Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Research Article Vol. 14, No.

1 / 1 Jan 2024 / Optical Materials Express 178

Multi-layered cladding based ultra-low loss,


single mode antiresonant hollow core fibers
R UHANA N ISHAD , 1,3 L UTFUN N AHAR A SHA , 1 K UMARY S UMI R ANI
S HAHA , 1,2 A. B. M. A RAFAT H OSSAIN , 1 AND A BDUL
K HALEQUE 1,4
1 Department of Electrical & Electronic Engineering, Rajshahi University of Engineering & Technology,
Rajshahi-6204, Bangladesh
2 Department of Electrical, Electronic and Communication Engineering (EECE), Pabna University of
Science and Technology (PUST), Pabna-6600, Bangladesh
3 nishadruhana54@gmail.com
4 abdul.khaleque.eee@gmail.com

Abstract: In reality, an efficient platform for high-power laser delivery is highly important,
which can be justified by readily available fiber lasers, and hollow-core fiber can be the best
platform for guiding high optical power over long distances. The constraints include designing
new cladding geometry, which may lead to a higher laser induced damage threshold in the
fiber’s structure, having low losses along with a single mode nature. This article reports a new
antiresonant fiber that has a hollow core and a triple-layered cladding configuration with only
circular tube elements. The effects of multiple layers corresponding to the number of tube
rings in the cladding geometry are well explored, which leads to understanding the physical
insight of inter-layers. In comparison to double-layered cladding elements fiber, the proposed
fiber significantly reduces loss by an order of two and reveals a minimum leakage loss of
5.82 × 10−5 dB/km at the chosen wavelength of 1.06 µm through the proper arrangement of
cladding elements. We achieved a maximal higher order mode extinction ratio of about 104
(indicates the excellent single mode properties) and comparatively a little bending-induced loss
of 9.00 × 10−4 dB/km, when the radius of bending is 14 cm. As a result, our studies on new
multilayer fiber designs are not only useful for delivering high laser power but also serve as
guidelines for the experimental realization of future multilayered cladding fibers.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction
Specialized fibers, referred as hollow core (HC) fibers, limit light in the hollow interior known as
the core [1]. They are able to overcome many limitations of their solid core counterparts due
to their distinctive characteristics, which include low loss [2], an increased optical breakdown
threshold [3], an extended frequency range, reduced spectral broadening, low nonlinearity, and
many more. As HC fibers guide light in the hollow region, it is capable to minimize the material
loss that is influenced by host material. This guiding property also helps to significantly reduce
the fiber dispersion, and also minimize other types of losses such as surface scattering loss and
absorption induced loss [4,5]. As a result, they have garnered substantial attention and fascination
to the researchers. HC fibers (HCFs) have several applications in different fields, such as optical
fiber communications [6–8], an advanced high-energy beam delivery system [9], gas based light
source [10,11], polarization controlling optical devices [12,13], and terahertz guiding [14–16] as
a result of the aforementioned promising characteristics. Among the above application areas, it
has become an important issue for many researchers to guide and control laser lights with high
power, and they chose HC fibers as a better platform for this [17–21]. For instance, Goel et al.
used HC fiber platform for delivering high power [18].

#504207 https://doi.org/10.1364/OME.504207
Journal © 2024 Received 25 Aug 2023; revised 7 Dec 2023; accepted 8 Dec 2023; published 18 Dec 2023
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 179

The guidance characteristic divides HCFs into groups that are useful for obtaining certain
promising qualities [22]. For instance, hollow core photonic bandgap fibers (HC-PBGFs) exploit
the photonic bandgap (PBG) phenomenon to trap light within the air-centered core, resulting
in reduced transmission loss, constrained bandwidth, and increased dispersion [23]. In HCFs
with Kagome structures, the restricted bandwidth may be addressed, and decreased attenuation is
shown as a result of the hypocycloid form [24,25]. On the other hand, the HC antiresonant fiber
(ARF)’s inhibited guiding mechanism [26] largely depends on the design of their cladding. As a
result, light leakage from the HC is prevented by the cladding’s design, which creates a significant
difference in refractive indices between the core and cladding modes [5]. Therefore, HC-ARFs
overcome many limitations due to their advancement [5], including low dispersion along with
broad bandwidth, negligible power overlapping [22], low loss [27], etc. These extraordinary
features lead this fiber to many applications [28], especially in the area of high-power laser
delivery systems. For example, Gu and co-workers found record low loss of 4.3 dB/km on
HC-ARF platform for high power laser delivery at 1.0 µm [17]. Moreover, Zhu and co-authors
[29] experimentally investigated the laser induced damage threshold on the cladding geometry of
AR-HCF at 1 µm of wavelength. Therefore, one interesting way in this area is to engineer the
cladding elements of HC-ARFs.
Several cladding geometries are utilized on HC-ARFs for different applications (mostly for
communication purposes), and the geometry includes anisotropic anti-resonant (AR) nested
tubes [30], anti-resonant ice-cream cone-shaped tubes [31], elliptic anti-resonant tubes (ARTs)
[32], etc. As an example, Debord et al. [33] showed untouching eight circular cladding tube
elements in a HC-AR fiber with 7.7 dB/km of leakage loss (LL) at a wavelength of 0.75 µm.
Besides that, Gao and his colleagues have suggested a fiber that has a LL of 2.2 dB/km at the
wavelength of 1.512 µm [34]. In addition, in our previous research, Shaha et al. [35] proposed
a nested HC-AR fiber with double cladding elements that exhibits low optical loss, in the year
2020. In our recent works, it is seen that hybrid cladding elements (circular and elliptical) [6] and
modified conjoined tubes [8] in the cladding geometry of HC-ARFs have significantly improved
the loss performances. Moreover, Wang et al. [36] introduced a two ring ART with curvature of
inverse nature and showed what happened at the wavelength of 1.06 µm when an additional ring
was added for high power laser applications. Once more, Shaha and colleagues [30] reported a
minimal loss anisotropic HC-AR fiber arrangement for reducing the loss to 12 × 10−4 dB/km at an
operating wavelength of 1.25 µm. A conclusion can be drawn from the literature that an efficient
guiding medium is highly expected to transmit and control high power, and HC-ARF would be
the best platform for this. Although many researchers tried to solve the limitations in this fiber
platform (high leakage as well as bending losses, weak single mode characteristics, and so on),
the study of the effect of increasing the number of layers in the cladding on the HC-ARF platform
might be highly demanding. Hence, innovative structures with efficient cladding configurations
are eagerly anticipated in HC-ARFs for transmitting and controlling high-power laser light with
extremely low leakage and bending-induced loss with superior singular mode characteristics.
With a view to examining the impacts of multi-layer cladding elements on the fiber performances,
we studied triple cladding antiresonant hollow core fibers (TC-AR-HCFs) and proposed the
best low loss guiding possibility of high-power laser light. It offers extremely minimal leakage
loss, comparably minimal bending-induced loss, and excellent single mode performance. Our
focusing point is at the wavelength of 1.06 µm where high power lasers are available. We
achieved impressive low-level losses: LL and bending loss (BL) are reduced by two and three
order, respectively, by utilizing the positive impact of additional third ring or layer of AR tubes
surrounding the core because it gives additional negative curvatures to the fiber. A minimum
LL, 5.82 × 10−5 dB/km, is calculated at the wavelength of 1.06 µm, and it maintains a loss of
0.08 dB/km in between 0.95 µm and 1.2 µm (250 nm bandwidth). Moreover, this fiber has reduced
bending-induced loss (9 × 10−4 dB/km) at 1.06 µm wavelength with a bend radius of 14 cm. The
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 180

