Vous êtes sur la page 1sur 15

Analytic model for light guidance in single-wall

hollow-core anti-resonant fibers


Wei Ding1,* and Yingying Wang2
1

Laboratory of Optical Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
2
Institute of Laser Engineering, Beijing University of Technology, Beijing 100124, China
*
wding@iphy.ac.cn

Abstract: We report an analytic model for quantitatively calculating the


transmission attenuation of single-wall hollow-core anti-resonant fibers. Our
calculations unveil the light leakage dependences on azimuthal angle,
polarization, and geometrical shape and have been examined in a variety of
fiber geometries. Based on our model, a simple and clear picture about light
guidance in hollow-core lattice fibers is presented. Formation of equiphase
surface at fibers outermost boundary and light emission ruled by Helmholtz
equation in transverse plane constitute the basis of this picture. Using this
picture, we explain how the geometrical shape of a single-wall hollow-core
fiber influences its transmission properties.
2014 Optical Society of America
OCIS codes: (060.2280) Fiber design and fabrication; (060.4005) Microstructured fibers;
(060.2400) Fiber properties.

References and links


1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.

F. Couny, F. Benabid, P. J. Roberts, P. S. Light, and M. G. Raymer, Generation and photonic guidance of
multi-octave optical-frequency combs, Science 318(5853), 11181121 (2007).
F. Luan, J. C. Knight, P. St. J. Russell, S. Campbell, D. Xiao, D. T. Reid, B. J. Mangan, D. P. Williams, and P. J.
Roberts, Femtosecond soliton pulse delivery at 800nm wavelength in hollow-core photonic bandgap fibers, Opt.
Express 12(5), 835840 (2004).
Y. Y. Wang, X. Peng, M. Alharbi, C. F. Dutin, T. D. Bradley, F. Grme, M. Mielke, T. Booth, and F. Benabid,
Design and fabrication of hollow-core photonic crystal fibers for high-power ultrashort pulse transportation and
pulse compression, Opt. Lett. 37(15), 31113113 (2012).
M. Tonouchi, Cutting-edge terahertz technology, Nat. Photonics 1(2), 97105 (2007).
J. C. Knight, J. Broeng, T. A. Birks, and P. S. J. Russell, Photonic band gap guidance in optical fibers, Science
282(5393), 14761478 (1998).
W. Belardi and J. C. Knight, Effect of core boundary curvature on the confinement losses of hollow antiresonant
fibers, Opt. Express 21(19), 2191221917 (2013).
F. Couny, F. Benabid, and P. S. Light, Large-pitch kagome-structured hollow-core photonic crystal fiber, Opt.
Lett. 31(24), 35743576 (2006).
N. M. Litchinitser, A. K. Abeeluck, C. Headley, and B. J. Eggleton, Antiresonant reflecting photonic crystal
optical waveguides, Opt. Lett. 27(18), 15921594 (2002).
P. J. Roberts, F. Couny, H. Sabert, B. J. Mangan, D. P. Williams, L. Farr, M. W. Mason, A. Tomlinson, T. A.
Birks, J. C. Knight, and P. S. J. Russell, Ultimate low loss of hollow-core photonic crystal fibres, Opt. Express
13(1), 236244 (2005).
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), 15371539 (1999).
T. D. Hedley, D. M. Bird, F. Benabid, J. C. Knight, and P. S. J. Russell, Modelling of a novel hollow-core
photonic crystal fibre, in Conference on Lasers and Electro-Optics/Quantum Electronics and Laser Science
Conference, Technical Digest (Optical Society of America, 2003), paper QTuL4.
G. J. Pearce, G. S. Wiederhecker, C. G. Poulton, S. Burger, and P. S. J. Russell, Models for guidance in
kagome-structured hollow-core photonic crystal fibres, Opt. Express 15(20), 1268012685 (2007).
S. Fvrier, B. Beaudou, and P. Viale, Understanding origin of loss in large pitch hollow-core photonic crystal
fibers and their design simplification, Opt. Express 18(5), 51425150 (2010).
M. Alharbi, T. Bradley, B. Debord, C. Fourcade-Dutin, D. Ghosh, L. Vincetti, F. Grme, and F. Benabid,
Hypocycloid-shaped hollow-core photonic crystal fiber Part II: Cladding effect on confinement and bend loss,
Opt. Express 21(23), 2860928616 (2013).
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), 669671 (2011).

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27242

16. A. D. Pryamikov, A. S. Biriukov, A. F. Kosolapov, V. G. Plotnichenko, S. L. Semjonov, and E. M. Dianov,


Demonstration of a waveguide regime for a silica hollow-core microstructured optical fiber with a negative
curvature of the core boundary in the spectral region > 3.5 m, Opt. Express 19(2), 14411448 (2011).
17. B. Debord, M. Alharbi, T. Bradley, C. Fourcade-Dutin, Y. Y. Wang, L. Vincetti, F. Grme, and F. Benabid,
Hypocycloid-shaped hollow-core photonic crystal fiber Part I: Arc curvature effect on confinement loss, Opt.
Express 21(23), 2859728608 (2013).
18. J. Hu and C. R. Menyuk, Understanding leaky modes: slab waveguide revisited, Adv. Opt. Photon. 1(1), 58106
(2009).
19. M. A. Duguay, Y. Kokubun, T. L. Koch, and L. Pfeiffer, Antiresonant reflecting optical waveguides in SiO2-Si
multilayer structures, Appl. Phys. Lett. 49(1), 1315 (1986).
20. D. Marcuse, Theory of Dielectric Optical Waveguides, 2nd Edition, (Academic, 1991).
21. A. W. Snyder and J. D. Love, Optical Waveguide Theory, (Chapman and Hall, 1983).
22. P. Yeh, A. Yariv, and E. Marom, Theory of Bragg fiber, J. Opt. Soc. Am. 68(9), 11961201 (1978).
23. S. Selleri, L. Vincetti, A. Cucinotta, and M. Zoboli, Complex FEM modal solver of optical waveguides with PML
boundary conditions, Opt. Quantum Electron. 33(4/5), 359371 (2001).
24. C. A. Balanis, Antenna Theory: Analysis and Design, 3rd Edition, (John Wiley & Sons, 2005).
25. 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),
95149519 (2013).
26. M. Born and E. Wolf, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction
of Light, 6th Edition, (Cambridge University Press, 1999).

