You are on page 1of 8

Optics Communications 364 (2016) 1–8

Contents lists available at ScienceDirect

Optics Communications
journal homepage: www.elsevier.com/locate/optcom

Dispersion properties of transverse anisotropic liquid crystal core


photonic crystal fibers
Naoki Karasawa
Chitose Institute of Science and Technology, 758-65 Bibi, Chitose 066-8655, Japan

art ic l e i nf o a b s t r a c t

Article history: The dispersion properties of liquid crystal core photonic crystal fibers for different core diameters have
Received 5 August 2015 been calculated by a full vectorial finite difference method. In calculations, air holes are assumed to be
Received in revised form arranged in a regular hexagonal array in fused silica and a central hole is filled with liquid crystal to
12 November 2015
create a core. In this study, three types of transverse anisotropic configurations, where liquid crystal
Accepted 13 November 2015
molecules are oriented in a transverse plane, and a planar configuration, where liquid crystal molecules
are oriented in a propagation direction, are considered. The large changes of the dispersion properties are
Keywords: found when the orientation of the liquid crystal molecules is changed from a planar configuration to a
Photonic crystal fiber uniform configuration, where all molecules are oriented in the same direction in a transverse plane.
Liquid crystal
Since the orientation of liquid crystal molecules may be controlled by applying an electric field, it could
Supercontinuum generation
be utilized for various applications including the spectral control of supercontinuum generation.
Finite difference method
& 2015 Elsevier B.V. All rights reserved.

1. Introduction infrared wavelength regions is relatively low [28]. The optical loss
due to the scattering in a nematic state is much higher than the
The generation of a broadband light pulse in a photonic crystal absorption loss in the visible region. However, it is reported that
fiber (PCF) using an optical pulse has been studied extensively [1]. the scattering loss can be reduced by filling LC in a glass capillary
Recently, liquid-filled PCFs for generating supercontinuum have to be less than about 3 dB/cm at wavelength 633 nm [29]. Thus it
been studied, where various liquids were selectively filled in PCFs is expected that LC core PCFs, whose lengths are about a few
[2–8]. Besides supercontinuum generation, liquid-filled PCFs and centimeters, can be utilized for generating supercontinuum in the
capillary fibers have been used to exploit their unique dispersive visible and near infrared regions [30]. Here, we consider the dis-
and nonlinear optical properties [9–14]. Various methods were persion properties of LC core PCFs because the generation of so-
proposed to fill liquid selectively in PCFs [15–18] and these liton pulses and the subsequent generation of dispersive waves are
methods may be used to create liquid crystal (LC) filled PCFs only possible when the GVD at the wavelength of an input optical
shown here by filling LC into a single hole in the clad region of pulse is negative. The ordinary and extraordinary refractive indices
PCFs with similar hole sizes and pitches shown here. of E7 LC considered here are higher than the refractive index of
LC filled PCFs have been studied for various applications. Since fused silica. Thus it is possible to make an optical fiber with an LC
the properties of LC filled PCFs can be controlled by changing core and fused silica clad. However, in that case, the effective re-
temperature or applying an electric field, these PCFs can be used as fractive index of the propagation mode has to be between the
variable optical filters [19–24] or sensors [25]. In addition to PCFs refractive index of the LC and that of fused silica, and GVD values
made of fused silica, LC filled PCFs made of polymer were reported become always positive except for the cases with very small core
[26]. When LC core PCFs are used for generating supercontinuum, diameters and long wavelengths. Thus, we consider the structure
it is expected that the spectra of the generated supercontinuum with maximum air fraction in clad such that the effective re-
can be controlled by changing temperature or by applying an fractive index can take a value close to the refractive index of air
electric field since the dispersion properties of the PCFs can be ( nair = 1). In practice, the director reorientation of LC molecules
controlled by these effects. The nonlinear refractive index for may occur due to the intense optical field and it may modify the
homogeneously aligned 5CB ( 4−cyano−4′ − n−pentylbiphenyl) LC GVD of LC core PCFs [31]. However, for ultrafast optical pulses
is measured to be more than 100 times larger than that of fused considered here for supercontinuum generations (pulse width
∼100 fs and peak power ∼104 W ) in transparent LC, very small
silica [27]. The optical absorption of LC in the visible and near
directorreorientation angle change, 2 × 10−6θss is expected where
θss is the steady state angle change by a continuous light (assuming
E-mail address: n-karasa@photon.chitose.ac.jp 20 mW power and the response time of LC to be 25 ms). Thus the

http://dx.doi.org/10.1016/j.optcom.2015.11.039
0030-4018/& 2015 Elsevier B.V. All rights reserved.
2 N. Karasawa / Optics Communications 364 (2016) 1–8