higher order mode extinction ratio (HOMER) formula is assessed for single-mode performance,
and so in this work, HOMER has an elevated value of 104 for a wavelength of 1.06 µm.

2. Fiber design and methodology


Figure 1 reports the studied TC-AR-HC fibers, in a two-dimensional platform, having three
cladding layers (i.e., three ART tubes ring). Instead of full geometry, half structures are shown
due to the easy presentation. Each of the four possible fiber geometries has a uniform core radius
of Dc = 30 µm and a similar thickness for all the circular elements of t = 0.42 µm. Besides that,
all the ARTs in three cladding rings are circular, and the diameters of first (near the core), second
(in the middle) and third (close to the outer capillary) rings are represented by d1 , d2 , and d3 ,
respectively. The core, which has a hollow region, is filled with air, and the background fiber
material is silica [37] in our fiber design. Figures 1(a)-1(d) show a schematic two-dimensional
arrangement of the studied three-layered circular shaped cladding ring geometries. In Figs. 1(a)
and (b), the fiber core is surrounded by a total of eight ARTs to form the first and second rings,
while the third ring consists of sixteen tubes connected to the outer capillary (connected to
the boundary). The total fiber diameter for the first fiber geometry is 79.74 µm (Fig. 1(a)). In
Fig. 1(b), the third ring ARTs are rotated by 45° in order to create more nodes or junctions in the
third layer. Here, the ratio of d1 /Dc and d2 /Dc for both the structures (Figs. 1(a) and (b)) are 0.57
and 0.70, respectively. The third ring tubes’ diameter ratio, d3 /Dc , is 0.29 for original unrotated
fiber structure (Fig. 1(a)) and 0.36 for rotated structure (Fig. 1(b)). The different ratio in third
ring tube diameters is due to the fixed total fiber diameter.

Fig. 1. Standardized cross-sectional representation of the considered TC-HC-ARFs. This


configuration contains four semi-structures of TC-HC-ARF having (a) eight tubes, and (c)
six tubes in the first ring (i.e., close to the core). In each of these diagrams, the core diameter
is uniform and represented as Dc , and the uniform thickness of the cladding AR tubes is
represented by t. Furthermore, d1 , d2 , and d3 correspond to the diameters of 1st , 2nd , and
3rd cladding ring, respectively. Third ring tubes are rotated by 45° in Figs. (b) and (d) in
order to see the effect of increased nodes in a multi-layer cladding configuration.

On the other hand, Figs. 1(c)-1(d) have six tubes in their first and second cladding rings, and the
third ring has twelve tubes connected to the outer capillary (close to the perfectly matched layer
(PML)). In Fig. 1(d), the structure is modified by rotating the third ring tubes by 45° compared
to the unrotated counterpart (Fig. 1(c)). Here, the ratio of d1 /Dc and d2 /Dc for both the rotated
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 181

and non-rotated structures, shown in Figs. 1(c) and 1(d), are 0.63 and 0.88, respectively. The
third ring tubes diameter with respect to the core diameter, d3 /Dc , is 0.60 and 0.73, respectively,
for our discussed unrotated and rotated (third ring tubes) cladding structures. In a nutshell, the
dimensions of all four structures studied (two have eight tubes and two have six tubes in both
the first and second ring) are uniform except for the diameter of the tubes, and the optimized
dimensions are summarized in Table 1. This is happening because we have tried to keep the fiber
core diameter fixed for the four considered structures which helps us to compare the performance
of the fibers. There is no overlap between the fiber core and any of the tubes. The positioning of
the ARTs in the first cladding ring is determined as [26]

Dc = (d1 + 2t + g)/sin(π/n) − (d1 + 2t) (1)

where Dc is the fiber core diameter, d1 is the fiber tube diameter for first ring tube, and n is the
number of tubes. The parameter g represents the distance or spacing between the tubes in the
first ring (close to the core). The tubes in the second and third rings naturally adhere to the
previously mentioned rules since their diameters are smaller compared to the diameter of the
tubes presented in the literature [26]. Taking into account the thickness of the tubes, denoted by t,
which is measured at 0.42 µm, and concentrated on the initial range of effective transmission,
the state of resonance is received approximately at a wavelength of 0.87 µm, as indicated by the
provided expression as [26] √︂
λ = 2t n2si − n2air /w1 (2)
where the refractive indices for silica and air are denoted as nSi and nair , respectively, and
the wavelength of the guided light is λ, while w1 represents the effective transmission range
possessing only counting or natural values. Here, silica is selected as the material for the tubes
surrounding the core (known as cladding tubes), and Sellmeier’s equation (material dispersion
is already included) is used for defining the properties of the fiber material [37]. Leakage loss,
which results from insufficiently confining light inside the fiber core, is a crucial loss that must
be estimated. The following equation can be used to determine this loss as [38]


(︃ )︃
LL = × 8.686 × Im (neff ) × 106 (dB/m) (3)
λ

where Im (neff ) is the imaginary coefficient of the effective refractive index (RI).

