You are on page 1of 85

“DESIGN AND FABRICATION

OF SPECIAL PHOTONIC
CRYSTAL FIBERS WITH HIGH
COUPLING OF THE
FUNDAMENTAL MODE TO
LEAKY MODES FOR
REFRACTIVE INDEX SENSING
DEVICES”

Trabajo de Grado
para optar al tı́tulo de Maestro en Ciencias Ópticas

Presenta: Johan Sebastian Buriticá Bolaños

Director de Tesis: Uladzimir Minkovich, Ph.D.

León, Guanajuato - México


Noviembre 2023
Acknowledgments

This Master’s Degree thesis was made under the guidance of Dr. Uladzimir Minkovich,
thanks to the Centro de Investigaciones en Óptica (CIO) which facilitated its facilities and
resources and to the Consejo Nacional de Humanidades Ciencia y Tecnologı́a (CONAHCYT)
who generously financed the student scholarship during these two years.

First of all, I would like to thank my entire family, from my grandparents to my siblings,
to the master’s teachers with whom I took courses during these two years of the Master’s
Degree in Optical Sciences, to my fellow graduate students and fellow nationals. for each
and every one of your contributions, both in my personal and professional training.

Thanks to the Technician of the optical fiber group at the Centro de Investigaciones en
Óptica, Myrian Cristina Jimenes Mares, for her efforts and collaboration with the splicing
and tapering of conventional SMF-28 fibers and special PCF fibers. Finally I personally want
to thank my advisor, Dr. Uladzimir Minkovich for all his patience, hard work , feedback and
guidance along these two years.
Abstract

This work explores the possibility of implementing a detection mechanism in optical fibers
called “Lossy Mode Resonance” (LMR), which has begun to grow in popularity because, com-
pared to other similar detection mechanisms in the area of optical fibers, has high flexibility
with respect to its implementation. By using photonic crystal fibers (PCF) instead of con-
ventional optical fibers, it is possible to obtain improved capacities, which can be controlled
by designing the microstructure of the photonic crystal fibers. For this work, 2 photonic
crystal fibers (F1 and F2) were manufactured that have 2 rings of air channels, which have a
size of davg = 3.31 µm . for F1 and davg = 3.93 µm F2 . Due to this configuration of the fiber
microstructure, the evanescent field of the fiber is strongly enhanced and consequently its
interaction with the medium external to the fiber makes it a good platform for sensing appli-
cations. To implement the Lossy Mode Resonance phenomenon in the photonic crystal fiber,
it was necessary to subject the fiber to different conditioning such as tapering and removing
part of the fiber coating to use another type of material (in this case polyethyleneimine PEI).
Objectives

General Objectives
• Carry out fabrication of a special photonic crystal fiber with high coupling from the
fundamental mode to Leaky Modes for sensing applications.

Specific Objectives
• To carry out the fabrication of silica composite preform for drawing special photonic
crystal fibers with high coupling of the fundamental mode to leaky modes for refractive
index sensing devices.

• Experimentally characterize the microstructure of the fabricated photonic crystal fibers


using an atomic force microscope and measure attenuation of said fibers.

• Perform numerical simulations of the characteristics of the fiber; chromatic dispersion,


normalized frequency, fundamental mode radius, among others from the structural
parameters of the fiber.
Introduction

Over time, humans have come up with ingenious ways to harness light to convey messages
and make meaningful connections. This fascinating practice has evolved from ancient smoke
and fire signals, reflective mirrors, and signal lamps, to the advanced optical communication
technologies that today enable data transfer at blazing speeds.
Approximately four years after the invention of the telephone, Graham Bell introduced a
device known as the “Photophone” [1]. This device used modulated sunlight through a
diaphragm to transmit voice over distances of up to 200 meters [2]. However, due to the
limitation of suitable light sources and the vulnerability of optical transmission to weather
disturbances, this invention did not find significant commercial application. On the other
hand, the use of signals with higher frequencies was found to increase the feasibility of
transmitting information over long distances. It was this reason that drove the success of
radio communications, consolidating it as the main method of transmission at the end of the
19th century and during the 20th century and the beginning of the 21nd century.
Today, communications play a vital and practically indispensable role in the development,
sustainability and adaptation of cities and metropolises around the world. This progress
has been driven primarily by the development of new communication technologies, which use
light manipulation and guidance to send information with relatively low loss. This technology
is known as fiber optics [3].

Long time before the invention of lasers, scientists faced the challenge of finding ways to
direct light around corners or between points that could not be directly connected by a
straight line. Various experiments were carried out using devices such as mirrors and special
tubes; none of them really caught the eye, however, until John Tyndall came on the scene.
In 1870, in front of members of the prestigious British Royal Society, Tyndall demonstrated
how it was possible to guide a beam of light through a stream of falling water [4]. Although
Tyndall’s demonstration of this light-guiding phenomenon was remarkable, it took almost
another century before this idea was successfully applied in concrete new technologies.

Albert Einstein’s contributions in the 1920s, especially the introduction of the concept of
stimulated emission, laid the foundation for what would become a revolutionary invention
that would take technology based on optical phenomena to unprecedented levels. In 1960,
there was a milestone in the field of optics that stimulated interest in communications based
on optical instruments: the invention of lasers [5]. This innovation generated a paradigm shift
in light technologies, leading to significant attention and effort in this area. In the same year,
the term “optical fiber” was officially mentioned in a 1960 scientific paper by Narinder Singh
Kapany [6]. Although the concept of an optical technology with the capacity to transport
a large amount of information between two points (transmitter-receiver) had already been
established since 1955, when Kapany presented his doctoral thesis [4], it was not until then
that he took shape significantly.