1. Introduction
Hollow core micro-structured optical fibers (HC-MOFs), consisting of air core surrounded by
arrangement of micron-scaled silica webs, have recently been proposed as unique hosts for
both photons and vaporous gases. Numerous applications in fiber optics and light-matter
interactions, such as optical frequency comb generation [1], optical soliton pulse shaping [2],
high-power laser delivery [3], and Terahertz wave transmission [4], to mention just a few,
severely rely on efficient light guidance in such fibers whose most important parameters are
transmission attenuation and bandwidth. According to different waveguiding mechanisms,
HC-MOFs can be categorized to two classes with the first one, i.e. the hollow-core photonic
bandgap fiber (HC-PBGF) [5], being result of photonic bandgap (PBG) effects in fibers
cladding area and the second one, i.e. hollow-core anti-resonant fiber (HC-ARF) [6], for
example Kagome-lattice photonic crystal fiber [7], being result of anti-resonant light
reflections occurring at core-cladding interface [8]. Opening PBG leads to annihilation of
propagating optical states in fibers cladding, which enables light confinement inside the air
core. In such kind of fibers, loss figure as low as 1.2dB/km has been demonstrated [9], and the
PBG-induced light confinement becomes stronger as the number of cladding layer increases
[10]. In contrast, HC-ARFs only have spectral regions with low density of optical states (DOS)
[11]. Adding cladding layer has minor and vague effects on transmission attenuation [1214].
However, in aspect of transmission bandwidth, the HC-ARF substantially surpasses the
HC-PBGF. At the expense of attenuation coefficient (two orders of magnitude worse), the
HC-ARF can provide much broader transmission band (one order of magnitude wider) than the
HC-PBGF. A widely accepted explanation to this tradeoff is that opening a full twodimensional photonic bandgap in fibers cladding area benefits light confinement in core but
imposes additional restrictions on the operating wavelength [10]. Obtaining advantages in both
aspects seems difficult.
However, this knowledge about HC-MOF is now gradually changing with the appearance
of low-loss HC-ARFs, especially when several groups reported that a hypocycloid shape [15],
or negative curvature [16], core-surround can efficiently lower the attenuation coefficient while
maintaining the transmission bandwidth. These findings revive the investigations on the
guidance mechanism of HC-ARF. But, previous efforts, such as low DOS in fibers cladding
area [11], radial light confinement induced by concentric glass rings [12], and spatial power
overlap between core mode and cladding glass [17], only qualitatively explain part of the
transmission attenuation properties. A quantitative calculation of the influence of geometrical
shape to the attenuation spectrum is urgently needed. In this paper, we present an analytic

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27243

model for quantitatively calculating attenuation spectra of single-wall HC-ARFs. We relate the
attenuation coefficient to the integral of leaked energies in all the transverse directions. Our
analytic model is inspired by the leaky mode solution in one-dimensional (1D) slab waveguide
[1820]. By using a proper geometry transformation, our analytical treatments can be extended
from 1D slab geometry to 2D circular ring fiber and then to 2D polygon shape fibers with
gradually reduced structural symmetry. In principle, our analytic model can be employed in
arbitrary single-wall HC-ARFs.
This paper is organized as follows. In section 2, a wave equation and its solution of leaky
mode in an M-type 1D slab waveguide are presented. The relationship between the attenuation
coefficient and the field amplitude at the outermost boundary is derived under a picture of
oblique plane wave radiation. In section 3, the equivalence of 2D circular ring fiber and 1D slab
waveguide is discussed. A geometry transformation and an analytic calculation approach are
presented. In section 4, our model is tested in different shapes of single-wall HC-ARFs, i.e. the
regular triangle, square, hexagon, octagon, and hypocycloid. Electric fields and attenuation
spectra are calculated and are compared with numerically simulated results. The light leakage
dependences on azimuthal angle, polarization, and geometrical shape are analyzed. In the last
section, fundamental principles and further developments of our model are summarized before
conclusions are drawn.
2. Leaky mode in M-type slab waveguide
Our studies begin with calculating propagation constant of leaky mode in an M-type slab
waveguide [18]. We limit our investigations in fundamental core modes, whose field profiles in
the core area are peaked at the central axis.
As shown in Fig. 1(a), the core (with the thickness of 2ain) and the surrounding areas of the
slab waveguide are filled with air (n1 = 1), while the cladding consists of single layer of silica
with the refractive index n2 = 1.45 and the thickness t = 0.67 m. From the Helmholtz equation,
2 E ( z, x) 2 E ( z, x)
+
+ k0 2 n 2 ( x) E ( z , x) = 0 , the electric field distributions in this slab can be
z 2
x 2
written as [21],
cos(k x1 x ),
(Core)
(s, p)
(s, p)
),
(Cladding ) (1)
s / p-Pol. : E y , x ( z , x) = exp(i z ) Acl cos(k x 2 x +
A( s , p ) exp[ik ( x a t )], (Surrounding )
x1
in
sur