Table 1. Fine Tuned Structural Properties of the Studied Fibers


Fiber Type based on the 1st and 2nd ring tubes Dc t d1 /Dc D2 /Dc D3 /Dc
8-tubes structure 0.57 0.70 0.29
8-tubes rotated structure (45° rotation in 3rd ring tubes) 0.57 0.70 0.36
30 µm 0.42 µm
6-tubes structure 0.63 0.88 0.60
6-tubes rotated structure (45° rotation in 3rd ring tubes) 0.63 0.88 0.73

3. Numerical results and discussion


For modeling and numerical analysis of the proposed structures, commercial software namely
COMSOL Multiphysics is used, which works based on the finite element method (FEM). A
finite computational domain is selected by implementing a PML boundary that is responsible for
absorbing radiating light energy. The background fiber material (silica) is also selected for this
layer. A proper selection of mesh is also an important issue. A convergence experiment is first
conducted, as shown in Figs. 2(a) and 2(b), on the suggested eight tube non-rotated TC-HC-ARF
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 182

(Fig. 1(a)) at the operating wavelength of 1.06 µm to justify the numerical accuracy with respect
to the PML thickness and meshing. We have used the maximum mesh element size as λ/6 m
for silica, and λ/4 m for the air regions, respectively, where m is a mesh size parameter and
λ displays the working wavelength in our case. It is noted that the minimum mesh element
size is 1.68 nm here, and actual meshing will depend on the maximum to minimum element
size limit, the curvature factor, the element growth rate, etc. We have now varied m and PML
thickness to see the leakage loss stability only (Figs. 2(a) and 2(b)) because LL is a vital fiber’s
performance parameter and more sensitive than the others. After initial oscillations, it is seen
that a constant loss is observed beyond m ≥ 0.8; hence, m = 1 is used for further investigation to
save computational time without losing computational accuracy. Similarly, Fig. 2(b) shows the
LL execution in relation to the thickness of the PML. It can be observed that beyond a thickness
of greater than 5 µm, PML consistently provides a stable response (nearly constant loss) for the
designed fiber. This is the reason why a 10 µm PML thickness was employed for this work.

Fig. 2. The convergence experiment: leakage loss analysis of our selected fiber (8 tube
non-rotated geometry shown in Fig. 1(a)) at the wavelength of 1.06 µm for (a) utilizing the
maximum mesh size parameter (m) with a value of λ/6 m for the silica region and λ/4 m for
the air region, and (b) considering a PML thickness range of up to 10 µm as chosen for this
study.

In addition, for the justification of numerical study, we reproduced a result (red curve with
DR-NCF marked) from a related double cladding structure [36] and found good agreement over
the whole wavelength of interest as shown in Figs. 3(a) (the result of the Ref. paper [36] along
with inset structure) and 3(b) (our regenerated result of the Ref. [36]). We focused primarily on
the wavelength of 1.06 µm since fiber lasers with high-power are widely accessible here, and are
chosen as our operating wavelength. As a results, most of the results are optimized at 1.06 µm
(can be seen in Figs. 2(a), 2(b), and Fig. 4 as few examples).

3.1. Analysis of the optimization of the structures


In our research, we have tried to optimize our structures first to achieve the minimum leakage loss
at 1.06 µm of wavelength. The process of structure optimization is similar for the four considered
fiber geometries, and one structure (shown in Fig. 1(a)) is analysed here to clarify the steps. Our
first target is to fine-tune the value of d1 (tube diameter of the nearest ring around the core), and
the corresponding structure is inserted in Fig. 4(a). As a result, the fundamental mode (FM)
leakage loss is depicted in Fig. 4(a) as a function of d1 /Dc , which ranges from 0.3 to 0.6. When
the d1 /Dc ratio rises, the loss begins to fall, and it stays practically constant across a large range
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 183

Fig. 3. Justification of numerical accuracy by comparing our simulation results with the
results obtained in a related existing double cladding structure [36]: (a) The confinement
loss in dB/m (also known as leakage loss) of the Ref. paper [36] where the red color curve
represents their DR-NCF (structure is inset that we considered), although other curves are
for their remaining studied structures [36], (b) The reproduced confinement loss of Ref. [36]
vs. wavelength where the regenerated electric field surface of DR-ARF (in our COMSOL
environment) is inset in the plot. The frame color is correlated with the curve color.

Fig. 4. The optimization process of cladding tubes diameters of our proposed structure (in
Fig. 1(a)) by taking into consideration the leakage loss, at 1.06 µm, with respect to the (a)
d1 /Dc for first layer cladding, (b) d2 /Dc for second layer cladding, and (c) d3 /Dc for third
layer cladding.

as 0.45 < d1 /Dc <0.58. The ARTs intersect with one another at d1 /Dc = 0.60, and loss increases
due to the creation of additional nodes. As a result, the value of d1 /Dc cannot be raised further,
and the minimum leakage loss is attained at d1 /Dc = 0.57.
We then proceeded to introduce the second ring ARTs in the structure (inset of Fig. 4(b)) with
a diameter of d2 after setting d1 /Dc = 0.57. Figure 4(b) shows the leakage loss with regard to
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 184

d2 /Dc , the loss reduces as the d2 /Dc ratio rises, and the leakage loss is nearly constant from
d2 /Dc = 0.65 to d2 /Dc = 0.72. The leakage loss is decreased by three orders of magnitude (at 1.06
µm) due to the addition of second ring as seen in Fig. 4(b). At about d2 /Dc = 0.70, the loss is
settled at its lowest level; thereafter, it steadily rises as d2 /Dc grows in its value. In a similar way,
we added a third layer of ARTs by keeping the already optimized diameters of d2 /Dc at 0.70 and
d1 /Dc = 0.57 (the corresponding geometry can be seen inset of Fig. 4(c)). According to Fig. 4(c),
the leakage loss decreases quite sharply with the raises of d3 /Dc up to its value of 0.29, and
then the loss raises for the further increment of d3 /Dc . As a result, the minimum loss is further
lowered by three orders of magnitude (at 1.06 µm) which is 5.82 × 10−8 dB/m and the optimized
diameter ratio of the third layer of ARTs is selected at around d3 /Dc = 0.29. A summary of the
optimized structural parameters can be seen in Table 1. It is noted that LL is considered in dB/m
(instead of dB/km) in Fig. 4 which is because of the clear understanding of the low loss level
during the optimization process.

3.2. Analysis of leakage loss


For simplicity of analysis, the leakage loss performance is calculated first and analysed for the
four considered fibers. Figure 5 shows the LL spectra for the studied structures within the
wavelength range of 0.90 µm to 1.60 µm. Among those structures, the eight tubes unrotated
fiber has the lowest LL because of the additional number of layers (negative curvatures) and
limited nodes, which leads to a better anti-resonant reflection. In the lower wavelength range,
each of these fibers offer a low loss level, i.e., 5.82 × 10−5 dB/km at the operating wavelength of
1.06 µm, 3.28 × 10−4 dB/km at 1 µm, 7.86 × 10−4 dB/km at 1.06 µm, and 5.14 × 10−4 dB/km at
1 µm, respectively, for the structure having 8 tubes, 8 tubes 45° rotated, 6 tubes, and 6 tubes 45°
rotated cladding arrangement in the third ring of the geometry.