Optical fibers proved their usefulness finding one of their first and most rudimentary ap-
plications in the medical field, specifically in endoscopy, using optical fibers as semi-flexible
endoscopes [7]. In 1965, Charles K. Kao and his team performed the necessary calculations
to make it possible to transmit light over long distances using optical fibers. The implemen-
tation of optical fibers made of glass allowed the transmission of light signals at distances
of up to 100 kilometers, significantly exceeding the 20 kilometers reached at that time [8].
This feat made optical fibers a candidate to replace coaxial cables. However, there was a
fundamental challenge related to attenuation in the fibers, which was around 1000 dB km−1
compared to 5 − 10 dB km−1 of coaxial cables. In addition to being lighter and more flexible
than coaxial cables, optical fibers also offered a lot of thing for improvement.

In just a few years since its introduction, the potential of optical fibers has been demonstrated.
Intensive effort devoted to these optical devices over a decade resulted in the reduction of at-
tenuation from 1000 dB km−1 to 5 dB km−1 . This achievement not only triggered advances
and developments in the optical fibers themselves, but also a series of new devices designed to
integrate with existing systems. These devices included detectors and light sources optimized
to work with optical fibers. This evolution fueled the growth of development and manufac-
turing of technology specifically adapted to optical fibers, which became the fundamental
platform for telecommunications [9].

The efforts made by companies such as Corning Glass in optimizing and improving man-
ufacturing processes for optical fibers have been significant. For example, the introduction
of dopants such as titanium in the silica structure made it possible for the first time to ob-
tain fibers with hundreds of meters in length and relatively low losses [11]. However, this
breakthrough was initially limited to the 0.8 − 0.9 µm wavelength range. Subsequently, by
beginning to dope the fibers with different alloys, the working range of the optical fibers was
extended to 1.1 to 1.6 µm, to take advantage of the improvements in performance that they
present in this interval. In this new wavelength window, remarkable results were achieved,
such as reducing fiber losses from 5 dB km−1 to as low as 0.15 dB km−1 , one of the lowest
values recorded in the literature. However, this range is also characterized by maximum
chromatic dispersion, which poses a new challenge in the design and operation of optical
systems.

To address the challenge of chromatic dispersion, various variants of optical fibers have been
developed, such as water zero peak fibers or modified dispersion fibers. These variants involve
altering the refractive index profiles of the fibers and even incorporating multiple nuclei into
the fiber structure. However, in the 1990s, there was a crucial breakthrough in photonic
crystal fiber technology, which is the central focus of this work: optical fibers with a regular
periodic microstructure of air holes along their length, known as as photonic crystal fibers.

The development of this type of fiber was based on the discoveries and work on photonic crys-
tals and the effect of the photonic band gap, published by Eli Yablonovitch and Sajeev John
in 1987 [12], [13]. Yablonovitch proposed an analogy with semiconductors, which present a
”band gap” of energy between valence electrons, related to chemical bonds, and conduction
electrons. Similarly, in the proposed structure, light with frequencies within the electromag-
netic band gap could not propagate, following the same principle as semiconductors.

In 1989, Yablonovitch, together with his collaborator T. J. Gmitter, made the first practical
demonstration of the existence of a complete three-dimensional photonic bandgap [14]. These
results came to the hands of Philip Russell, who in 1991 first proposed the concept and the
term ”Photonic Crystal Fiber” [15]. Russell and his team successfully engineered the first
solid-core photonic crystal fiber in 1996, using 217 silica capillaries (8 layers around the central
capillary), which were specially machined to be hexagonal on the outside and circular on the
inside. This achievement marked a milestone, almost five years after theoretical conception,
in the technological development of [16] photonic crystal fibers.

Photonic crystal fibers, also known as microstructure or ”holey fibers” due to their perfo-
rated structure, are mainly classified into two categories: solid core photonic crystal fibers
and hollow core photonic crystal fibers. Solid-core photonic crystal fibers have regions with
an unusual property: they transmit a single mode of light in any wavelength range. There-
fore, they form a completely innovative type of single-mode fiber capable of carrying more
optical power than conventional fibers. On the other hand, the transmission bands in hollow-
core photonic crystal fibers are limited, since there is only transmission in restricted bands
of wavelengths that coincide with the photonic band gaps. Photonic crystal fibers offer new
or improved characteristics beyond what conventional fiber provides, which means they are
finding an increasing number of applications in ever-expanding areas of science and technol-
ogy. Some of the most relevant applications in which photonic crystal fibers have ventured
are: lasers and high power amplifiers doped with rare earths, sensors (which will be studied
in this work) and fibers with highly nonlinear effects.

Since these types of fibers not only had unprecedented properties, but could also overcome
many of the intrinsic limitations of conventional optical fibers, they were quickly applied in
the field of sensing. The strength of photonic crystal fibers lies in their ability to adjust
the fiber’s transmission spectrum, mode shape, nonlinearity, dispersion, air-hole refraction,
and birefringence, among others, by varying the size and location of the holes in the cladding
and/or core. This makes it possible to achieve values that are unattainable using conventional
PCFSA fibers. In addition, compared to electronic sensors, optical fibers and photonic crystal
fibers have certain advantages, such as immunity to electromagnetic interference, light weight,
remote sensing capability, ease of design, wavelength multiplexability, and high sensitivity.
with a wide detection range.
Specifically, in the area of sensors, two phenomena that have been traditionally used in con-
ventional fibers ([18], [19]) have been adapted to photonic crystal fibers taking advantage of
their properties. These phenomena are SPR and LMR. The phenomenon of Surface Plas-
mon Resonance has been widely used to develop fiber optic sensors in environmental [21],
industrial [22], biological [23] and medical [24] applications. However, to induce the SPR
phenomenon in an optical fiber, whether photonic crystal or not, it is necessary to deposit
a thin layer of metal or a material that can excite collective oscillations of electrons, known
as surface plasmons, and efficiently couple the light [25]. Consequently, to achieve maximum
sensitivity in SPR-based fiber optic sensors, metals such as gold and silver, or their alloys,
are generally used, which significantly increases manufacturing costs. However, in recent
years there have been notable advances in fiber sensors based on the LMR phenomenon,
which offers a high sensitivity comparable or equal to that obtained by SPR [26]. Unlike
SPR, LMR has great flexibility in the materials that can be used to induce this phenomenon,
from polymers to metals, making LMR-based sensors a more versatile and potentially less
expensive option [28].