where z (x) denotes the direction along (transverse to) the waveguide, the s and p polarizations
are depicted in Fig. 1(a). The propagation constant in the z direction, , equals to k0 neff with k0
the propagation constant in vacuum and neff the modal index. The transverse wave-vector
number kxj is defined as k0 n j 2 neff 2 . The complex field amplitude A, the phase , and the
eigenvalue equation relevant to can be derived from boundary continuity conditions [21]. In
the rest of this paper, the eigenvalue equation of , which can be found in many textbooks, will
be frequently mentioned and belongs to one part of our analytic calculation approach.
As to the leaky mode character of Eq. (1), we stress that the field in the surrounding area
only contains outward-propagating wave component, whereas the fields in the core and the
cladding areas are standing waves. Additionally, the transverse wave-vector number in the
surrounding area, kx1, has a non-zero imaginary part, leading to an exponential growth of field
amplitude as x . Strictly speaking, for a dielectric waveguide, a finite number of discrete
guided modes and a continuum of radiation modes constitute the complete set of orthogonal
basis. Leaky modes, which are mathematical solutions under the assumption that no
inward-propagating wave component exists in the outermost layer, are not members of this
orthogonal basis. However, detailed analyses have verified the equivalence between this

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27244

mathematical solution of leaky mode and the realistic physical process of energy diffusion in
the radiation mode continuum [18].

Fig. 1. (a) Schematic diagram of an M-type slab waveguide. (b) Field amplitudes (logarithmic
scale) and phases of the dominant electric field components for the s- and p-polarization waves
as a function of the x coordinate. Parameters used in calculations include ain = 2 m, t = 0.67 m,
n2 = 1.45, and 0 = 0.938 m.

Figure 1(b) plots the field amplitudes and phases of the leaky modes calculated in a slab
waveguide. For the p-polarization, only the transverse electric field, Ex, is presented. The
longitudinal field component, Ez, is weak and ignored. Note that, in these calculations, the
modal index, neff, is regarded as a complex number with its imaginary part proportional to the
attenuation coefficient, [dB m] = 8.69k0 Im(neff ) . Moreover, the complex nature of neff can

also be seen in the field amplitude profiles, which grow exponentially in the direction
transverse to the propagation as illustrated by the blue dashed line in Fig. 1(b).
The field distribution of the leaky mode, i.e. Equation (1), can also be understood with the
energy conservation. In a lossless dielectric waveguide, energy decrease in the longitudinal
direction, due to attenuation, should be equal to energy leakage in the transverse directions. As
illustrated in the insert of Fig. 2, two outward-inclined plane waves lie in the surrounding area
and satisfy the phase-matching condition in the longitudinal direction. The oblique angle of
these plane waves can be approximately estimated, and a formula relating the complex modal
index and the field amplitude at the outermost boundary can be derived,
Im(neff ) =

n1 E ( x = ain + t +)

n12 [Re(neff )]2

2k0 Re(neff )

ain + t

n1 E ( x = ain + t + )

E ( x) dx

(2)

n12 [Re(neff )]2

E ( x = 0) k0 ain Re(neff )

Here, only the transverse field components, i.e. the major field components, are taken into
account, E(x + ) is defined to be lim E ( x + x) , and the resonant condition of the fundamental
x 0 +

core mode, i.e. Re(k x1 ) 2ain , is also used. In order to check accuracy of Eq. (2), we read
out the field amplitudes at the outermost boundary from Fig. 1(b) (marked by the blue circles),
estimate Re(neff ) 1 (0 4ain ) 2 = 0.9931 , and then derive imaginary parts of the modal
indices for the s- and p-polarization waves to be 1.08 104 and 4.77 104 respectively, which
agree well with the precisely calculated results based on the eigenvalue equation [Im(neff) =
1.08 104 and 4.73 104 respectively]. More importantly, by reversely using Eq. (2), we can
also estimate the field amplitude at outermost boundary of the waveguide from the complex

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27245

modal index. Below, this method will be frequently used. In contrast, obtaining full field
distribution by solving Eq. (1) requires explicit geometrical information besides the knowledge
of modal index.
To exhibit spectral properties of this M-type slab waveguide, Fig. 2 plots the attenuation
spectra of the fundamental core modes for both the s- and p-polarizations. It is seen that the two
high-loss wavelength regions, determined by the anti-crossings between the fundamental core
modes and different order glass confined modes, appear at 0 0.7 and 1.4 m. The m-th order
resonant wavelength, 0 2t n2 2 n12 m [8,19], has little relationship with the polarization.
However, inside each transmission band, the p-polarization wave exhibits much worse light
confinement than the s-polarization, implying that, in the case of a hybrid polarization wave,
the p-polarization component will play the primary role in light leakage. This anti-resonant
reflecting optical waveguide (ARROW) mechanism has been well studied in the context of
planar waveguides [8] and fiber geometries [19]. Figure 2 also plots the phases of the electric
fields at the outermost boundary. Surprisingly, these phases seem only determined by the
anti-resonant order of the transmission band, irrelevant to polarization and geometrical
dimension (data not shown). This phase locking effect is very important and will be used in the
following treatment.

Fig. 2. Attenuation coefficient and phase of the field at the outermost boundary (the pink line in
the insert) for the fundamental core mode in an M-type slab waveguide. The vertical gray lines
represent the resonant wavelengths of the glass modes. ain = 2 m, n2 = 1.45, and t = 0.67 m.