Table 1
The parameters for the extended Cauchy equations off E7 LC at 298 K used in
calculations for ordinary (no) and extraordinary (ne) refractive indices [42].

no ne

A 1.4994 1.6933
B 0.0070 0.0078
C 0.0004 0.0028

Table 2
Parameters used for finite difference calculations (in μm) of LC core PCFs for dif-
ferent hole diameters (d). In calculations, Δy was set to be equal to Δx .

d L Δx Δ

0.5 2.35 0.005 0.2


1.0 3.2 0.01 0.2
1.5 4.25 0.01 0.2
2.0 5.5 0.015 0.21

Fig. 1. The cross section of a LC filled PCF. Gray areas show air holes and a central
hole is filled with LC. The pitch Λ is set to be d + 0.1 μm in calculations, where d is
the hole diameter. Calculations were performed in a square region with a length L,
where perfectly matched boundary layers with widths Δ were used.

effect of director reorientation is not considered in this study. The


dispersion properties of liquid core PCFs were calculated using a
multipole method [32] and those of LC core PCFs were calculated
using the modified multipole method to treat anisotropic inclu-
sions rigorously for a planar configuration [30] to understand the
difference between the nematic and isotropic states since the
planar configuration was observed in experiment [22]. Previously,
the dispersion properties of LC core PCFs were calculated by a
perturbative scheme for the planar and the axial geometries of LC
[33]. Also, modal properties of LC filled PCFs were calculated by a
finite difference method [34,35] when an electric field was
applied.
In this study, dispersion properties of LC core PCFs are calcu-
lated by a mode solver for anisotropic waveguides based on a full
vectorial finite difference method [36] for the transverse aniso-
tropic configurations of LC, where all LC molecules are aligned in
the transverse plane of a PCF, as well as for a planar configuration.
The results of calculations by a finite difference method are com-
pared with those by a modified multipole method for a planar Fig. 3. Effective refractive indices ( neff,r ) versus wavelength for LC core PCFs with
different hole diameters for a planer (solid and dotted curves) and uniform (dot
configuration, since only uniaxial PCFs can be treated by a multi-
dashed curves) configurations. For planar configurations, both results by a multi-
pole method at present. When an electric field is applied in the pole method (solid curves) and those by a finite difference method (dotted curves)
perpendicular direction of a fiber, it is expected that LC molecules are shown. However, these curves overlap almost completely. Here, hole diameters
align in the same direction as the electric field, which we call a are 2.0, 1.5, 1.0, and 0.5 μm from top to bottom for both uniform and planar
uniform configuration. Since this configuration may be realized configurations.

experimentally, the changes of dispersion properties between

Fig. 2. The configurations of transverse anisotropic LC molecules in a central hole of a PCF for (a) uniform, (b) radial, and (c) circular director configurations.
N. Karasawa / Optics Communications 364 (2016) 1–8 3

Fig. 4. Electric fields near the central LC core of fundamental modes for planar (a) and uniform (b) configurations (d ¼ 2.0 μm at 800 nm). Both modes are HE-like.

