Schneebeli, R. A. Norwood, and N. Peyghambarian,
“Integrated liquid-core optical fibers for ultra-efficient nonlinear liquid
We have considered the practical terms in our model. Thus, photonics,” Opt. Express 20, 8148–8154 (2012).
we recommend that readers implement the proposed structure 14. L. Xiao, N. V. Wheeler, N. Healy, and A. C. Peacock, “Integrated hol-
experimentally to achieve the same results. The fiber-based low-core fibers for nonlinear optofluidic applications,” Opt. Express 21,
supercontinuum sources proposed in this paper represent a 28751–28757 (2013).
revolution in broadband mid-infrared region (MIR) light 15. G. Fanjoux, S. Margueron, J.-C. Beugnot, and T. Sylvestre,
“Supercontinuum generation by stimulated Raman-Kerr scattering
sources for different MIR applications.
in a liquid-core optical fiber,” J. Opt. Soc. Am. B 34, 1677–1683
(2017).
16. G. J. Raj, R. V. J. Raja, N. Nagarajan, and G. Ramanathan, “Tunable
4. CONCLUSION broadband spectrum under the influence of temperature in IR region
We have proposed the generation of supercontinuum in the using CS2 core photonic crystal fiber,” J. Lightwave Technol. 34,
3503–3509 (2016).
visible and near-infrared frequency ranges, utilizing a silica-
17. M. Vieweg, S. Pricking, T. Gissibl, Y. V. Kartashov, L. Torner, and H.
based liquid core-cladding PCF approach. Simulations have Giessen, “Tunable ultrafast nonlinear optofluidic coupler,” Opt. Lett.
shown that when a 50 fs input optical pulse of peak power 37, 1058–1060 (2012).
10 kW and center wavelength 1560 nm is launched into a 18. C. J. Pouchert, The Aldrich Library of Infrared Spectra, 3rd ed. (Aldrich
10 mm long PCF infiltrated by appropriate fluidics, spectral Chemical Co., 1981).
19. M. C. P. Huy, A. Baron, S. Lebrun, R. Frey, and P. Delaye,
broadening as wide as 2 μm in both the visible and near-
“Characterization of self-phase modulation in liquid filled hollow core
infrared regions can be obtained. photonic bandgap fibers,” J. Opt. Soc. Am. B 27, 1886–1893
(2010).
20. S. Couris, M. Renard, O. Faucher, B. Lavorel, R. Chaux, E.
REFERENCES Koudoumas, and X. Michaut, “An experimental investigation of the
1. K. M. Hilligsøe, H. N. Paulsen, J. Thøgersen, S. R. Keiding, and J. J. nonlinear refractive index (n2) of carbon disulfide and toluene by spec-
Larsen, “Initial steps of supercontinuum generation in photonic crystal tral shearing interferometry and z-scan techniques,” Chem. Phys.
fibers,” J. Opt. Soc. Am. B 20, 1887–1893 (2003). Lett. 369, 318–324 (2003).
2. J. M. Dudley, L. Provino, N. Grossard, H. Maillotte, R. S. Windeler, B. 21. M. J. Weber, Handbook of Optical Materials (CRC Press,
J. Eggleton, and S. Coen, “Supercontinuum generation in air-silica 2002).
microstructured fibers with nanosecond and femtosecond pulse 22. S. Pekarek, A. Klenner, T. Südmeyer, C. Fiebig, K. Paschke, G.
pumping,” J. Opt. Soc. Am. B 19, 765–771 (2002). Erbert, and U. Keller, “Femtosecond diode-pumped solid-state
3. O. P. Kulkarni, V. V. Alexander, M. Kumar, M. J. Freeman, M. N. Islam, laser with a repetition rate of 4.8 GHz,” Opt. Express 20,
F. L. Terry, Jr., M. Neelakandan, and A. Chan, “Supercontinuum 4248–4253 (2012).
generation from ∼1.9 to 4.5 μmin ZBLAN fiber with high average 23. S. Kedenburg, M. Vieweg, T. Gissibl, and H. Giessen, “Linear refrac-
power generation beyond 3.8 μm using a thulium-doped fiber tive index and absorption measurements of nonlinear optical liquids in
amplifier,” J. Opt. Soc. Am. B 28, 2486–2498 (2011). the visible and near-infrared spectral region,” Opt. Mater. Express 2,
4. F. Omenetto, N. Wolchover, M. Wehner, M. Ross, A. Efimov, A. 1588–1611 (2012).
Taylor, V. R. K. Kumar, A. George, J. Knight, and N. Joly, 24. G. P. Agrawal, “Nonlinear fiber optics,” in Nonlinear Science at the
“Spectrally smooth supercontinuum from 350 nm to 3 μm in sub- Dawn of the 21st Century (Springer, 2000), pp. 195–211.
centimeter lengths of soft-glass photonic crystal fibers,” Opt. 25. S. Kedenburg, A. Steinmann, R. Hegenbarth, T. Steinle, and H.
Express 14, 4928–4934 (2006). Giessen, “Nonlinear refractive indices of nonlinear liquids: wavelength
5. A. Fedotov, E. Serebryannikov, A. Ivanov, and A. Zheltikov, dependence and influence of retarded response,” Appl. Phys. B 117,
“Spectral transformation of femtosecond Cr:forsterite laser pulses 803–816 (2014).
in a flint-glass photonic-crystal fiber,” Appl. Opt. 45, 6823–6830 26. D. McMorrow, W. T. Lotshaw, and G. A. Kenney-Wallace,
(2006). “Femtosecond optical Kerr studies on the origin of the nonlinear
330 Vol. 35, No. 2 / February 2018 / Journal of the Optical Society of America B Research Article

responses in simple liquids,” IEEE J. Quantum Electron. 24, 28. J. K. Ranka, R. S. Windeler, and A. J. Stentz, “Visible continuum gen-
443–454 (1988). eration in air-silica microstructure optical fibers with anomalous
27. J. H. Price, X. Feng, A. M. Heidt, G. Brambilla, P. Horak, F. Poletti, G. dispersion at 800 nm,” Opt. Lett. 25, 25–27 (2000).
Ponzo, P. Petropoulos, M. Petrovich, and J. Shi, “Supercontinuum 29. J. M. Dudley, G. Genty, and S. Coen, “Supercontinuum generation in
generation in non-silica fibers,” Opt. Fiber Technol. 18, 327–344 photonic crystal fiber,” Rev. Mod. Phys. 78, 1135–1184 (2006).
(2012).

You might also like