In this study, a solid core Photonic crystal Fiber was fabricated, this fiber operates by
the Modified Total Internal Reflection mechanism. The objective of varying the periodic
microstructure of the photonic crystal is to make the fiber monomodal, that is, that it can only
guide the fundamental mode, and that it presents significant losses. In this way, by applying a
suitable coating, it is sought to stimulate the phenomenon of Lossy Mode Resonance (LMR),
with the purpose of using as a sensor for refractive index detection applications. Extensive
research has been carried out in the field of optical fiber-based physical and chemical sensors
employing Lossy Mode Resonance (LMR) [27, 28]. When the cladding layer modes are in
phase with the fiber modes, coupling occurs between the fiber modes and the cladding layer
modes, resulting in a lossy mode. The LMR undergoes variations when there are changes in
the properties of the coating film, either in its refractive index or in its thickness, as well as
when alterations occur in the optical properties of the surrounding media. These variations
result in a detectable modulation in the transmission spectrum of the fiber, which constitutes
a measurable sensory effect.

In most cases, these devices employ standard optical fibers, whose propagation constants
in the core exceed the propagation constants of the coating. This results in the in-phase
modes of the fiber and film generating evanescent fields in the fiber cladding, requiring close
contact between the fiber core and the absorbent film. This contact is achieved by removing
or modifying a portion of the fiber coating and replacing it with a different coating material
[27, 28]. In this study, we will explore another alternative to achieve LMR by using a
special photonic crystal fiber (PCF), which is planned to be manufactured at the Centro
de Investigaciones en Óptica (CIO). When the PCF is composed by a limited number of air
channels in a dielectric matrix, the propagation constants of its modes tend to be smaller than
the refractive index of the matrix [29, 30]. Consequently, if the matrix is not constrained,
the fields of these modes expand beyond the fiber core, i.e., the modes experience losses due
to lack of confinement [29, 30]. In real PCFs, which feature bounded coatings and absorbent
films, these leaky modes can only be in phase with the leaky modes that are present in
the film coatings. The coupling of leakage modes from different waveguides can occur at
considerable distances between said waveguides. This allows obtaining the LMR without
the need to make local modifications to the fiber coating, simply by applying an absorbent
film directly on the outer surface of the fiber cladding. Since, in the current state of CIO
Infrastructure, PCFs can be manufactured as single-mode ones [29], the described LMR has
the advantage of allowing single-mode operation of various sensors and greatly reduces losses
due to irregularities in the PCF caused by diffraction.
List of Figures

1.1 Schematic representation of a beam incident to a medium with a lower refrac-


tive index (n0 ) at three different angles. . . . . . . . . . . . . . . . . . . . . . 8
1.2 Schematic representation of a beam incident to a medium with a lower refrac-
tive index (n0 ) at three different angles and the stepped-type refractive index
profile. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3 Schematic representation of a rectangular waveguide. The letter R in the
figure refers to the guide lining which has a refractive index n0 and the letter
N refers to the guide core which has a refractive index n1 . . . . . . . . . . . 11
1.4 Schematic representation of an optical fiber. . . . . . . . . . . . . . . . . . . 13
1.5 Distribution of the intensities for 4 modes LPlm for an optical fiber with V=10.
Image taken from [32] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.6 Variation of the part of the left (solid line) and right (dotted line) terms of the
equation (1.53) for the value of V = 6. Each intersection between the solid
lines and the dotted lines implies a specific value of b and also a mode LPlm . 17
1.7 Schematic representation of the Gaussian profile associated with the funda-
mental mode propagating in the optical fiber. L is the distance from the end
point of the fiber and the point where the light projected. . . . . . . . . . . . 18
1.8 Schematic representation of the cut-back method. . . . . . . . . . . . . . . . 19
1.9 Attenuation spectrum in a silica optical fiber in the infrared band. The image
was taken from [35] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.10 Schematic representation of the generation of losses due to bending in the fiber. 22
1.11 Schematic representation of the phenomenon of chromatic dispersion. . . . . 24

2.1 Schematic representation of a Photonic Crystal Fiber (PCF). . . . . . . . . . 26


2.2 Image of the first manufactured PCF reported by Philip Russell and his team
in 1996. Figure taken from [44] . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3 Representative schematic of the mTIR mechanism in a PCF . . . . . . . . . 27
2.4 Representative schematic of the Bandgap Mechanism in a PCF . . . . . . . . 27
2.5 Schematic representation of the equivalence between a photonic crystal fiber
and a conventional optical fiber. . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.6 Graph of the parameter V as a function of λ/∧ using the empirical method of
Saitoh and Koshiba and FEM. This image was taken from [45] . . . . . . . . 30
2.7 Graph of the effective refractive index nf sm as a function of λ/∧ using the
empirical method of Saitoh and Koshiba and FEM. This image was taken
from [45] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.8 Graph of the parameter W as a function of λ/∧ using the empirical method
of Saitoh and Koshiba and FEM. This image was taken from [45] . . . . . . 32
2.9 Graph of the effective refractive index nf sm as a function of λ/∧ using the
empirical method of Saitoh and Koshiba and FEM. This image was taken
from [45] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.10 Graph of the effective refractive index nf sm as a function of λ/∧ using the
empirical method of Saitoh and Koshiba and FEM. Image taken from [45] . 34
2.11 Dispersion parameter of a photonic crystal fiber as a function of wavelenght
for ∧ = 2 µm. Dashed curves, results of eq. (2.9) This image was taken from
[45] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.12 Graph of the MFD as a function of the wavelength. Image taken from [45] . 36
2.13 Graph of the beam divergence as a function of the wavelength. This image
was taken from [45] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3.1 Schematic of the manufacturing process for a PCF. . . . . . . . . . . . . . . 39