treated in this study since these are fully anisotropic and cannot be
treated by a method used here. A finite-difference method for
arbitrary permittivity tensor is proposed [39], but the calculation
times are expected to be much longer than a method used here,
since both electric and magnetic fields have to be considered in
the fully anisotropic case. On the other hand, only magnetic fields
are considered in a transverse anisotropic case for calculating ei-
genvalues [36].
Here, GVD curves are calculated for LC core PCFs for different
hole diameters (d), where the pitches (Λ, the distance between
centers of neighboring holes) are set to be d + 0.1 μm such that the
air fractions in the clad are maximized to reduce the effective in-
dices of clad. In this way, the wall thickness between air holes was
thick enough to be manufactured in practice. For a fused silica core
PCF to be single moded, the air hole diameter to hole pitch ratio d/
Λ to be less than 0.41 [40]. All cases considered here, the LC core
PCFs are multi-moded since d/Λ is between 0.833 and 0.952. In
this study, dispersion properties of only fundamental modes are
considered, since the excitation of high order modes is believed to
be only significant when the excitation pulses into the fiber are not
properly matched to the fiber parameters. Spatial and temporal
Fig. 5. GVDs versus wavelength for LC core PCFs with different hole diameters (d)
distortions of the generated supercontinuum due to the nonlinear
for planar (solid and dotted curves) and uniform (dot dashed curves) configura- excitation of high order modes may be expected for a long fiber at
tions. For planar configurations, both results by a multipole method (solid curves) high input power as shown in a fused silica core PCF in [40]. For a
and those by a finite difference method (dotted curves) are shown. However, these 6-m-long fused silica core PCF with the nonlinear coefficient
curves overlap almost completely except for d ¼0.5 μm.
80 W−1 km−1, high order mode generation could be eliminated by
reducing the input power to ∼0.15 W [40]. For a few-cm-long LC
planar and uniform configurations are investigated. In addition to core PCF considered here, it is expected that high order mode
these configurations, other LC configurations have been con- generation can be eliminated by reducing the input power to a few
sidered theoretically and experimentally [37,38], including plane mW assuming that the nonlinear coefficient of LC is 100 times
radial, plane axial, and escaped radial configurations for a home- larger than fused silica.
otropic anchoring condition and plane circular, circular planar, and
escaped circular configurations for a circular anchoring condition.
In these configurations, plane radial and plane circular configura- 2. Methods
tions are transverse anisotropic and their dispersion properties are
investigated in this study. It is because in these configurations, the The PCF used in this study was assumed to be made of fused
LC directors are perpendicular to those of a planar configuration silica and it contained a central hole and two rings of air holes in a
and the differences of dispersion properties between different regular hexagonal lattice arrangement as shown in Fig. 1. The
configurations may be seen most clearly. On the other hand, the LC diameter d and the pitch Λ of air holes were selected to satisfy
directors of escaped configurations near the hole center are Λ = d + 0.1 μm , which means that the fraction of air in the clad is
identical to those of a planar configuration. Although escaped ra- maximized assuming the minimum wall thickness to be 0.1 μm.
dial and escaped circular configurations are more commonly ob- The orientations of LC directors in a transverse plane (xy plane)
served than radial and circular configurations, these are not of a PCF considered in this study are shown in Fig. 2. In Fig. 2(a),
4 N. Karasawa / Optics Communications 364 (2016) 1–8

Fig. 6. Effective refractive indices ( neff,r ) versus wavelength for LC core PCFs with different hole diameters (d). (a) HE-like modes (solid curves) and TM-like modes (dotted
curves) in radial configurations. (b) HE-like modes (solid curves) and TE-like modes (dotted curves) in circular configurations.

Fig. 7. Electric fields near the central LC core of HE-like (a) and TM-like (b) modes for a radial configuration (d¼ 2.0 μm at 800 nm).

the directors are aligned uniformly in x direction, which we call a ⎡ n2 0 0 ⎤


uniform configuration. This configuration may be achieved by ⎢ e ⎥
⎢ 0 no2 0 ⎥.
applying an electric field in the x direction. In Fig. 2(b), the di- ⎢ ⎥
rectors are aligned in a radial direction, which we call a radial ⎣ 0 0 no2 ⎦ (2)
configuration. In Fig. 2(c), the directors are aligned in a circular For a radial configuration, it is given by
direction, which we call a circular configuration. The radial con-
figuration may be observed when the surface condition favors a ⎡ n 2 sin2 θ + n 2 cos2 θ (n 2 − n 2 ) sin θ cos θ 0 ⎤
⎢ o e e o ⎥
homeotropic boundary condition, while the circular configuration ⎢ (ne2 − no2 ) sin θ cos θ no2 cos2 θ + ne2 sin2 θ 0 ⎥,
may be observed when the surface condition favors a circular ⎢ ⎥
⎣ 0 0 no2 ⎦ (3)
boundary condition.
In a central LC core, the relative dielectric constant tensor for a where θ is an angle measured from a horizontal (x) axis. Finally for
planar configuration is given by a circular configuration, it is given by
⎡ n2 0 0 ⎤ ⎡ n 2 sin2 θ + n 2 cos2 θ (n 2 − n 2 ) sin θ cos θ 0 ⎤
⎢ o ⎥ ⎢ e o o e ⎥
⎢ 0 no2 0 ⎥, ⎢ (no2 − ne2 ) sin θ cos θ ne2 cos2 θ + no2 sin2 θ 0 ⎥.
⎢ ⎥ ⎢ ⎥
⎣ 0 0 ne2 ⎦ (1) ⎣ 0 0 no2 ⎦ (4)

where no and ne are ordinary and extraordinary refractive indices In calculations, the Sellmeyer equation for fused silica [41] was
of LC, respectively. For a uniform configuration, it is given by used and the refractive index of air was assumed to be 1 for all
N. Karasawa / Optics Communications 364 (2016) 1–8 5

Fig. 8. Electric fields near the central LC core of HE-like (a) and TE-like (b) modes for a circular configuration (d ¼ 2.0 μm at 800 nm).