Fig. 5. Comparison of leakage loss performance among four studied structures within a
range of wavelengths, although our focusing point is at 1.06 µm. All the structures feature
an identical core radius (Rc = Dc /2 = 15 µm) and a strut thickness (t) of 0.42 µm. The
red, blue, black and green color curves correspond to the fiber structures having eight
tube non-rotated, eight tube rotated, six tube non-rotated, and six tube rotated cladding
arrangement, respectively.

Figure 5 displays a spectrum of leakage loss for the 8 tube nonrotated three layered cladding
components of the best TC-AR-HCF along with the other three counter configurations. Due to
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 185

the close proximity of the core and the outer capillary layer in the cladding, the 6-tube rotating
ARF recorded the greatest LL of all those structures, which is 1.55 × 10−3 dB/km (green color
dotted indicated) at 1.06 µm of wavelength. Then, within the wavelength of interest (Fig. 5),
the 8-tube rotated structure reported a second reduced LL (blue color dotted) compared to the
prior one (6 tube rotated one). The loss performance improved, in eight tube rotated fiber’s
structure, due to the increased number of negative curvatures in the cladding, but the number
of nodes creates limitation to achieve the best performance. Besides that, the absence of node
between the outer and second circular capillaries in the 6-tube non-rotated structure improves the
loss performance of the fiber and leads a comparatively third lower LL of 7.86 × 10−4 dB/km
at 1.06 µm (black color solid line). The above limitations have been improved in the 8-tube
non-rotated configuration due to the increased number of cladding tubes in the third ring along
with less number of nodes creation, which offers the lowest LL (5.82 × 10−5 dB/km, red color
solid line) among the others. To have a better visibility of the structural insight, the side panel of
Fig. 5 illustrates the four studied structure’s constructions with indicated dimensions where the
box colour correlates with the line color of the leakage loss spectra.
The performances based on LL at 1.06 µm among the four investigated structures are tabulated
in Table 2. Therefore, among all of the investigated structures, the proposed 8-tube non-rotated
TC-ARF (Fig. 1(a)) offers the minimal LL (red color solid line): the combination of three circular
tube rings results in the lowest loss because they offer more antiresonance strength by using more
negative curvatures. The best structure (shown in Fig. 1(a)) offers a LL of 5.82 × 10−5 dB/km at
a wavelength of 1.06 µm. Besides that, a LL of <∼ 0.08 dB/km and <10 dB/km are obtained
across a spectral width of 250 nm (from 0.95 µm to 1.06 µm) and 400 nm (from 0.90 µm to 1.30
µm), respectively, in our best optimized arrangement.
Table 2. Leakage Loss of the Compared Fiber Structures at 1.06 µm
Fiber Type Actual Description of Structure Type LL at 1.06 µm (dB/km)
8-tubes Eight tubes are around HC and nonrotated sixteen tubes in the 3rd 5.82 × 10−5
ring
8-tubes rotated Eight tubes are around HC and sixteen tubes, rotated by 45°, in the 4.85 × 10−4
3rd ring (close to the outer capillary)
6-tubes Six tubes are around HC and nonrotated twelve tubes are in the 3rd 7.86 × 10−4
ring
6-tubes rotated Six tubes are around core and twelve tubes, rotated by 45°, are in 1.55 × 10−3
the 3rd ring (close to the outer capillary)

3.3. Analysis of bending loss


A fiber should possess a natural resistance to the effects available in real-world applications. Due
to structural distortion and the coupling effect, the bending-induced loss in the HC-ARF emerges.
The subsequent mathematical expression is known as the equivalent refractive index model and
is applied to determine fiber bending-induced loss [35,36,39] along the x direction as

nb = nm (1 + x/Rb ) (4)

where nb implies the equivalent RI after bending, nm is the material refractive index of the
undistorted fiber. Moreover, Rb represents the bending radius (distance between fiber center to
the outer capillary under bending).
The bending-induced loss relative to the curved bend radius for the studied structures is shown
in Fig. 6(a) at 1.06 µm of wavelength. It is evident that as the curved bending radius increases,
the bending loss decreases, which holds true for all fiber types [5,20,26,36,38–41]. As a result of
the phenomenon known as the anti-crossing effect [6,26,35,36], it is shown that the 6-tube rotated
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 186

(45°) configuration has the highest BL (green color dotted line) among all. The maximum BL
was recorded at a curved bend radius of 4 cm because of the tendency for coupling between this
structure’s core mode and cladding modes under bending state. The 6-tube’s (without rotation)
structure thus offers comparatively better BL (black colour solid line) than the 6-tube turned
(45°) structure, but it is still quite high. Better BL than the previous two constructions is provided
by the 8-tube rotated structure (blue colour dotted line curve) and 8-tube non-rotated cladding
arrangements (red colour solid line), although both suffer from anti-crossing effects at 3 cm
(only 8-tube non-rotated one), 6 cm, and 9 cm of bend radius. Although the 8-tube rotated one
provides better BL in the higher bending regions (bend radius of less than 14 cm), eventually, the
suggested fiber (red color solid line) showed the best BL among the evaluated fiber architectures,
thanks to a limited air zone in the cladding and inhibited coupling propensity between core mode
and cladding modes.

Fig. 6. (a) Comparison of the bend radius-related bending loss performance of various
constructions studied in this research. The same strut thickness, t = 0.42 µm and core radius,
Rc = 0.15 µm, are employed. The corresponding field profiles are depicted in right side
panel for some specific bend radius of (b) 2 cm, (c) 4 cm, (d) 6 cm, and (e) 8 cm, where field
profile’s box color is matched with line color of the BL curve of the optimized structure only
(red color).