3.2 Diagram of the stretching tower that is located in the facilities of building A
of the CIO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3 Schematic design of the preform used to Fabricate the PCF. . . . . . . . . . 41
3.4 Cross-sectional images of the manufactured PCFs: (a) PCF (F1), (b) PCF (F2). 42
3.5 Diagram of the experimental setup used to carry out the attenuation measure-
ments of the fabricated PCFs. . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.6 Attenuation spectra for the fabricated PCFs. . . . . . . . . . . . . . . . . . . 43
3.7 Normalized frequency curves for each of the manufactured fibers. The gray
dashed line represents the value of Vef f of the fiber regime (single-mode or
multi-mode). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.8 Curves obtained by the empirical method of Koshiba and Saitoh of the nor-
malized attenuation constant for each of the fibers manufactured . . . . . . . 45
3.9 Effective Refractive index of the PCF coating as a function of λ for fibers F1
and F2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.10 Effective refractive index of the PCF as a function of λ for fibers F1 and F2. 46
3.11 Dispersion Behavior of the PCFs as a function of λ for the manufactured fibers
F1 and F2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.1 PCF schematic drawing after coating removal and cutting. . . . . . . . . . . 49


4.2 Schematic design of the configuration to measure the transmission in both
sections of the PCF before making the splice. . . . . . . . . . . . . . . . . . 50
4.3 Graph of the initial transmission spectrum for the fiber F1 data obtained using
the optical spectrum analyzer. . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.4 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.5 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 51
4.6 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.7 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 52
4.8 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.9 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 53
4.10 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.11 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 54
4.12 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.13 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 55
4.14 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.15 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 56
4.16 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.17 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 57
4.18 Photographs taken from the display of the splicing machine before (a) and
after splicing (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.19 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 58
4.20 Results of the transmission spectrum measurements for before (blue line) and
after (orange line) splicing the two PCF sections. . . . . . . . . . . . . . . . 59
4.21 Schematic representation of a tapered PCF. . . . . . . . . . . . . . . . . . . 60
4.22 Profile of the PCF taper #9 measured by using the CIO drawing tower. Pa-
rameters used to obtain the taper profile: Power 55 watts, Speed 1.4 units/s,
time 0.1 ms, diameter of waist requested is 44 µm . . . . . . . . . . . . . . . 61
4.23 Profile of the PCF taper #11 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.4 units/s,
time 0.1ms, diameter of waist requested is 44.2 µm . . . . . . . . . . . . . . 62
4.24 Profile of the PCF taper #13 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.5 units/s,
time 0.1 ms, diameter of waist requested is 42 µm. . . . . . . . . . . . . . . . 62
4.25 Profile of the PCF taper #16 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.6 units/s,
time 0.1 ms, diameter of waist requested is 45.5 µm . . . . . . . . . . . . . . 63
4.26 Profile of the PCF taper #21 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.7 units/s,
time 0.1 ms, diameter of waist requested is 45 µm . . . . . . . . . . . . . . . 63

5.1 Schematic representation of the outcome produced by the LMR phenomenon


on an optical fiber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.2 Transmission spectra for the investigated F1 fiber taper # 21 with the polyethyleneimine
coating in air: (a) for wavelengths of 550-1650 nm, (b) for wavelengths of 1180-
1210 nm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.3 Transmission spectrum for the investigated F1 fiber taper # 21 with the
polyethyleneimine coating: in water, 20% glycerin, 40% glycerin, 60% glyc-
erin for wavelengths of 1180-1210 nm. . . . . . . . . . . . . . . . . . . . . . . 68
List of Tables

1.1 Sellmeier Coefficients for SiO2 (Silica) . . . . . . . . . . . . . . . . . . . . . 23

2.1 Values of the coefficients ai and bi for the equation (2.6). . . . . . . . . . . . 30


2.2 Values of the coefficients ci and di for the equation (2.8) . . . . . . . . . . . . 32

3.1 Results of measurement of geometric dimensions and results of our MFD cal-
culations for fabricated PCFs. . . . . . . . . . . . . . . . . . . . . . . . . . 42

5.1 Conditions on the dielectric constant of the special coating that is used to
generate some of the resonance phenomena (SPR and LMR). The variables ϵ′r
and ϵ′r are the real and imaginary parts of the dielectric constant, respectively. 66
Contents

1 Fundamentals of Optical Fibers and Light Guide Theory 8


1.1 Total internal reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 Waveguides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.1 Numerical aperture and angle of acceptance . . . . . . . . . . . . . . 9
1.2.2 Planar waveguide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Optical Fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.1 TE modes of a step-index OF . . . . . . . . . . . . . . . . . . . . . . 14
1.3.2 Fundamental Mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4 Losses in OFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.4.1 Attenuation: Total Losses . . . . . . . . . . . . . . . . . . . . . . . . 19
1.4.2 Rayleigh scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.3 Losses due to absorption . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.4 Curvature Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.5 Dispersion in Optical Fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.5.1 Multimodal Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.5.2 Material Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.5.3 Waveguide Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.5.4 Polarization Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2 Photonic Crystal Fibers 25