Fig. 9. GVDs versus wavelength for LC core PCFs with different hole diameters (d) for radial configurations. In (a), GVD curves for HE-like modes are shown and in (b), GVD
curves for TM-like modes are shown.

wavelengths. The LC considered here was E7 and the extended x¼ L and y¼ L boundaries, perfectly matched boundary layers were
Cauchy equations obtained by [42] were used. Here, the wave- used to eliminate the effects of boundaries to calculate modes
length (λ in μm) dependences of the extraordinary (ne) or ordinary accurately with a finite computational area [44]. In calculations,
(no) refractive indices were given using three parameters A, B, and the square region was divided by small regions with sides Δx and
C as follows (n is ne or no) Δy , in which the relative dielectric constant tensors were assumed
to be uniform. In our calculations, we set Δx and Δy to be uniform
B C
n=A+ + 4. except for a region inside perfectly matched boundary layers. In-
λ2 λ (5)
side a perfectly matched boundary layer ( x > L − Δ for x direction),
The values of parameters at 298 K are shown in Table 1. Δx was multiplied by 1 − 8i (x − (L − Δ))2 /Δ2. An identical scaling
In this study, the dispersion properties of a fundamental mode, was used for Δy for y > L − Δ. Because of the use of this complex
where the real part of its effective refractive index had the largest stretching factor, the calculated value of an effective refractive
value, were calculated for various LC configurations. In calcula- index became complex and its imaginary part represented the
tions, only a square region with sides L shown in Fig. 1 was used to confinement loss. Parameters used in calculations for different PCF
calculate modes accurately using symmetry properties of modes in structures were adjusted to obtain accurate dispersion properties
a PCF [43], which was also helpful to reduce the computational as shown in Table 2.
times. In the boundaries at x ¼0 or y¼0, either an anti-symmetric GVD at angular frequency ω was calculated by
or a symmetric boundary condition was used, where Hx was
d2β (ω) λ 3 d2neff,r (λ )
symmetric (anti-symmetric) and Hy was anti-symmetric (sym- β2 (ω) = = ,
dω 2 2π c 2 dλ 2 (6)
metric) for a symmetric (an anti-symmetric) boundary condition,
where Hx (Hy) was the x (y) component of a magnetic field. Near where β (ω) = neff,r (ω) ω/c is the propagation constant of a PCF,
6 N. Karasawa / Optics Communications 364 (2016) 1–8

Fig. 10. GVDs versus wavelength for LC core PCFs with different hole diameters (d) for circular configurations. In (a), GVD curves for HE-like modes are shown and in (b), GVD
curves for TE-like modes are shown.

Table 3 7.6 × 10−5, and 2.4 × 10−5 for d ¼0.5, 1.0, 1.5, and 2.0 μm, respec-
The zero dispersion wavelengths (in nm) of LC core PCFs for different configura- tively. The effective indices of uniform configurations are much
tions of LC molecules for different hole diameters (d). In radial and circular con- larger than those of planar configurations for the same hole
figurations, results of HE-like, TM-like, and TE-like modes are shown.
diameters.
d (μm) Planar Uniform Radial Circular In Fig. 4, electric fields of these configurations near the central
core are shown for d¼ 2 μm at 800 nm. As shown in this figure,
HE-like TM-like HE-like TE-like fundamental modes of these configurations are HE-like and line-
arly polarized. For a planar configuration, a vertically-polarized
0.5 – – – – – 1345, 1900
1.0 857, 1581 992 1253, 1459 – 1196, 1612 874, 1323 mode is degenerate with a horizontally-polarized mode. However,
1.5 961 1081 1465 860, 1463 1442 937 for a uniform configuration, a vertically-polarized mode is a higher
2.0 1060 1173 1754 926, 1960 1730 1010 order mode whose effective refractive index is much lower than
that of a horizontally-polarized fundamental mode. Since there are
many modes with higher effective refractive indices than a verti-
neff,r (λ ) is the real part of an effective refractive index at wave-
cally-polarized mode in short wavelengths in a uniform config-
length λ for each mode, and c is the speed of light.
uration, it is expected that a mode conversion from a vertically-
polarized mode to different modes occurs easily, which cause the
sudden change of dispersion properties. Thus, it will be important
3. Results and discussions to use a horizontally-polarized mode in practical applications in-
cluding supercontinuum generation.
Effective refractive indices ( neff,r ) versus wavelength of LC core In Fig. 5, GVD curves of LC core PCFs for fundamental modes of
PCFs with planar and uniform configurations for different hole planar and uniform configurations are shown. In this figure, results
diameters are shown in Fig. 3. In these configurations, funda- by a multipole method and those by a finite difference methods
mental modes are HE-like and electric fields are aligned in the are shown for a planar configuration. However, these curves are
same direction (x direction) of LC directors in uniform configura- overlapped almost completely except for the case d ¼0.5 μm at
tions. For HE-like modes, anti-symmetric boundary conditions wavelengths longer than 1200 nm. The maximum differences be-
were used at both x¼ 0 and y¼0 boundaries. For planar config- tween these methods are 7.2, 4.4, 2.7, and 1.9 fs2/mm for d ¼0.5,
urations, both results obtained by a multipole method and those 1.0, 1.5, and 2.0 μm, respectively. As shown in this figure, GVD
by a finite difference method are shown in the figure. But these values of uniform configurations are much larger than those of
curves are overlapped almost completely. The maximum differ- planar configurations at wavelengths shorter than 800 nm and
ences between these methods are 2.2 × 10−4 , 1.3 × 10−4 , zero dispersion wavelengths of uniform configurations are shifted