The bending loss curve exhibits multiple peaks at different bending radii before reaching a bend
radius value of 10 cm. This behaviour is attributed to the influence of the anti-crossing effect [24].
Notably, at a bend radius of 14 cm, the bending loss measures approximately 9 × 10−4 dB/km and
remains less than 9 × 10−2 dB/km and 6 × 10−3 dB/km of loss while the bend radius exceeds 7 cm
and 11 cm, respectively. Moreover, with a curved bend radius of 20 cm, a BL of 0.000272 dB/km
is attained. As a result of the debilitated propensity of coupling between the central (core) mode
and surrounding (cladding) modes, it exhibits three increasing peaks at bend radii of 3 cm, 6 cm,
and 9 cm because of the structural distortion that occurs due to too much bending. It is noted that
a smaller bend radius indicates larger fiber bending. However, the loss reduces monotonically as
the bend radius grows (for example, a decreasing loss tendency can be seen after 9 cm of bend
radius in Fig. 6(a)). Hence, the fiber with eight tube non-rotated structure performs better than
the remainder equivalent. So, for the remaining discussion, we will focus on the eight tubes’
non-rotated construction only. In the right-side panel, Figs. 6(b)–6(e) shows the corresponding
electric field profiles of our finalized structure at the bending radius of 2 cm, 4 cm, 6 cm, and 8 cm,
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 187

respectively, to clearly understand the bending loss behaviour of the best performed structure
(red colour marked). It is seen that field profile distorted much (core region of Fig. 6(b)) because
of the too much bending radius of 2 cm which is expected. The distortion in field profile reduces
as the bending radius increases further (4 cm to 8 cm) as can be seen in Figs. 6(c) to 6(e).
It is also crucial to examine the bending-induced loss behaviour of our proposed best structure
(eight tubes non-rotated structure) in relation to the relevant wavelength. Figure 7 displays the
variation in bending loss as a function of wavelength for different values of Rb (bending radius).
The bending-induced loss is illustrated in this study with a given curved bend radius of 14 cm. It
is chosen as 14 cm because the bending loss decreases after 9 cm bend radius and becomes stable
in 12 cm, which provide stable loss up to 20 cm. To be more specific, it is an arbitrary bend
radius chosen just for discussion, we can choose any bend radius from the range of 2 cm to 20 cm
(the similar range can be found in the literature [6,26,32,36]). The lowest loss (< 10−3 dB/km)
is obtained at the wavelength of 1.06 µm as expected. The loss is still less than 0.087 dB/km
between wavelengths of 1.00 µm and 1.22 µm. The presence of the silica probe connection (the
connection between two silica tube layers, i.e., the junction between one-layer tubes to other layer
tubes) results in Fano-resonance, leading to the specific oscillations observed in the bending
loss spectra (Fig. 7) [8,39]. There is some abrupt change in the BL curve, especially at the
wavelengths of around 1.14 µm, 1.20 µm, 1.25 µm, and 1.35 µm which is because of structural
distortion due to anticrossing effect [24,35,36] happened in the numerical simulation at some
particular critical point.

Fig. 7. The behaviour of bending-induced loss of the best designed structure in relation to
the wavelength. In this analysis, a constant curved bend radius of Rb = 14 cm is employed
for our chosen best fiber structure (actual schematic is in Fig. 1(a)).

3.4. Analysis of single mode performance


The ratio between the minimal higher order modes (HOMs) loss and the FM loss is known
as HOMER, which stands for single mode properties of a fiber [32]. The fiber can effectively
operate as a single mode if HOM losses are sufficiently greater than the basic mode (FM) loss. In
our fiber, FM is HE11 and HOMs are considered as TM01 , HE21 , TE01 , HE31 , and EH11 (modes
can be differentiated with the help of different color’s legend). According to Fig. 8(a), FM (HE11 )
has the lowest loss level, while the TE01 mode exhibits the least level of loss among all the
HOMs. It is noted that the above explanation regarding the lowest loss level of FM and HOMs
is valid over the whole wavelength of interest although our focusing point specially is around
1.06 µm due to the high-power laser delivery. As a result, the wavelength range that is being
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 188

studied is restricted to 1.00 µm to 1.30 µm. The corresponding electric field patterns of HOMs
and FM are presented at the apex of Fig. 8, where the box color of field profile corresponds with
the line color of the loss spectra. Figure 8(b) displays the nature of HOMER variation for the
change in wavelength. As the wavelength increases, it is seen that the value of HOMER keeps
its high value (> 103 ) until it reaches its maximal level (8590) at around 1.06 µm. After that,
the HOMER maintains much better than the well-known single mode standard (HOMER = 10
[6,8,36]). The aforementioned explanation leads to the conclusion that the provided device can
function as a top-notch single mode fiber. The other three studied structures also show good
HOMER performance, but the selected one provides the best performance by considering other
loss properties.

Fig. 8. (a) Spectra regarding leakage loss of HOMs and FM showcasing dependency on
wavelength variation, and (b) HOMER versus wavelength, specifically focusing on the FM
of HE11 and TE01 mode due to its minimal loss compared to other HOMs. The top section
showcases the field profiles of HOMs and FM, visually represented with matching colors.
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 189

3.5. Comparative analysis of our achievements


Table 3 displays a comparative behavioral analysis of our proposed fiber with relevant and current
research on the platform of HC-ARF. One such example is the solitary ring, eight-tube AR-HC
fiber reported in [33], which exhibits a two orders higher fiber bending-induced loss (compared
to our case) of 3 × 10−2 dB/turn at Rb (bend radius) of 15 cm and a considerable higher loss of
7.70 dB/km in the wavelength area of 0.75 µm.

Table 3. Comparative Summary of our Fiber Performances with the Existing Related and Recent
Literatures
Ref. Dc (µm) Minimum LL (λ) Minimum LL Range HOMER Bending Loss
(Yr)
[33] 27.5 to 7.7 dB/km (0.75 µm) 10-20 dB/km over 600 nm —– 0.03 dB/turn at
2017 37.9 Rb = 15 cm
[22] 30.5 2 dB/km (1.512 µm) 16 dB/km over 335 nm —– < 1 dB/km at
2018 Rb = 10 cm
[41] 40 0.05 dB/km (1.35 µm) <1 dB/km over 400 nm —– —–
2019
[35] 33 0.001 dB/km (1.40 <0.10 dB/km over 320 nm ∼103 ∼0.01 dB/km at
2020 µm) Rb = 8 cm
[36] 30 0.01 dB/km (1.06 µm) —– ∼103 ∼0.01 dB/km at
2020 Rb = 10 cm
[42] 30 0.00369 dB/km (1.06 <1 dB/km over 800 nm ∼1400 0.4405 dB/km at
2021 µm) Rb = 5 cm
[43] 26 3.66× 10 −3 dB/m <0.4 dB/m over 45 nm —– 0.03 dB/m at Rb = 6 cm
2022 (1.55 µm)
Prop. 30 5.82 × 10−5 dB/km <∼0.08 dB/km over 250 nm ∼104 ∼0.0009 dB/km at
2023 (1.06 µm) Rb = 14 cm