2.1 Light-guiding mechanisms in PCFs . . . . . . . . . . . . . . . . . . . . . . . 26
2.1.1 Guided by Modified Total Internal Reflection (mTIR) . . . . . . . . 27
2.2 Empirical Model for the Theoretical Design of a PCF . . . . . . . . . . . . . 28
2.2.1 The V parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.2 The W Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2.3 Monomodal and Multimodal Regime . . . . . . . . . . . . . . . . . . 33
2.3 PCF Transmission Characteristics . . . . . . . . . . . . . . . . . . . . . . . . 33

3 Fabrication Process of the PCFs 38


3.1 Fabrication Process of Special PCF for Sensing Applications . . . . . . . . . 40
3.1.1 Fabrication Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.2 Theoretical features of the PCF given by the empirical model of Koshiba
and Saitoh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4 Conditioning and Evaluation of the PCF for Sensing Application 48
4.1 PCF-PCF, SMF-PCF-SMF splices . . . . . . . . . . . . . . . . . . . . . . . 48
4.2 PCF tapering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5 Refractive Index Sensing at the use of the Fabricated PCF 65


5.1 Principle of Detection: Lossy Mode Resonance . . . . . . . . . . . . . . . . . 65
5.2 Refractive Index Sensing Trials . . . . . . . . . . . . . . . . . . . . . . . . . 66

6 Conclusions 69
Chapter 1

Fundamentals of Optical Fibers and


Light Guide Theory

1.1 Total internal reflection


In the field of optics, the phenomenon that deals with the transmission of light as it moves
from a medium with a certain optical density (or refractive index) to another with a different
optical density, constitutes one of the fundamental pillars of optics study. However, it is
important to highlight that this phenomenon presents variations depending on the optical
density characteristics of the source medium compared to that of the new medium. That
is, if the optical density of the medium through which the light is propagating is lower or
higher than that of the medium into which it is about to enter, not only the transmission of
light can be observed, but also a reflection phenomenon in the direction of the light-emitting
medium. Next, we will look specifically at the scenario where light originates from a medium
with a higher optical density and how it interacts with that medium.

Figure 1.1: Schematic representation of a beam incident to a medium with a lower


refractive index (n0 ) at three different angles.

When light propagates from a medium of high optical density to a medium of lower optical
density, there are three possible results that are conditioned by the angle of incidence of light
on the medium of lower optical density. Figure (1.1) illustrates these three possible results
when varying the angle of incidence θi . When this angle is increased, a critical value known
as the critical angle θc is reached. Beyond this point, ie when θi > θc , the transmitted beam
disappears and the incident beam experiences total reflection at the interface.
Snell’s law establishes the relationship between the angles of incidence and refraction of light
when crossing a limit between two media. It is expressed by the following equation:
n0
sin(θi ) = sin(θt ) (1.1)
n1

Applying Snell’s law, we consider the scenario in which the angle of incidence (θi ) is such that
the angle of transmission (θt ) reaches π/2. In this situation, the light does not propagate
towards the medium with the lowest refractive index, but follows the interface, as illustrated
in the second diagram of the Figure (1.1). This particular angle at which this occurs is called
the critical angle (θc ). According to Snell’s law, this relationship implies that:
n0
sin(θi ) = sin(π/2) (1.2)
n1
n0
θi = θc = . (1.3)
n1

On the other hand, when the angle of incidence exceeds or equals the value of the critical
angle (θi ≥ θc ), the light fails to penetrate the second medium and gives rise to a phenomenon
of total internal reflection. In this circumstance, all the light is reflected back towards the
first medium, as observed in the third scheme of the Figure (1.1). Total internal reflection is
a highly relevant phenomenon in the field of optics, especially in fiber optic technology. Here,
this phenomenon is used to transmit light signals over long distances with minimal energy
losses. By making use of the total internal reflection at the interface between the core and
cladding of an optical fiber, light is confined and propagates along the fiber with exceptional
efficiency.

1.2 Waveguides
1.2.1 Numerical aperture and angle of acceptance
Generally speaking, both optical fibers and Planar waveguides exhibit a highly similar struc-
ture. Both consist of a core, whose function is to confine light, and in the case of optical
fibers, of a cladding (or substrate that covers the core). A schematic representation of how a
conventional planar waveguide or optical fiber directs a light beam is presented in the figure
(1.2). In essence, the core of a waveguide or optical fiber has a refractive index (n1 ) greater
than the refractive index of the cladding (n0 ). This gives rise, as explained in the previous
section, to a space where light can be confined and guided from one end of the structure to
the other, taking advantage of the phenomenon of total internal reflection.
Figure 1.2: Schematic representation of a beam incident to a medium with a lower
refractive index (n0 ) at three different angles and the stepped-type refractive index
profile.

In the context of a planar waveguide configuration, similar to the one illustrated in Figure
(1.2), it can be determined that the occurrence of the total internal reflection phenomenon
is subject to a specific condition at the interface that separates the core and cladding. This
condition can be expressed using Snell’s law as follows:
π 
n1 sin − ϕ ≥ n0 , (1.4)
2
It is then possible to obtain a mathematical expression as a function of the angle of incidence
(θ) since the angle ϕ is related to the angle of incidence by Snell’s law as:

sin(θ) = n1 sin(ϕ). (1.5)

Additionally, it is p
necessary to guarantee that the angle of incidence complies with the rela-
tionship sin(θ) ≤ n21 − n20 . In other words, there is a limit value for the angle of incidence
that allows incident light to be admitted into the waveguide configuration. This value is
related to the sine of the angle and is known as the numerical aperture (N A). By applying
this concept to the equation (1.4), it follows that:
q
−1
θ ≤ sin n21 − n20 = θmax , (1.6)
q
sin(θ) = n1 sin(ϕ) ≤ n21 − n20 , (1.7)