Briefly speaking, all the information about the electric field at the outermost boundary of an
M-type slab waveguide can be quickly obtained once the complex modal index of the leaky
mode is known. The field amplitude can be estimated from Eq. (2), and its phase is fixed to a
constant value only relevant to the anti-resonant order of the transmission band.
3. Analogy between single-wall circular ring fiber and M-type slab waveguide

Comparing with slab waveguide, a circular ring fiber loses the continuous translation symmetry
in the y direction. The continuous rotational symmetry of the geometry is also impaired by the
linear polarization of the fundamental core mode. To solve this problem, a reasonable geometry
transformation from 2D geometry to 1D slab may be helpful.
Figure 3(a) depicts cross section of a circular ring fiber (with the inner radius ain and the
glass thickness t) in a cylindrical coordinate. The radial light leakage through the glass wall at
any azimuthal angle [denoted by the green arrow in Fig. 3(a)] is supposed to be equivalent to
that in a slab waveguide having the inner radius r and the glass thickness t. We assume
r '( ) = ain 2 and rotate the glass segment to parallel to the y axis. The first procedure is

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27246

based on the consideration that, as a result of geometry transformation, the transverse light
confinement of the core mode will be relaxed from the x-y plane (2D) to the x direction (1D).
The second procedure will be accompanied with a rotation of polarization direction having the
same degree. The whole geometry transformation is depicted in the insert of Fig. 3(a).
After this geometry transformation, the complex modal indices of the leaky modes,
(s, p)
neff
( ) , and the field amplitudes at fibers outermost boundary, E ( s , p ) ( ) , can be quickly
solved in the equivalent slab structure, whose inner radius and glass thickness are r '( ) and t
respectively.
12

E 2 k a Im[n ( s , p ) ( )] Re[n ( s , p ) ( )]

eff
eff
E ( s , p ) (at fiber's outer boundary) = 0 0 in
(3)
(s, p)
2
2
1 Re[neff
( )]

Here, stands for the azimuthal angle, and the superscripts represent the s- and p-polarizations.
Equation (3) is equivalent to Eq. (2). The additional factor 21 2 is because of the geometry
transformation from 2D to 1D.

Fig. 3. (a) Equivalence of a circular ring fiber and a series of slab structures. (b) Attenuation
coefficients and propagation constants (real parts of the effective modal indices) of a single-wall
circular ring fiber as a function of the wavelength. The precisely calculated results (hollow
squares) are from a transfer matrix approach [22]. The results of our model (red curves) are from
Eqs. (4) and (5). ain = 9.76 m, n2 = 1.45, and t = 0.67 m.

Taking into account the proportions of s- and p-polarization components in different


angular segments (cos2 and sin2, respectively), the overall modal index of the whole fiber can
be expressed as an arithmetical average,
1 2
{Re[neff ( s ) ( )] cos 2 + Re[neff ( p ) ( )] sin 2 }d
(4)
2 0
In order to obtain the overall attenuation coefficient, we exploit the phase-matching
condition in the longitudinal direction and the leaky mode characteristics in the surrounding
area. According to the phase-matching condition, the 2D Helmholtz equation in the transverse
plane can be written as T 2 E ( x, y ) + k0 2 [1 Re(neff ) 2 ]E ( x, y ) = 0 . And, according to the leaky
Re(neff ) =

mode characteristics, the far-field radiation at any azimuthal angle, [the pink arrow in Fig.
3(a)], can be obtained by integrating the electric field along fibers outermost boundary. The
amplitudes of these electric fields are products of Eq. (3) and cos or sin, depending on
different polarizations [depicted in Fig. 3(a)], and the phases of these fields are fixed (Fig. 2).
Therefore, the overall attenuation coefficient of the circular ring fiber can be expressed as,

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27247

[dB m] = 8.69k0 Im( neff ) = 8.69k0 eRad 1 Re( neff ) 2 Re( neff )
eRad

1
8 ain 2

[ ex ( ) + ey ( ) ]d

(5a)
(5b)

E ( p) E (s)
2
cos( ) + 1
ex ( ) =
exp[ikT (cos x + sin y )] dl
cos sin

k0 ain
E0
2

E ( s ) cos 2 + E ( p ) sin 2 cos( ) + 1

2
=

exp[ikT (cos x + sin y )] dl


e
(
)

y
k0 ain
E0
2

(5c)

Equation (5a) represents the energy conservation in a leaky mode, Eq. (5b) averages the
contributions from all the radiation directions and from both the s/p-polarizations, and the
integral of Eq. (5c) takes into account the angle between x/y-axis and s/p-polarization (), the
inclination angle of each glass segment relative to the radiation direction ( - ), and the phase
delay in the radiation direction (). The transverse wave-vector, Re(kT ) k0 1 Re(neff ) 2 ,
utilizes the result of Eq. (4). A closed loop integral in Eq. (5c) is implemented along fibers
outermost boundary and can be understood by Greens theorem. Detailed derivation of Eq. (5)
is given in the Appendix.
Our model is then compared with precisely calculated result of single-wall circular ring
fiber. The precise calculation is carried out by using Bessel functions and standard transfer
matrix technique [22]. In the calculation, the azimuthal number of the Bessel functions is set to
be 1, corresponding to the HE11-like leaky core mode, and the surrounding area only contains
the outward-propagating Hankel function. Figure 3(b) shows that, with respect to the
attenuation coefficient, a very good agreement between our model and precise calculation is
achieved in a broad wavelength range. This result verifies that the geometry transformation
proposed in Fig. 3(a) is reasonable. Our analytic model not only demonstrates the capability of
quantitatively calculating attenuation coefficient of single-wall circular ring fiber but also
solves the problem of the symmetry decrease from 1D slab to 2D fiber, which is very useful in
the following treatment. Note that the geometrical sizes and spectral range used in Fig. 3 are
exactly the same with those in [12], and our results coincide with theirs as well.
4. Light leakage dependences on azimuthal angle, polarization, and geometrical shape