Table 4
The confinement losses (in dB/cm) of LC core PCFs at 1600 nm for different configurations of LC molecules for different hole diameters (d). In radial and circular config-
urations, results of HE-like, TM-like, and TE-like modes are shown. For d ¼ 0.5 μm, the losses at 800 nm are also shown in parentheses.

d (μm) Planar Uniform Radial Circular

HE-like TM-like HE-like TE-like

0.5 4.9 × 103 2.6 × 103 3.7 × 103 2.8 × 104 3.6 × 103 2.8 × 104
(1.5) ( 5.1 × 10−6 ) ( 1.2 × 10−1) ( 5.4 × 102 ) ( 6.8 × 10−2 ) ( 3.7 × 101)
1.0 1.4 4.6 × 10−3 2.0 × 10−1 6.4 × 102 1.2 × 10−1 4.3 × 10
1.5 1.4 × 10−4 5.5 × 10−9 5.0 × 10−7 1.0 × 10−2 6.0 × 10−8 1.2 × 10−4
2.0 1.3 × 10−8 4.3 × 10−14 3.1 × 10−12 2.3 × 10−9 1.3 × 10−13 3.9 × 10−11
N. Karasawa / Optics Communications 364 (2016) 1–8 7

to longer wavelengths significantly compared with those of planar parentheses). In this study, only two rings of air holes were used
configurations for the same d. and the losses at 1600 nm for d ¼0.5 μm were too large for prac-
In Fig. 6, effective refractive indices ( neff,r ) versus wavelength tical use. However, losses at 800 nm (and shorter wavelengths) are
are shown for radial and circular configurations. For radial con- less than a scattering loss (about 3 dB/cm [29]) except for radial
figurations (Fig. 6(a)), the fundamental modes for d larger than TM-like and circular TE-like modes. For the cases with larger hole
0.5 μm at short wavelengths were found to be TM-like modes diameters, confinement losses at 1600 nm (and shorter wave-
instead of HE-like modes, since the effective refractive indices of lengths) are much lower than a scattering loss except for radial
TM-like modes became larger than those of HE-like modes when TM-like mode and circular TE-like modes at d ¼1.0 μm. It is pos-
the wavelengths were decreased. The transition wavelengths from sible to decrease confinement losses by increasing the rings of air
HE-like to TM-like modes were 424, 588, and 763 nm for d ¼1, 1.5, holes and it may be necessary depending on the wavelength, the
and 2 μm, respectively. For circular configurations (Fig. 6(b)), the hole diameter, and the propagation mode for a few centimeter
fundamental modes for d larger than 0.5 μm at short wavelengths propagation.
were found to be TE-like modes instead of HE-like modes, since In summary, we have calculated dispersion properties of LC
the effective refractive indices of TE-like modes became larger core PCFs using a full vectorial finite difference method. For the LC
than those of HE-like modes when the wavelengths were de- core PCFs in planar configurations, calculated results by a finite
creased. The transition wavelengths from HE-like to TE-like modes difference method were compared with those by a multipole
were 549, 772, and 1009 nm for d ¼1, 1.5, and 2 μm, respectively. method modified for anisotropic inclusions and excellent agree-
In Fig. 7, electric fields near a central LC core for a radial con- ments were found. Three different transverse anisotropic config-
figuration are shown for both HE-like and TM-like modes in the urations of LC, namely, uniform, radial, and circular configurations,
case of d ¼2 μm at 800 nm. In calculations for radial configura- were considered, in which uniform configurations may be realized
tions, anti-symmetric boundary conditions were used at both x¼ 0 by applying a transverse external electric field. The large effects on
and y ¼0 boundaries for HE-like modes and an anti-symmetric dispersion properties of LC core PCFs were found by applying an
boundary condition at y¼ 0 and a symmetric boundary condition external electric field to obtain a uniform configuration of LC
at x¼ 0 were used for TM-like modes. As shown in Fig. 7(b), the molecules. The shifts of the zero dispersion wavelengths from
electric fields in a TM-like mode are parallel to the LC directors and planar to uniform configurations were more than 100 nm for the
it reduces the dielectric energy for short wavelengths, where the hole diameters larger than 1 μm. On the other hand, the shifts of
confinement of the field in the core is large. the zero dispersion wavelengths from either radial or circular
In Fig. 8, electric fields near a central core for a circular con- configurations to uniform configurations were about 200, 400, and
figuration are shown for both HE-like and TE-like modes in the 600 nm for d¼ 1, 1.5, and 2 μm, respectively. The directions of
case of d¼ 2 μm at 800 nm. In calculations for circular configura- shifts were positive when initial configurations were planar and
tions, anti-symmetric boundary conditions were used at both x¼ 0 were negative when initial configurations were either radial or