The difference between the refractive index of the core and the cladding is commonly on the
order of 10−2 . Therefore, we have that for θmax it can be approximated

n21 − n20 n1 − n0
q
2 2
N A = θmax ≈ n1 − n0 , ∆ = 2
≈ , (1.8)
2n1 n1

N A ≈ n1 2∆. (1.9)
1.2.2 Planar waveguide
So far, no mathematical formulation has been provided to address the propagation of light
from a wave perspective. However, in the case of a light beam traveling along a waveguide, it
is possible to carry out a mathematical description. This three-dimensional description for a
rectangular waveguide is achieved by Maxwell’s equations. This formulation can be extrapo-
lated to the context of waveguides with cylindrical symmetry, as is the case of optical fibers.
Since we are considering dielectric waveguides, it is relevant to introduce the permittivity
and permeability into Maxwell’s equations, such that ϵ = ϵ0 n2 and µ = µ0 , respectively.

Maxwell’s equations for a lossless, homogeneous, dielectric medium are written as

∂ H̃
∇ × Ẽ = −µ0 , (1.10)
∂t
∂ Ẽ
∇ × H̃ = −ϵ0 n2 , (1.11)
∂t
∇ · Ẽ = 0 (1.12)
∇ · H̃ = 0 (1.13)

Figure 1.3: Schematic representation of a rectangular waveguide. The letter R in


the figure refers to the guide lining which has a refractive index n0 and the letter N
refers to the guide core which has a refractive index n1 .

Here, Ẽ and H̃ are represented as the electric and magnetic fields, respectively. Equation
(1.10) corresponds to Gauss’s equation for electricity, while equation (1.11) is Faraday’s
equation of magnetic induction; both are presented in their vector form. In a first approach,
we will approach these equations using the rectangular coordinate system, since the waveguide
under consideration presents symmetry in this configuration. The structure of the waveguide
is illustrated in the figure (1.3). Therefore, we consider that the electric and magnetic fields
behave like plane waves in their propagation:
Ẽ = E(x, y)ei(ωt−βz) , (1.14)
H̃ = H(x, y)ei(ωt−βz) , (1.15)
In this situation, β denotes the propagation constant of the mode of light that is guided
along the waveguide. The two previous equations illustrate the distributions of the electric
and magnetic fields, which only undergo phase changes during their propagation along the
coordinate z in the waveguide. In contrast, the transverse fields (E(x, y) and H(x, y)) remain
constant when propagating in the waveguide. Substituting these considerations into the
equations (1.10) and (1.11), results in a system of equations involving both the electric field
components and the magnetic field components, adopting the following structure:
∂Ez
+ iβEy = −iωµ0 Hx (1.16)
∂y
∂Ez
− − iβEx = −iωµ0 Hy (1.17)
∂x
∂Ey ∂Ex
+ = −iωµ0 Hz (1.18)
∂x ∂y

∂Ez
+ iβEy = −iωµ0 Hx (1.19)
∂y
∂Ez
− − iβEx = −iωµ0 Hy (1.20)
∂x
∂Ey ∂Ex
+ = −iωµ0 Hz (1.21)
∂x ∂y
Given the symmetry shown in Figure (1.3) the electric and magnetic fields do not have a
dependency on the component and therefore the dependent part of y in the equations ((1.14))
y ((1.15)) can be written as e−iγy , with which the above systems of equations have the form
iβEy = −iωµ0 Hx , (1.22)
∂Ey
= −ωµ0 HZ , (1.23)
∂x
∂Hz
−iβHx − = iωϵ0 n2 (x)Ey , (1.24)
∂x

∂Ey
−iβEx − = −iωµ0 Hy , (1.25)
∂x
iβHy = iωϵ0 n2 (x)Ey , (1.26)
∂Hy
= iωϵ0 n2 (x)Ez . (1.27)
∂x
Maxwell’s equations simplify into these two systems of equations. In the first system, the
components Ey , Hx and Hz are involved, giving rise to the propagating modes in the guide
known as Electrical Transverse Modes, abbreviated as TE. This nomenclature comes from
the fact that the electric field has exclusively a transverse component in these modes. In
the second system of equations, the components Hy , Ex and Ez are found, generating the
propagating modes in the guide called Transversal Magnetic Modes, abbreviated as TM.
Here, the name is derived from the characteristic that the magnetic field only has a transverse
component. This analysis results in that in a planar waveguide, like the one shown in the
Figure 1.3, the propagation of electromagnetic waves confined to the guide is described by
the modes TE and TM.

1.3 Optical Fibers


Unlike planar waveguides, optical fibers (abbreviated as OF in English) have a cylindrical
structure that modifies the approaches discussed in the previous section, which were based
on Maxwell’s equations for a planar waveguide. In this section, we will carry out a theoretical
development of an optical fiber whose refractive index profile adopts a stepped shape. This
approach will allow us to understand the fundamental properties of optical fibers.

Figure 1.4: Schematic representation of an optical fiber.