Our analytic model will be examined in more complicated geometries, i.e. regular polygon and
hypocycloid shape single-wall hollow-core fibers. With the loss of cylindrical symmetry of
circular ring fiber, precise calculation of attenuation coefficient becomes impossible. We,
therefore, use a commercial finite-element mode solver (Comsol Multiphysics) to provide
comparison for our model. The precision of our numerical simulation is kept better than 0.5%
by choosing appropriate mesh size and perfectly matched layer (PML) configuration [23].
Figure 4(a) depicts cross section of a regular triangle single-wall fiber, whose inscribed
radius is ain. In a cylindrical coordinate, the inner boundary of the fiber is expressed as r ( ) . In
order to construct an equivalent slab waveguide as Fig. 3(a) does, a hypothetical inner radius
r '( ) = r ( ) 2 , a glass thickness t, and a polarization rotation 2 (relative to the y
axis) are employed as indicated in Fig. 4(a). Here, is a function of the azimuthal angle and
stands for the tangential direction of glass segment. After geometry transformation, a series of
equivalent slab waveguides are obtained at different azimuthal angles. Analytical calculation

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27248

can then be applied to obtain the modal index neff ( s , p ) ( ) and to estimate the field amplitude at
fibers outermost boundary based on Eqs. (1)-(3).
12

(s, p)

E 2 k a Im[n ( s , p ) ( )] Re[n ( s , p ) ( )]

eff
eff
(at fiber's outer boundary) = 0 0 in
( )
(s, p)
2
2
1 Re[neff
( )]

(6)

The factor ( ) = sin( ) ain r ( ) is introduced due to the energy conservation.

Fig. 4. (a) Geometry transformation from a triangular fiber to a series of slab structures. The
green arrows represent the equivalent light leaking processes occurring in 2D and 1D. The pink
arrow denotes the radiation direction of . (b) Analytically modelled (red curves) and
numerically simulated (hollow squares) attenuation coefficients and propagation constants of
the fundamental core mode in this triangle fiber as a function of the working wavelength
(logarithmic scale). Parameters used in calculations: ain = 9.76 m, n2 = 1.45, t = 0.67 m.

The overall modal index, Re(neff ) , is equal to the arithmetic mean of Re[neff ( s , p ) ( )] over
different angular segments. The proportions of the s- and p-polarization components need to be
taken into account.
1 2
{Re[neff ( s ) ( )] sin 2 + Re[neff ( p ) ( )] cos 2 }d
(7)
2 0
Based on this overall modal index, the transverse wave-vector number in the surrounding
Re(neff ) =

area, Re(kT ) k0 1 Re(neff ) 2 , is derived and used in the 2D Helmholtz equation in the x-y
plane. Integrating electric field along fibers outermost boundary yields the emission energy at
any transverse angle [the pink arrow in Fig. 4(a)]. Contributions from x- and y-polarizations
should be both taken into account. Then, the overall attenuation coefficient of the fiber can be
derived by integrating the radiations over all the transverse directions,
2

[dB m] = 8.69k0 eRad 1 Re( neff ) 2 Re(neff )


eRad

1
8 ain 2

[ ex ( ) + ey ( ) ]d

(8a)
(8b)

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27249


E ( p) E (s)
2
sin( ) + 1 ikT (cos x + sin y )
ex ( ) =
e
dL
cos sin

k0 ain
E0
2

(8c)

E ( s ) sin 2 + E ( p ) cos 2 sin( ) + 1 ik (cos x + sin y )

e T
dL
e y ( ) =
k0 ain
E0
2

Note that the tangential angle of glass segment, , has been used. In Eq. (8c), the integral length
element is derived from dL = dl = dl/sin(-) [see the insert in Fig. 4(a)], whereas, in Eq. (5c),
the length element, dl, is equal to (ain + t).
Figure 4(b) compares our model (red curves) with numerically simulated results (hollow
squares) for a triangle single-wall fiber. In terms of transmission attenuation, two calculations
agree very well across more than two spectrum octaves (from 0 = 0.3 m to 1.7 m), which
explicitly verifies the quantitative calculation capability of our model.
To inspect our model more carefully, Fig. 5 plots the electric field amplitudes and phases at
fibers outermost boundary [the red dashed lines in Fig. 5(a)]. The field amplitudes along a
far-distance circle R (the gray dashed lines) are also plotted. Under two orthogonal
polarizations [denoted by the double arrows in Fig. 5(a)], on the aspect of both near fields [Fig.
5(b)] and far fields [Fig. 5(c)], our model (the red curves) agrees well with the numerical
simulation (the black curves). The working wavelength is selected to be 680 nm [marked by the
blue dot in Fig. 4(b)], which is very close to the edge of one transmission band. Under such a
rigorous condition, the quantitative calculation capability of our model is still evident.

Fig. 5. Analytically modelled (red curves) and numerically simulated (black curves) field
amplitudes and phases at (b) the outermost boundary and (c) a far-distance circle R. The
schematic diagrams (a) depict a triangle single-wall fiber (orange), its outermost boundary (red),
a far-distance circle R (gray), and the light polarizations (double arrows). Under each
polarization, only the dominant electric field component, either Ex or Ey, is plotted. All the field
amplitudes have been normalized for the convenience of comparison. Parameters used in
calculations: ain = 9.76 m, n2 = 1.45, t = 0.67 m, R = 60 m and 0 = 0.68 m.