and y¼0 boundaries for HE-like modes and a symmetric boundary circular, which may be used to estimate the initial configurations
condition at y¼0 and an anti-symmetric boundary condition at of LC in a PCF. By selecting the wavelength of an input pulse, it may
x ¼0 were used for TE-like modes. As shown in Fig. 8(b), the be possible to design an LC core PCF, whose GVD changes sign
electric fields in a TE-like mode are parallel to the LC directors and when applying an external electric field to create a uniform con-
it reduces the dielectric energy for short wavelengths, where the figuration of LC. The large spectral modification of super-
confinement of the field in the core is large. continuum generated from the LC core PCF is expected when the
In Fig. 9, GVD curves for HE-like modes and TM-like modes for electric field is applied, since the mechanism of the super-
radial configurations are shown. As shown in this figure, the zero continuum generation depends on the sign of GVD significantly.
dispersion wavelengths for HE-like modes are much longer than
those of TM-like modes for the same d. Also, in Fig. 10, GVD curves
for HE-like modes and TE-like modes for circular configurations
References
are shown. As shown in this figure, the zero dispersion wave-
lengths for HE-like modes are much longer than those of TE-like
[1] M. Dudley, J.R. Taylor (Eds.), Supercontinuum Generation in Optical Fibers,
modes for the same d. The zero dispersion wavelengths for all Cambridge University Press, Cambridge, 2010.
configurations are summarized in Table 3. [2] R. Zhang, J. Teipel, H. Giessen, Theoretical design of a liquid-core photonic
From this table, the effects of applying a transverse electric field crystal fiber for supercontinuum generation, Opt. Express 14 (2006)
6800–6812.
to create uniform configurations of LC on the zero dispersion [3] R.V.J. Raja, A. Husakou, J. Hermann, K. Porsezian, Supercontinuum generation
wavelengths for various hole diameters (d) can be estimated as in liquid-filled photonic crystal fiber with slow nonlinear response, J. Opt. Soc.
follows. If the initial configuration of LC is planar, the zero dis- Am. B 27 (2010) 1763–1768.
[4] H. Zhang, S. Chang, J. Yuan, D. Huan, Supercontinuum generation in chloro-
persion wavelengths become longer more than 100 nm for d larger
form-filled photonic crystal fibers, Optik 121 (2010) 783–787.
than 1 μm. If the initial configurations are radial or circular, the [5] R.V.J. Raja, K. Porsezian, S.K. Varshney, S. Sivabalan, Modeling photonic crystal
zero dispersion wavelengths become shorter and the shifts are fiber for efficient soliton pulse propagation at 850 nm, Opt. Commun. 283
about 200, 400, and 600 nm for d ¼1, 1.5, and 2 μm, respectively, (2010) 5000–5006.
[6] A. Bozolan, C.J.S. de Matos, C.M.B. Cordeiro, E.M. dos Santos, J. Travers, Su-
where values of HE-like modes are used, since these are the fun- percontinuum generation in water-core photonic crystal fiber, Opt. Express 16
damental modes at zero dispersion wavelengths. Thus very large (2008) 9671–9676.
shifts of zero dispersion wavelengths can be expected when the [7] J. Bethge, A. Husakou, F. Mitschke, F. Noack, U. Griebner, G. Steinmeyer,
J. Herrmann, Two-octave supercontinuum generation in water-filled photonic
transverse electric field is applied to obtain uniform configurations crystal fiber, Opt. Express 18 (2010) 6230–6240.
of LC. Moreover, the direction and the amount of shifts depend [8] M. Vieweg, T. Gissibl, S. Pricking, B.T. Kuhlmey, D.C. Wu, B.J. Eggleton,
strongly on initial configurations of LC, which are determined by a H. Giessen, Ultrafast nonlinear optofluidics in selectively liquid-filled photonic
crystal fibers, Opt. Express 18 (2010) 25232–25240.
boundary condition of LC molecules at the hole surface. Thus the
[9] C.R. Rosberg, F.H. Bennet, D.N. Neshev, P.D. Rasmussen, O. Bang,
measurement of the shift of zero dispersion wavelengths may W. Krolikowski, A. Bjarklev, Y.S. Kivshar, Tunable diffraction and self-defo-
indicate an initial configuration of LC in a fiber. cusing in liquid-filled photonic crystal fibers, Opt. Express 15 (2007)
In Table 4, the calculated confinement losses (in dB/cm) for 12145–12150.
[10] P.D. Rasmussen, F.H. Bennet, D.N. Neshev, A.A. Sukhorukov, C.R. Rosberg,
various configurations at 1600 nm by a finite difference method W. Krolikowski, O. Bang, Y.S. Kivshar, Observation of two-dimensional non-
are shown (for d ¼0.5 μm, losses at 800 nm are also shown in local gap solitons, Opt. Lett. 34 (2009) 295–297.
8 N. Karasawa / Optics Communications 364 (2016) 1–8