In the case of an OF whose symmetry is cylindrical, the mathematical representation of the


electric and magnetic fields, unlike the previous chapter, is expressed as

Ẽ = E(r, θ)ei(ωt−βz) , (1.28)


i(ωt−βz)
H̃ = H(r, θ)e , (1.29)

substituting equations (1.28) and (1.29) into Maxwell’s equations (1.10) and (1.11) and con-
sidering that the refractive index profile does not depend on the coordinate θ it is possible
to obtain that
i  ∂E
z ωµ0 ∂Hz 
Er = − 2 β + , (1.30)
k n(r)2 − β 2 ∂r r ∂θ
i  β ∂E
z ∂Hz 
Eθ = − 2 + ωµ0 , (1.31)
k n(r)2 − β 2 r ∂θ ∂r
i  ∂H
z ωµ0 n(r)2 ∂Ez 
Hr = − 2 β + , (1.32)
k n(r)2 − β 2 ∂r r ∂θ
i  β ∂H
z 2 ∂Ez

Er = − 2 + ωµ0 n(r) . (1.33)
k n(r)2 − β 2 r ∂θ ∂r
The above system of equations allows us to know how the transverse electric (Ez = 0) and
transversal magnetic (Hz = 0) fields are related. In general the modes in an OF consist of
TE modes, TM modes and Hybrid modes (Ez ̸= 0, Hz ̸= 0)

1.3.1 TE modes of a step-index OF


To analyze the case where TE modes propagate through the fiber, we consider that EZ = 0
and therefore this generates that the system of equations shown above becomes:
i  ωµ ∂H 
0 z
Er = − 2 , (1.34)
k n(r)2 − β 2 r ∂θ
i  ∂Hz 
Eθ = − 2 ωµ 0 , (1.35)
k n(r)2 − β 2 ∂r
i  ∂H 
z
Hr = − 2 2 2
β , (1.36)
k n(r) − β ∂r
i  β ∂H 
z
Er = − 2 2 2
. (1.37)
k n(r) − β r ∂θ
Also using Maxwell’s equations we can obtain for Hz the equation
∂Hz 1 ∂Hz 2 2 2 l2
+ + [k n(r) − β − 2 ]Hz = 0, (1.38)
∂r2 r ∂r r
where l is an integer. The magnetic field in the core and cladding can be expressed as

Hz (r, θ, z) = R(r) cos(lθ), sin(lθ) (1.39)

Substituting the above in the equation (1.38) we have:


2
∂Hz ∂Hz 2r
r2 + r + [U − l2 ]Hz = 0, 0<r<a (1.40)
∂r2 ∂r a2
2
2 ∂Hz ∂Hz 2r
r 2
+r + [W 2 + l2 ]Hz = 0, r>a (1.41)
∂r ∂r a
(1.42)
where a is the radius of the OF kernel and the variables U and W are defined as

U = a(k 2 n21 − β 2 )1/2 , (1.43)


W = a(β 2 − k 2 n20 )1/2 , (1.44)

The equations (1.40) and (1.41) take the form of the Bessel equations. The solution to the
first of these equations takes the form Jl (x) and Yl (x), where x = U r/a. However, since
the solution Yl (x) becomes divergent as x approaches zero, this solution must be discarded,
leaving us only with the solution Jl (x) applicable to the core region of the fiber. . As for the
solution of the second equation, it involves the well-known modified Bessel functions: Kl (x′ )
and Il (x′ ), where x′ = W r/a. The function Il (x′ ) diverges when x′ tends to infinity, so it
must be discarded as a solution. With this reasoning, we obtain that the solutions for the
transverse magnetic field are expressed as follows:
A  U r 
H(r, θ) = J cos(lθ), sin(lθ) r<a (1.45)
Jl (U ) a
A  W r 
H(r, θ) = K cos(lθ), sin(lθ) r>a (1.46)
Kl (W ) a
By analyzing the continuity of ∂Hz /∂r for r = a, it is obtained that:
U Jl+1 (U ) W Kl+1 (W )
= l>0 (1.47)
Jl (U ) Kl (W )
U J1 (U ) W K1 (W )
= l=0 (1.48)
J0 (U ) K0 (W )
The above equations are transcendental equations that allow drawing curves that indicate
the number of modes present in the optical fiber, depending on the number of intersections
between the right and left sides of the equation, as well as the value of l. These equations can
be reformulated in terms of two crucial quantities in the analysis of optical fibers. On the
one hand, there is the normalized frequency V or the normalized waveguide parameter, and
on the other hand, the normalized propagation constant b. These magnitudes are defined as
follows:

V = (U 2 + W 2 )1/2 = ka(n21 − n20 )1/2 , (1.49)


β2
k02
− n20 W2
b= = , (1.50)
n21 − n20 V2
2
where βk2 is the modal index, and therefore we can have that the condition that there are
0
guided modes within the fiber have to fulfill the condition
β
n0 < < n1 , (1.51)
k
0 < b < 1, (1.52)
With which we can express the equations (1.47) and (1.48) as a function of b and V as

Jl−1 [V (1 − b)1/2 ] W Kl−1 [V (b)1/2 ]


V (1 − b)1/2 = l > 0, (1.53)
Jl [V (1 − b)1/2 ] Kl [V (b)1/2 ]
J1 [V (1 − b)1/2 ] W K1 [V (b)1/2 ]
V (1 − b)1/2 = l = 0. (1.54)
J0 [V (1 − b)1/2 ] K0 [V (b)1/2 ]

For a specific fixed value of l, a limited number m of solutions will arise. These solutions
are commonly known as LPlm and correspond to the modes that propagate inside an optical
fiber. The Figure (1.6) shows the modes allowed in a fiber with a normalized frequency of
V = 6.5. On the other hand, the spatial arrangement of some of these modes is illustrated
in the Figure (1.5).

Figure 1.5: Distribution of the intensities for 4 modes LPlm for an optical fiber with
V=10. Image taken from [32]

The above treatment, in general, can be adapted for both TE and TM modes and even HE
Hybrid modes with the appropriate notation and similar development.