With regard to the electric field at fibers outermost boundary, Fig. 5(b) shows that except
around corner apexes the phases of the electric fields are fixed to a constant. Additionally, in
comparison with the s-polarization, the p-polarization component gives rise to stronger electric
fields outside the glass wall. These features are coincident with those shown in 1D slab
geometry [Fig. 1(b)]. Figure 5(b) also shows that our model precisely predicts the ratio of the
field amplitudes in the central regions of three edges of the triangle. However, when the

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27250

observation point approaches the corner apexes, the field amplitude predicted by our model is
over-estimated. This may be because our approximate model ignores the influences from
adjacent glass segments [see the geometry transformation in Fig. 4(a)]. Actually, at the corner
apexes, our geometry transformation procedure stops working.
With regard to the far-field properties, Fig. 5(c) plots the field amplitudes at a circle R,
whose radius of 60 m is comparable with Fraunhofer far-field criterion ( 2 D 2 T 90 m ,
where D is the largest dimension of the structure and T

1 Re(neff ) 2 is the effective

transverse wavelength [24]). In Fig. 5(c), both our model and numerical simulation ensure that
the light leakage in the transverse plane is far from isotropic. The radiations pointing at the
corner apexes are suppressed, whereas those in the directions vertical to the triangle edges are
enhanced. Most radiation energies are emitted out through the triangle edges nearly vertical to
the polarization direction, where there exist more p-polarization wave components. With this
understanding, it may be possible to design new HC-ARF structures having lower attenuation
coefficients. More light confinement structures, for example PBG cladding, could be mainly
deployed in the stronger light leaking directions.
We emphasize that the above azimuthal angle dependence of light leakage [Fig. 5(c)] will
not cause birefringence for a regular triangle fiber. In calculating the overall modal index and
overall attenuation coefficient, our model averages contributions from all the angular segments
and all the transverse directions, which eliminates the azimuthal angle dependence of these two
parameters.
Before stepping forward, we discuss again the problem of corner apex. In Fig. 6, a
single-wall square fiber is compared with its variant, whose corners have been rounded with 4
circles having radius rc = 2.5 m. Simulated attenuation spectra of these two fibers exhibit one
distinct difference. The square fiber shows many spiky features (the black curve), whereas the
variant square fiber does not (the green curve). We believe new resonances, rather than those
relevant to the uniformly thick glass wall, bring forth these spikes. We attribute them to either
back reflection or non-uniformity of the glass thickness occurring at the corner apexes (see the
schematic illustrations in Fig. 6). With regard to the second conjecture, we notice Kolyadin et
al. have found that touching capillaries in a negative curvature hollow core fiber dramatically
degrade transmission performance [25]. In their structure, the back reflection from the touching
point can be excluded, so that, the non-uniformity of the glass thickness at the touching point
should play an important role. In our case of Fig. 6, the four corner apexes bring about the
non-uniformity of glass thickness. Replacing them with rounded curvatures, the variant square
fiber shows a much smoother attenuation spectrum, agreeing well with the prediction of our
analytic model (the red curve).
Additionally, we cannot find spiky structures in the attenuation spectrum of the triangle
fiber [Fig. 4(b)]. One possible explanation is the corner apexes in the triangle fiber lie far away
from fibers central axis and their influences to the fundamental core mode are weak.

Fig. 6. Simulated attenuation spectra of a single-wall square fiber (black squares) and its variant
(green squares), comparing with the calculated result based on our analytic model (red curve).
ain = 9.76 m, n2 = 1.45, t = 0.67 m, and rc = 2.5 m.

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27251

Next, we consider another square fiber geometry in order to illuminate the influence of
glass thickness non-uniformity. In Fig. 7, the inscribed radius of the fiber is ain, and the glass
thickness in the vertical direction (t2 = 0.6 m) is slightly thinner than that in the horizontal
direction (t1 = 0.67 m). With the variation of polarization, both the simulated and the modelled
attenuation spectra exhibit notable changes. First, two sets of resonant conditions
( 0 (1,2) = 2t1,2 n2 2 n12 m ) commonly determine the edges of the transmission bands for both
polarizations. Second, the shapes of the attenuation spectra are dramatically changed under two
polarizations. Our analytic model (the red and the green curves) agrees very well with the
numerical simulation (the black and the blue ones), implying that our model has grasped the
physical essence of the light leaking process occuring in such a complex 2D structure. For
reference, we also plot attenuation spectra of two normal square fibers, which have uniform
glass thicknesses of t = 0.6 m and 0.67 m in four sides respectively (denoted by the dashed
cyan and pink lines respectively). With uniform glass thickness, the polarization dependence
vanishes, but different ts lead to a shift of transmission band. It is seen that at specific
polarizations and wavelengths (e.g. the reddish and the greenish shaded areas) the modified
square fiber, which has non-uniform glass thickness, exhibits lower attenuation coefficient than
the normal square fiber, which has uniform glass thickness. This result is against the widely
accepted viewpoint that a uniform glass thickness across whole lattice fiber leads to better light
guidance [12]. The principle underneath this phenomenon may be that most light leakages are
caused by the p-polarization components, and the anti-resonance condition of the
p-polarization wave 0 ( anti ) = 4t n2 2 n12 (2m + 1) is only relevant to the glass thickness
vertical to the polarization direction.

Fig. 7. Simulated (black and blue curves) and modelled (red and green curves) attenuation
spectra of a modified square fiber under vertical (black and red curves) and horizontal (blue and
green curves) polarizations. The cyan and the pink dashed curves represent the two normal
square fibers having uniform thicknesses of 0.6 and 0.67 m respectively. In the vertical
polarization, the modified square fiber shows lower loss than the normal square fiber having t =
0.67 m in the wavelength range marked by the reddish shaded area. In the horizontal
polarization, the modified square fiber shows lower loss than the normal square fiber having t =
0.6 m in the greenish shaded area. ain = 9.76 m, and n2 = 1.45.