Moreover, Chen and colleagues [41] reported a double negative curvature AR-HCF at 1.35
µm and improved the level to 0.05 dB/km (still far away from the current reported loss level)
with superior bandwidth of 400 nm by maintaining a loss level of < 1.0 dB/km. In 2020 [36], a
remarkable HOMER of 103 was discovered, which is lower than the reported HOMER here by
one orders of magnitude, while Wang’s and co-authors demonstrated a significant reduction in
leakage loss, achieving an impressive value of 0.01 dB/km (which is much higher compared to
the current value) at the corresponding wavelength of 1.06 µm. Importantly, nearly two orders of
magnitude higher LL and bending losses are observed in very recent research outcomes [42,43].
There are more recent and related fibers [2,30,34,35] in the literature (not summarized in Table 3)
that have attempted to improve leakage loss, bending loss, and single mode properties. For
instance, in [35], an enhanced performance is seen in terms of higher loss (10−3 dB/km) with a
similar HOMER of 103 . Therefore, it is clear from the above summary that the suggested fiber
may be able to perform significantly better than the reported fibers due to its minimal leakage
loss (5.82 × 10−5 dB/km), high HOMER ( 104 ), and better bending-induced loss (0.0009 dB/km).

3.6. Analysis of fabrication feasibility


From a manufacturing perspective, it is crucial to assess the variation in structural parameters to
ensure the perspective of the achieved results throughout the fabrication process. The presented
fiber is made with three layers of cladding circular tubes, and it is necessary to consider the
overlapping of the tubes at the junction points. We are considering this overlapping or penetration
as an ingression length (p) in percent. The ingression of the tubes is happened between two
successive layers i.e., rings. We have increased this length from ideally optimized (t/10) to 10%
of its ideal value. The corresponding geometry of every considered ingression length is depicted
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 190

in the right panel of Fig. 9 and can be identified by color matching between box (right side) and
line curve. It is noted that we have considered ingression in both layer’s node uniformly which
can be visualized from two geometries located in each row. For example, two geometries marked
as black color (top one) represent zoom view of nodes created in between first (around the core)
to second (middle) ring and second (middle) to the third (close to the outer capillary) ring of our
optimized structure with initial ingression (t/10), respectively. The following black, green, blue
color box are the corresponding geometries for the ingression length (p) of 4%, 6%, and 10%,
respectively.

Fig. 9. Leakage loss curve for the fabrication test with respect to wavelength where the
error is shown in two layers tube ingression with the ideal one. Here, p is ingression length
in percent due to ingression of tubes between two consecutive rings.

Figure 9 depicts the leakage loss as a function of ingression length variation with respect to
wavelength. From Fig. 9, it is perceived that the LL is elevated with the increment of ingression
length over the wavelength, thus the edge length has a negative impact on the fiber’s performance.
Due to the increase of ingression length (p) from its optimum value, the LL increases gradually.
This happens because increasing ingression results in a wider nodes area, and those nodes lead to
light leakage through the cladding, which results in high leakage loss. However, up to 10% of
ingression of the tubes, the LL is maintaining less than 0.001 dB/km which is still a significantly
lower loss than the existing reported loss [22,33,41]. Hence, it is concluded that the fiber can
maintain the lower loss performance after suffering enough ingression tolerance. This tolerance
level may be useful if the ingression is created from cladding tube pressure as well as pull and
feed speeds during the drawing process of fabrication [36].
At the time of fabrication, the thickness of the structure might be deviated. So, by increasing
and decreasing the values of the thickness, t, we can evaluate the performance of the selected
structure. By comparing the LL spectra of ideal (optimized) tube thickness with −2% to +2% of
the ideal tube thickness, the tolerance of tube thickness, t is observed thoroughly as shown in
Fig. 10(a). When there is an increase of 2% in the tube thickness, the LL initially increases due
to the change in resonance wavelength because resonance wavelength is highly sensitive to the
tube thickness [6,35,36]. From Fig. 10(a), it is noticed that a relatively limited oscillation in the
loss curve is observed for the optimized t and for its variation from −2% to +2%: the pattern
looks almost the same, and in higher wavelength regions, the loss variation is nearly the same.
The LL spectra are suffering from oscillation due to Fano resonance [35,36] that occurs at the
node between ARTs. Therefore, it is concluded that, after the postprocessing, the fiber can obtain
its desired performance.
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 191

Fig. 10. (a) Leakage loss variation for the tube thickness (t) variation of +2% to −2% for
finding the effect on real time application with respect to wavelength comparing the results
with optimized tube thickness. (b) Leakage loss performance with respect to wavelength
while the core diameter (Dc ) is varied from + 2% to −2% along with its optimized core
diameter.

At the end, from the real point of view of fabrication, we go through the analysis of tolerance
for the fiber core diameter (Dc ). We vary the core diameter from −2% to +2% of its optimized
value and summarized the effects in terms of LL over the specified wavelength range (Fig. 10(b)).
It is seen that a bit of oscillation is observed in the LL spectra, and the amount of loss variation is
within a tolerated level. Because, the oscillation originally observed in the loss curves of every
multi-layered structure, as explained earlier.
Moreover, one pioneering work which is based on second cladding layer is successfully
fabricated by Huang and co-workers [44]. Another related structure namely split cladding ARF
having triple cladding has already been fabricated in previous days [45]. On the other hand,
conjoined tube based antiresonant fiber was found 0.8 dB/km as simulated loss, and 2 dB/km loss
was found after fabrication at 1.55 µm [34]. Thus, we have chosen a tolerance level of ±2% by
following previous literature [5,34,44,45] and found negligible difference with the optimized
one. This means that the structure could tolerate more. It is noted that we have also discussed
overlapping or penetration as an ingression length (p) in percent which may be helpful regarding
the fabrication challenge. Therefore, we hope that the proposed structure may maintain its
numerical achievements in its fabrication stages along with few postprocessing.