[11] P.D. Rasmussen, A.A. Sukhorukov, D.N. Neshev, W. Krolikowski, O. Bang, [27] P. Palffy-Muhoray, H.J. Yuan, L. Li, M.A. Lee, J.R. DeSalvo, T.H. Wei, M. Sheik-
J. Lægsgaard, Y.S. Kivshar, Spatiotemporal control of light by Bloch-mode Bahae, D.J. Hagan, E.W. Van Stryland, Measurements of third order optical
dispersion in multi-core fibers, Opt. Express 16 (2008) 5878–5891. nonlinearities of nematic liquid crystals, Mol. Cryst. Liq. Cryst. 207 (1991)
[12] M. Vieweg, S. Pricking, T. Gissibl, Y.V. Kartashov, L. Torner, H. Giessen, Tunable 291–305.
ultrafast nonlinear optofluidic coupler, Opt. Lett. 37 (2012) 1058–1060. [28] S.-T. Wu, Absorption measurements of liquid crystals in the ultraviolet, visible,
[13] D. Churin, T.N. Nguyen, K. Kieu, R.A. Norwood, N. Peyghambarian, Mid-IR and infrared, J. Appl. Phys. 84 (1998) 4462–4465.
supercontinuum generation in an integrated liquid-core optical fiber filled [29] M. Green, S.J. Madden, Low loss nematic liquid crystal cored fiber waveguides,
with CS2, Opt. Mater. Express 3 (2013) 1358–1364. Appl. Opt. 28 (1989) 5202–5203.
[14] S. Kedenburg, A. Steinmann, R. Hegenbarth, T. Steinle, H. Giessen, Nonlinear [30] N. Karasawa, Dispersion properties of liquid crystal core photonic crystal fibers
refractive indices of nonlinear liquids: wavelength dependence and influence calculated by a multipole method modified for anisotropic inclusions, Opt.
of retarded response, Appl. Phys. B 117 (2014) 803–816. Commun. 338 (2015) 123–127.
[15] C. Martelli, J. Canning, K. Lyythikainen, N. Groothoff, Water-core Fresnel fiber, [31] I.C. Khoo, S. Zhao, Multiple time scales optical nonlinearities of liquid crystals
Opt. Express 13 (2005) 3890–3895. for optical-terahertz-microwave applications, Prog. Electromagn. Res. 147
[16] S. Yiou, P. Delaye, A. Rouvie, J. Chinaud, R. Frey, G. Roosen, P. Viale, S. Février, (2014) 37–56.
P. Roy, J.-L. Auguste, J.-M. Blondy, Stimulated Raman scattering in an ethanol [32] N. Karasawa, Dispersion properties of liquid-core photonic crystal fibers, Appl.
core microstructured optical fiber, Opt. Express 13 (2005) 4786–4791. Opt. 51 (2012) 5259–5265.
[17] Y. Wang, C.R. Liao, D.N. Wang, Femtosecond laser-assisted selective infiltration [33] P.D. Rasmussen, J. Lægsgaard, O. Bang, Chromatic dispersion of liquid-crystal
of microstructured optical fibers, Opt. Express 18 (2010) 18056–18060. infiltrated capillary tubes and photonic crystal fibers, J. Opt. Soc. Am. B 23
[18] F. Wang, W. Yuan, O. Hansen, O. Bang, Selective filling of photonic crystal fibers (2006) 2241–2248.
using focused ion beam milled microchannels, Opt. Express 19 (2011) [34] M.F.O. Hameed, S.S.A. Obayya, K. Al-Begain, M.I. Abo el Maaty, A.M. Nasr,
17585–17590. Modal properties of an index guiding nematic liquid crystal based photonic
[19] T.T. Larsen, A. Bjarklev, D.S. Hermann, J. Broeng, Optical devices based on li- crystal fiber, J. Lightwave Technol. 27 (2009) 4754–4762.
quid crystal photonic bandgap fibres, Opt. Express 11 (2003) 2589–2596. [35] G. Ren, P. Shum, X. Yu, J. Hu, G. Wang, Y. Gong, Polarization dependent guiding