To elucidate the influence of geometrical shape on fibers transmission properties, Fig. 8


plots attenuation spectra of various single-wall regular polygon fibers with regular triangle,
square, hexagon, and octagon. Both numerical simulations and our analytic model are carried
out with the inscribed radii ain = 9.67 m. Apart from some spiky features, our model agrees
well with numerical simulation, especially in terms of their variation tendency. As the number
of edges of the polygon increases, the polygon fiber becomes more and more alike a circular
ring and results in a worse and worse light confinement.

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27252

Fig. 8. Simulated and modelled attenuation spectra of a regular triangle (black), square (red),
hexagon (green), and octagon (blue) single-wall fiber. The gray dashed curves in both graphs
represent a circular ring fiber. ain = 9.76 m, n2 = 1.45, and t = 0.67 m.

Since a circular ring fiber is not favored for light guidance, we study another HC-ARF
structure, which has a hypocycloid shape core-surround [15,16]. Our analytic model is able to
elucidate why light guidance is improved in such a fiber. Figure 9 compares two single-wall
fibers with one having a square shape and the other having a hypocycloidal shape. Both
numerical simulation and analytical model corroborate that the latter one has lower loss. As
pointed out above, the electric fields at fibers outermost boundaries form equiphase surfaces.
A hypocycloidal, or concave, equiphase surface seems good at suppressing the overall light
leakage. In Fig. 9, although the decreasing extents of the simulated and the modelled
attenuation spectra show some discrepancy, the light leakage suppression effect caused by the
hypocycloid core-surround is clearly demonstrated. We believe the small disagreement in Fig.
9 can be overcome with further development of our model.

Fig. 9. Simulated and modelled attenuation spectra of a square shape fiber (black curves) and a
hypocycloidal square fiber (red curves). The hypocycloid shape fiber is schematically depicted
in the insert. ain = 9.76 m, n2 = 1.45, and t = 0.67 m.

5. Discussions and conclusions

Shortly speaking, the analytic and quantitative characteristics of our calculation approach have
been corroborated. Throughout this paper, except those spiky features, the discrepancy between
our model and numerical simulation is roughly less than 0.5 dB, or 10%, in terms of
transmission attenuation. The logarithmic scales used in Figs. (3,4,6-9) clearly exhibit this.
More importantly, thanks to the following three techniques, our model correctly predicts the
#220643 - $15.00 USD
Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27253

change of the shape of the attenuation spectrum and the variation tendency of the loss figure as
fibers geometrical shape changes.
The phase of electric field at the outermost boundary is only determined by the order of
the anti-resonant transmission band, irrespective of polarization, geometrical
dimension and working wavelength. This property can be approximately derived in
1D slab geometry [from Eq. (1)]. We apply it in 2D fiber structures. Fortunately, our
attempt has been verified by simulation [Fig. 5(b)] and leads to a very simple picture
of equiphase surface.
The relationship between the longitudinal and the transverse wave-vector numbers,
k0 2 neff 2 + kT 2 = k0 2 n 2 , combining with the leaky mode character, provide us

conveniences to deal with the light leaking problem. First, the real part of the
longitudinal wave-vector number, k0 Re(neff ) , represents the propagation constant of
the fiber, and the real part of the transverse wave-vector number, Re(kT ) , defines a
2D Helmholtz wave equation in the x-y plane, T 2 E + Re(kT ) 2 E = 0 . Second, the
imaginary part of the longitudinal wave-vector number, k0 Im(neff ) , is related to the
attenuation coefficient, , and has relationship with the field amplitude at fibers
outermost boundary. Third, once knowing the fields at the outermost boundary, rather
than the full field distribution in the cross section, the leaked energies in all the
transverse directions can be derived from the 2D Helmholtz equation. The shape of the
fibers outermost boundary influences the overall light leakages and determines the
attenuation coefficient.
In order to simplify mathematical treatments, we propose a geometry transformation
from 2D to 1D. By splitting the cross section of a single-wall fiber to different angular
segments and converting them to a series of slab structures, the modal indices and the
fields at the outermost boundary can be quickly and analytically solved. For
fundamental core mode, the transverse field distribution inside the air core can be
approximated to be a linearly polarized Gaussian beam. Each angular segment and the
corresponding slab waveguide, therefore, should contribute equally to the overall
modal index and light leakage. Besides, in order to further simplify this problem, our
model deals with s- and p-polarization wave components separately, and their
proportions in each angular segment are determined in the geometry transformation
procedure.
We have to admit our model still has many drawbacks. The first one is the geometry
transformation procedure. As depicted in Fig. 4(a), each angular segment split in the cross
section of the fiber is treated independently. The influences from adjacent segments have not
been incorporated, which may explain why the field amplitude in Fig. 5(b) is over-estimated as
the observation point approaches the corner apexes. The second drawback is that, at the current
stage, our model is only suitable for single-wall hollow core fibers. Although many people
believe the core-surround layer plays the most important role in determining the transmission
attenuation, it would be better to include the influences from other cladding structures. This
work is ongoing now and will be reported later. The third drawback is that our calculations have
not considered influences of corner apexes. Although the field amplitude over there is very
weak according to simulation [see Fig. 5(b)], appearance of new resonances will bring about
spiky features in many spectral regions.
In conclusion, our analytic model exhibits the capability of quantitatively calculating
transmission attenuation of single-wall hollow core fibers. This model elucidates the light
leakage dependences on azimuthal angle, polarization, and geometrical shape and has been
examined in a variety of fiber geometries. A simple and clear physical picture about the light