4. Conclusions
In order to reveal the physical insight of multilayered cladding in ARF, the design of a few new
multilayer cladding based AR-HCFs are studied, analysed, and the best proposal is obtained
through the finite element numerical method. The effect of nodes and the number of layers in the
cladding are clearly investigated and as a result, the leakage loss reduces significantly by a few
orders of magnitude. Numerical result shows that the leakage loss in the presented fiber is reduced
by two and three orders of magnitude, respectively, compared to the double-layered cladding
nested and double-layered negative curvature AR-HCFs (can be seen in Table 3). Additionally, a
better bending-induced loss and excellent single mode performance (HOMER = 104 ) is obtained.
We believe that the proposed low losses multilayered cladding based single mode fiber hold
enough potential for high power fiber laser systems.
Funding. Department of Research & Extension, Rajshahi University of Engineering & Technology (RUET) through
the University Grant Commission (UGC), Bangladesh (DRE/7/RUET/640(53)/PRO/2023-24/14).
Acknowledgments. The authors acknowledged the computational and other logistic support provided by the
Department of Electrical & Electronic Engineering (EEE), Rajshahi University of Engineering & Technology (RUET),
Bangladesh. We also want to thank Md. Sarwar Hosen (EEE, RUET) for his suggestions during the period of this
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 192

research. Dr. Abdul Khaleque wants to acknowledge the financial support (DRE/7/RUET/640(53)/PRO/2023-24/14)
provided by Research & Extension of RUET through the University Grant Commission (UGC) of Bangladesh. Moreover,
Kumary Sumi Rani Shaha wants to acknowledged the support provided by the Department of Electrical, Electronic and
Communication Engineering (EECE), Pabna University of Science and Technology (PUST), Pabna-6600, Bangladesh.
Disclosures. The authors declare no conflicts of interest.
Data availability. Data underlying the results presented in this paper are not publicly available at this time but may
be obtained from the authors upon reasonable request.

References
1. R. F. Cregan, B. J. Mangan, J. C. Knight, T. A. Birks, P. S. J. Russell, P. J. Roberts, and D. C. Allan, “Single-mode
photonic band gap guidance of light in air,” Science 285(5433), 1537–1539 (1999).
2. K. S. R. Shaha, A Khaleque, and M. I. Hasan, “Nested antiresonant hollow-core fiber with ultralow loss,” in 2020
11th International Conference on Electrical and Computer Engineering (ICECE), IEEE, pp. 29–32 (2020).
3. A. N. Kolyadin, A. F. Kosolapov, A. D. Pryamikov, A. S. Biriukov, V. G. Plotnichenko, and E. M. Dianov, “Light
transmission in negative curvature hollow core fiber in extremely high material loss region,” Opt. Express 21(8),
9514–9519 (2013).
4. D. Wu, F. Yu, and M. Liao, “Understanding the material loss of anti-resonant hollow-core fibers,” Opt. Express 28(8),
11840–11851 (2020).
5. F. Poletti, “Nested antiresonant nodeless hollow core fiber,” Opt. Express 22(20), 23807–23828 (2014).
6. K. S. R. Shaha, A. Khaleque, and M. S. Hosen, “Wideband low loss hollow core fiber with nested hybrid cladding
elements,” J. Lightwave Technol. 39(20), 6585–6591 (2021).
7. M. S. Hosen, A. Khaleque, K. S. R. Shaha, R. Nishad, and A. S. Sultana, “Highly Birefringent Low Losses
Hollow-Core Antiresonant Fiber,” in 2021 3rd International Conference on Electrical & Electronic Engineering
(ICEEE), IEEE, pp. 141–144 (2021).
8. K. S. R. Shaha and A. Khaleque, “Low-loss single-mode modified conjoined tube hollow-core fiber,” Appl. Opt.
60(21), 6243–6250 (2021).
9. G. Humbert, J. C. Knight, G. Bouwmans, P. S. J. Russell, D. P. Williams, P. J. Roberts, and B. J. Mangan, “Hollow
core photonic crystal fibers for beam delivery,” Opt. Express 12(8), 1477–1484 (2004).
10. P. S. J. Russell, P. Hölzer, W. Chang, A. Abdolvand, and J. C. Travers, “Hollow-core photonic crystal fibres for
gas-based nonlinear optics,” Nat. Photonics 8(4), 278–286 (2014).
11. M. I. Hasan, N. Akhmediev, A. Mussot, and W. Chang, “Midinfrared pulse generation by pumping in the normal
dispersion regime of a gas-filled hollow-core fiber,” Phys. Rev. Appl. 12(1), 014050 (2019).
12. K. S. R. Shaha, A. Khaleque, M. T. Rahman, and M. S. Hosen, “Broadband and short-length polarization splitter on
dual hollow-core antiresonant fiber,” IEEE Photonics Technol. Lett. 34(5), 259–262 (2022).
13. K. S. R. Shaha, A. Khaleque, and M. S. Hosen, “Dual-Core Antiresonant Fiber Based Compact Broadband Polarization
Beam Splitter,” in 25th International Conference on Computer and Information Technology (ICCIT), IEEE, pp.
254–259 (2022).
14. J. Anthony, R. Leonhardt, S. G. Leon-Saval, and A. Argyros, “THz propagation in kagome hollow-core microstructured
fibers,” Opt. Express 19(19), 18470–18478 (2011).
15. A. S. Sultana, A. Khaleque, K. S. R. Shaha, M. M. Rahman, and M. S. Hosen, “Nodeless antiresonant hollow core
fiber for low loss flatband THz guidance,” Opt. Continuum 1(8), 1652–1667 (2022).
16. A. Rahman, A. Khaleque, M. Y. Ali, and M. T. Rahman, “THz spectroscopic sensing of liquid chemicals using a
photonic crystal fiber,” OSA Continuum 3(11), 2982–2996 (2020).
17. S. Gu, X. Wang, H. Jia, Z. Lian, X. Shen, Y. Mai, and S. Lou, “Single-ring hollow-core anti-resonant fiber with a
record low loss (4.3 dB/km) for high-power laser delivery at 1 µm,” Opt. Lett. 47(22), 5925–5928 (2022).
18. C. Goel, H. Li, M. R. A. Hassan, W. Chang, and S. Yoo, “Anti-resonant hollow-core fiber fusion spliced to laser gain
fiber for high-power beam delivery,” Opt. Lett. 46(17), 4374–4377 (2021).
19. X. Zhu, D. Wu, Y. Wang, F. Yu, Q. Li, Y. Qi, J. Knight, S. Chen, and L. Hu, “Delivery of CW laser power up to 300
watts at 1080 nm by an uncooled low-loss anti-resonant hollow-core fiber,” Opt. Express 29(2), 1492–1501 (2021).
20. H. Li, C. Goel, J. Zang, S. Raghuraman, S. Chen, M. R. A. Hassan, W. Chang, and S. Yoo, “Integration of an
anti-resonant hollow-core fiber with a multimode Yb-doped fiber for high power near-diffraction-limited laser
operation,” Opt. Express 30(5), 7928–7937 (2022).
21. H. Zhang, Y. Chang, Y. Xu, C. Liu, X. Xiao, J. Li, X. Ma, Y. Wang, and H. Guo, “Design and fabrication of a
chalcogenide hollow-core anti-resonant fiber for mid-infrared applications,” Opt. Express 31(5), 7659–7670 (2023).
22. F. Yu and J. C. Knight, “Negative curvature hollow-core optical fiber,” IEEE J. Sel. Top. Quantum Electron. 22(2),
146–155 (2016).
23. B. J. Mangan, L. Farr, A. Langford, P. J. Roberts, D. P. Williams, F. Couny, M. Lawman, M. Mason, S. Coupland, R.
Flea, H. Sabert, T. A. Birks, J. C. Knight, and P. S. J. Russell, “Low loss (1.7 db/km) hollow core photonic bandgap
fiber,” in Optical Fiber Communication Conference, Optical Society of America, pp. 3, vol. 2 (2004).
24. F. Benabid, J. C. Knight, G. Antonopoulos, and P. S. J. Russell, “Stimulated raman scattering in hydrogen-filled
hollow-core photonic crystal fiber,” Science 298(5592), 399–402 (2002).
Research Article Vol. 14, No. 1 / 1 Jan 2024 / Optical Materials Express 193