[20] F. Du, Y.-Q. Lu, S.-T. Wu, Electrically tunable liquid-crystal photonic crystal in liquid crystal filled photonic crystal fibers, Opt. Commun. 281 (2008)
fiber, Appl. Phys. Lett. 85 (2004) 2181–2183. 1598–1606.
[21] M.W. Haakestad, T.T. Alkeskjold, M.D. Nielsen, L. Scolari, J. Riishede, H. [36] A.B. Fallahkhair, K.S. Li, T.E. Murphy, Vector finite difference modesolver for
E. Engan, A. Bjarklev, Electrically tunable photonic bandgap guidance in a li- anisotropic dielectric waveguides, J. Lightwave Technol. 26 (2008) 1423–1431.
quid-crystal-filled photonic crystal fiber, IEEE photon. Technol. Lett. 17 (2005) [37] B.V. Burylov, Equillibrium configuration of a nematic liquid crystal confined to
819–821. a cylindrical cavity, J. Exp. Theor. Phys. 86 (1997) 873–886.
[22] T.T. Alkeskjold, J. Lægsgaard, A. Bjarklev, D.S. Hermann, J. Broeng, J. Li, [38] G.P. Crawford, D.W. Allender, J.W. Doane, Surface elastic and molecular-an-
S. Gauza, S.-T. Wu, Highly tunable large-core single-mode liquid-crystal pho- choring properties of nematic liquid crystals confined to cylindrical cavities,
tonic bandgap fiber, Appl. Opt. 45 (2006) 2261–2264. Phys. Rev. A 45 (1992) 8693–8708.
[23] T.R. Wolinski, K. Szaniawska, S. Ertman, P. Lesiak, A.W. Domanski, [39] M.Y. Chen, S.M. Hsu, H.C. Chang, A finite-difference frequency-domain method
R. Dabrowski, E. Nowinowski-Kruszelnicki, J. Wojcik, Influence of temperature for full-vectorial mode solutions of anisotropic optical waveguides with an
and electrical fields on propagation properties of photonic liquid-crystal fi- arbitrary permittivity tensor, Opt. Express 17 (2009) 5965–5979.
bres, Meas. Sci. Technol. 17 (2006) 985–991. [40] Y. Vidne, M. Rosenbluh, Spatial modes in a PCF fiber generated continuum,
[24] J. Du, Y. Liu, Z. Wang, B. Zou, B. Liu, X. Dong, Liquid crystal photonic bandgap Opt. Express 13 (2005) 9721–9728.
fiber: different bandgap transmissions at different temperature ranges, Appl. [41] I.H. Malitson, Interspecimen comparison of the refractive index of fused silica,
Opt. 47 (2008) 5321–5324. J. Opt. Soc. Am. 35 (1965) 1205–1209.
[25] S. Mathews, G. Farrell, Y. Semenova, Liquid crystal infiltrated photonic crystal [42] J. Li, S.-T. Wu, S. Brugioni, R. Meucci, S. Faetti, Infrared refractive indices of
fibers for electric field intensity measurements, Appl. Opt. 50 (2011) liquid crystals, J. Appl. Phys. 97 (2005) 073501.
2628–2635. [43] F. Zolla, G. Renversez, A. Nicolet, B. Kuhlmey, S. Guenneau, D. Felbacq, Foun-
[26] W. Yuan, L. Wei, T.T. Alkeskjold, A. Bjarklev, O. Bang, Thermal tunability of dations of Photonic Crystal Fibres, Imperial College Press, London, 2005.
photonic bandgaps in liquid crystal infiltrated microstructured polymer op- [44] W.C. Chew, Complex coordinate stretching as a generalized absorbing
tical fibers, Opt. Express 17 (2009) 19356–19364. boundary condition, Microw. Opt. Technol. Lett. 15 (1997) 363–369.

You might also like