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27254

leaking process has been presented. First, an equiphase surface is formed at fibers outermost
boundary. Then, the light energy is transversely emitted out ruled by a 2D Helmholtz wave
equation. Our model not only grasps the physical essences but also simplifies mathematical
treatments by introducing many approximations.
Since fiber geometry can influence the light leakage and the attenuation coefficient, many
interesting low-loss HC-ARF designs are in prospect.
Appendix

Below, we present derivation of Eq. (5). Based on the 2D Helmholtz equation in the x-y plane,
T 2 Ex , y (r ) + kT 2 Ex , y (r ) = 0 , and Greens theorem, the x/y-component of the electric field in the
surrounding area can be expressed as [26],
Ex , y (r ) = Ex , y ( G n ) G ( Ex , y n ) dl

(9)

A closed loop integral is implemented along fibers outermost boundary, which has a unit
normal vector, n, as shown in Fig. 10. The 2D Greens function, G = i H 0(1) (kT s ) 4 , satisfies
T 2 G + kT 2 G = (r r ') , where () represents the Dirac delta function and the Hankel
function of the first kind of order zero has asymptotic form of

kT s

exp[i (kT s )] as s
4

approaches infinity.

Fig. 10. Energy flow in the transverse plane. Field integral is implemented along the outermost
boundary (the pink line) of the circular ring fiber (the orange ring). The polarization directions of
E(s,p) and Ex,y, and the two azimuthal angles (, ) are depicted.

Since fibers outermost boundary constitutes an equiphase surface, we hypothesize each


segment on this closed loop produces a plane wavelet pointing at its normal direction n, i.e.
Ex , y
G
1
cos(n , s ) (ikT )G cos(n , s ) ikT G , as s approaches
ikT Ex , y . Meanwhile,
n
2s
n
infinity. With r s, Eq. (9) can be approximately expressed as,
Ex , y (r ) e3 i 4

kT
cos(n , s ) + 1
Ex , y (r ')
exp(ikT s )dl

2 r
2

(10)

where [cos(n , s ) + 1] 2 is the Kirchhoffs inclination factor in the Huygens-Fresnel principle,


and the x/y-components of the electric field at fibers outermost boundary can be derived from
the s/p-components,
Ex (r ') = E ( p ) sin cos E ( s ) cos sin

E y (r ') = E ( s ) cos 2 + E ( p ) sin 2

(11)

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27255

Integrating all the energy flows in transverse outward directions yields


2
2
k 2
2
FT = kT z E dl = kT z T [ Ex ( ) + E y ( ) ]d
R
0
2
cos(n , s ) + 1

( p)
(s)
exp(ikT s )dl
Ex ( ) [ E E ]sin cos
2

E ( ) [ E ( s ) cos 2 + E ( p ) sin 2 ] cos(n , s ) + 1 exp(ik s )dl


T

y
2

(12)

where z is the differential element in the z direction, kT is the transverse wave-vector number,
and the integrating loop of a circular ring R (the dashed line in Fig. 10) lies in the far-field
region. Note that in Eq. (12) both the x- and y-components are taken into account because the
2D scalar Helmholtz equation allows both polarized waves to propagate toward the infinity.
The transverse-wave characteristic of a propagating electromagnetic wave is maintained in
three-dimensional space. However, in the case of glancing incidence, the transverse k-vector
and the transverse electric field vector can be both in the radial direction.
On the other hand, the exponential energy attenuation in the longitudinal direction can be
expressed as
2
2
2
= [2k Im(n )z ] k
FL = k z E ds E ds

eff
z Core E ( x, y ) dA
0
Core
Core

z
z +z
(13)
a2
2
[2k0 Im(neff )z ] k0 Re(neff ) E0 2
2

The field distribution in fibers core area is approximated to be a Gaussian function


E ( x, y ) = E0 exp ( r 2 2a 2 ) . Utilizing the energy conservation condition, Eq. (12) = Eq.

(13), and the definition of E(s,p) [Eq. (3)], we obtain the following expression, i.e. Equation (5),
Im(neff )

1 Re(neff ) 2
8 ain Re(neff
2

[ ex ( ) + ey ( ) ]d

cos(n , s ) + 1

( p)
(s)
exp(ikT s )dl
ex ( ) ( e e ) sin cos
2

e ( ) ( e( s ) cos 2 + e( p ) sin 2 ) cos(n , s ) + 1 exp(ik s )dl


T

y
2

(14)

12

(s, p)

Im[n ( s , p ) ( )] Re[n ( s , p ) ( )]

eff
eff

(s, p)
2
1 Re[neff
( )]

Here, the resonant condition of the fundamental core mode, kT ain 2 , and an approximate
relationship a ain

2 have been used.

Acknowledgments

This work was supported by the National Basic Research Foundation of China (No.
2011CB922002), the National Natural Science Foundation of China (No. 61275044,
61377098, and 11204366), the Key Programs of the Chinese Academy of Sciences (No.
YZ201346), the Start-up Funding of Institute of Physics (No. Y1K501DL11), and the Beijing
National Science Foundation (No. 4142006).

#220643 - $15.00 USD


Received 8 Aug 2014; revised 25 Sep 2014; accepted 15 Oct 2014; published 27 Oct 2014
(C) 2014 OSA
3 November 2014 | Vol. 22, No. 22 | DOI:10.1364/OE.22.027242 | OPTICS EXPRESS 27256

Vous aimerez peut-être aussi