25. Y. Y. Wang, N. V. Wheeler, F. Couny, P. J. Roberts, and F. Benabid, “Low loss broadband transmission in
hypocycloid-core kagome hollow-core photonic crystal fiber,” Opt. Lett. 36(5), 669–671 (2011).
26. C. Wei, R. J. Weiblen, C. R. Menyuk, and J. Hu, “Negative curvature fibers,” Adv. Opt. Photonics 9(3), 504–561
(2017).
27. D. Wu, F. Yu, C. Wu, M. Zhao, J. Zheng, L. Hu, and J. Knight, “Low-loss multi-mode anti-resonant hollow-core
fibers,” Opt. Express 31(13), 21870–21880 (2023).
28. W. Ding, Y. Y. Wang, S. F. Gao, M. L. Wang, and P. Wang, “Recent progress in low-loss hollow-core anti-resonant
fibers and their applications,” IEEE J. Sel. Top. Quantum Electron. 26(4), 1–12 (2020).
29. X. Zhu, F. Yu, D. Wu, S. Chen, Y. Jiang, and L. Hu, “Laser-induced damage of an anti-resonant hollow-core fiber for
high-power laser delivery at 1 µm,” Opt. Lett. 47(14), 3548–3551 (2022).
30. K. S. R. Shaha, A. Khaleque, and M. T. Rahman, “Low loss anisotropic nested hollow core antiresonant fiber,” in
2020 2nd International Conference on Advanced Information and Communication Technology (ICAICT), IEEE, pp.
71–76 (2020).
31. F. Yu, W. J. Wadsworth, and J. C. Knight, “Low loss silica hollow core fibers for 3-4 µm spectral region,” Opt.
Express 20(10), 11153–11158 (2012).
32. M. S. Habib, O. Bang, and M. Bache, “Low-loss single-mode hollow-core fiber with anisotropic anti-resonant
elements,” Opt. Express 24(8), 8429–8436 (2016).
33. B. Debord, A. Amsanpally, M. Chafer, A. Baz, M. Maurel, J. M. Blondy, E. Hugonnot, F. Scol, L. Vincetti, F.
Gérôme, and F. Benabid, “Ultralow transmission loss in inhibited-coupling guiding hollow fibers,” Optica 4(2),
209–217 (2017).
34. S. F. Gao, Y. Y. Wang, W. Ding, D. L. Jiang, S. Gu, X. Zhang, and P. Wang, “Hollow-core conjoined-tube
negative-curvature fibre with ultralow loss,” Nat. Commun. 9(1), 2828 (2018).
35. K. S. R. Shaha, A. Khaleque, and I. Hasan, “Low loss double cladding nested hollow core antiresonant fiber,” OSA
Continuum 3(9), 2512–2524 (2020).
36. Y. Wang, M. I. Hasan, M. R. A. Hassan, and W. Chang, “Effect of the second ring of antiresonant tubes in
negative-curvature fibers,” Opt. Express 28(2), 1168–1176 (2020).
37. G. P. Agrawal, “Nonlinear fiber optics,” in Nonlinear Science at the Dawn of the 21st Century (Springer, 2000), pp.
195–211.
38. M. H. Rahman, A. Khaleque, M. S. Hosen, K. S. R. Shaha, M. Mizan, and M. T. Rahman, “Dual-Function Plasmonic
Device on Photonic Crystal Fiber for Near to Mid-infrared Regions,” Opt. Mater. Express 13(9), 2526–2540 (2023).
39. M. S. Hosen, A. Khaleque, K. S. R. Shaha, L. N. Asha, A. S. Sultana, R. Nishad, and M. T. Rahman, “Highly
birefringent polarization maintaining low-loss single-mode hollow-core antiresonant fiber,” Opt. Continuum 1(10),
2167–2184 (2022).
40. L. Vincetti and V. Setti, “Extra loss due to fano resonances in inhibited coupling fibers based on a lattice of tubes,”
Opt. Express 20(13), 14350–14361 (2012).
41. X. Chen, X. Hu, L. Yang, J. Peng, H. Li, N. Dai, and J. Li, “Double negative curvature anti-resonance hollow core
fiber,” Opt. Express 27(14), 19548–19554 (2019).
42. B. Yang, X. Liu, W. Jia, S. Liu, K. Mei, T. Ge, D. Yang, Y. Liu, T. Lan, and Z. Wang, “Low Loss Hollow-Core
Connecting-Circle Negative-Curvature Fibres,” IEEE Photonics J. 13(1), 1–10 (2021).
43. L. N. Asha, M. S. Hosen, A. Khaleque, and K. S. R. Shaha, “Highly Birefringent Low Loss Hollow Core Nested
Antiresonant Fiber,” In 2022 4th International Conference on Electrical, Computer & Telecommunication Engineering
(ICECTE) (pp. 1–4). IEEE.
44. X. Huang, S. Yoo, and K. T. Yong, “Function of second cladding layer in hollow core tube lattice fibers,” Sci. Rep.
7(1), 1618 (2017).
45. X. Huang, W. Qi, D. Ho, K. T. Yong, F. Luan, and S. Yoo, “Hollow core anti-resonant fiber with split cladding,” Opt.
Express 24(7), 7670–7678 (2016).

You might also like