1.3.2 Fundamental Mode


Specifically, in this research, there is a focus on single-mode optical fibers, which means that
they can only guide the fundamental mode. In this type of fibers, unlike multimodal fibers, it
is essential to carry out a modal analysis. To address this, the Gaussian Approximation is
generally employed, which involves representing the fundamental mode as a Gaussian beam.
This representation takes the form of a Gaussian function:
2 /w 2
ψ(r) = Ae−r , (1.55)

where w is the point size of the mode field pattern, which is associated with the quantity
d = 2w which is known as the mode field diameter, and r is the radius of the kernel of the
fiber. The intensity can then be calculated based on the equation (1.55) as
2 /w 2
I = I0 e−2r . (1.56)
(a) (b)

(c) (d)

(e) (f)

Figure 1.6: Variation of the part of the left (solid line) and right (dotted line) terms
of the equation (1.53) for the value of V = 6. Each intersection between the solid
lines and the dotted lines implies a specific value of b and also a mode LPlm .
Figure 1.7: Schematic representation of the Gaussian profile associated with the
fundamental mode propagating in the optical fiber. L is the distance from the end
point of the fiber and the point where the light projected.

Said intensity decays because its argument is a negative quadratic value and modulated by
the width of the fundamental mode w. Thanks to this approximation, it is also possible to
find a relationship between the diameter di of a beam and the diameter df of the light beam
when it passes and travels a distance as shown in the Figure (1.7), the above is represented
as
4 f
di = f. (1.57)
π df
Based on the previously exposed, the Marcuse equation ([33]) provides the opportunity to
establish the correspondence between the size of the radius of the fundamental mode that
propagates in the fiber (wf ) and the radius (a) of the fiber core itself, under the condition
that this is a single-mode optical fiber. The mathematical relationship that intertwines these
magnitudes is expressed as follows:
wf 1.619 2.879
= 0.65 + 3/2 + . (1.58)
a V V6
Since the radius of the OF core (a) is a value, generally determined in its manufacture, and
L being the distance from the end point of the fiber and the point where the light projected
it is possible to know the value wf = w and therefore it is possible to know the value of the
numerical aperture ( N A) as follows:

N A = n1 sin tan−1 (w/L) .



(1.59)
1.4 Losses in OFs
Similar to any physical system, energy losses always play a fundamental role in its develop-
ment and study, due to its direct influence on the performance and efficiency of the system
under consideration. Although optical fibers (OFs) are among the most efficient platforms
when it comes to guiding light, they are also susceptible to loss. These losses refer to the
alteration of the relationship between the energy supplied to the system and the energy
measured at its output. Over time, continuous effort has been devoted to improving the
performance of [34] OFs in this aspect.
In the context of optical fibers, various types of losses that impact light propagation can be
identified. In general terms, the losses present in optical fibers can be classified as follows:
• Absorption Loss
• Loss due to Curvature/Bending
• Scattering Loss

1.4.1 Attenuation: Total Losses


Attenuation, which refers to the reduction in light intensity along the fiber, represents the loss
of optical power that arises as a result of various sources of absorption, scattering, and other
effects within the transmission medium. Generally speaking, attenuation (σdB ) is quantified
in decibels (dB). One of the most widely used methods to measure attenuation is known as
the Cut-back method, which is illustrated in Figure 1.8.

Figure 1.8: Schematic representation of the cut-back method.

By using this method, the coupling loss factor between the light source and the fiber is
omitted. First, the power (Pi n) of the light that is guided along the fiber to the Optical
Analyzer (OSA) is measured. The second stage of this method involves cleaving the fiber to
reduce its length, followed by repeating the power measurement (Pout ). The expression that
defines the attenuation σdB is the following:
10 Pin
σdB = − log , (1.60)
∆L Pout
where ∆L represents the difference between the lengths L1 and L2 in Figure (1.8).

1.4.2 Rayleigh scattering


Within the field of optical fibers, one of the most prominent forms of loss is Rayleigh scat-
tering. This type of dispersion is manifested in optical fibers due to the small irregularities
present in the microscopic structure of the material that constitutes the fiber. These ir-
regularities can manifest as variations in the density or composition of the material at the
microscopic level.

Tf
αR ∝ kB , (1.61)
λ4
Here, kB denotes Boltzmann’s constant and Tf refers to a specific temperature known as
the cooling temperature. This cooling temperature reflects the thermal level to which the
material was subjected during its manufacturing process.

1.4.3 Losses due to absorption


In the previous section, we discussed how the material used in the manufacture of optical
fibers can generate substantial losses during the process of propagating light through the fiber.
In this context, various sources of absorption in optical fibers are identified, among which
impurities and defects in the material, absorption induced by doping elements and nonlinear
effects stand out. Since the scope of this research does not cover the study of nonlinear

phenomena present in the fiber, it is valid to state that, excluding these phenomena, mainly
three factors play a predominant role in the absorption in optical fibers: ultraviolet absorption
, infrared absorption and ionic resonance absorption.

Ultraviolet Absorption
When ultraviolet light falls on the fiber material, some atoms can undergo electronic transi-
tions. This phenomenon is a consequence of the atomic structure of the fiber material and its
optical properties. When ultraviolet light falls on the optical fiber, it ionizes the conducting
valence electrons, which causes a decrease in the power of the optical signal during its propa-
gation through the fiber. This reduction in signal strength can make it difficult to accurately
detect and recover transmitted information, which in turn can result in signal degradation
and lower transmission quality.

Infrared Absorption
With regard to infrared absorption, two fundamental mechanisms predominate: the absorp-
tion bands of the material and the vibrations inherent in the structure of the material. When
infrared light interacts with the optical fiber, losses occur due to the absorption of light that
is transformed into heat. This occurs due to specific absorption bands in the infrared range,
which are a result of the atomic or molecular structure of the material. These absorption
bands can arise from molecular vibrations or from electronic transitions internal to the ma-
terial.

Ion Resonance Absorption


Ionic resonance absorption in optical fibers occurs when dopant ions or impurities present
in the fiber material interact with incident light and absorb energy at a specific wavelength
that is consistent with an electronic transition of the ion. Dopant ions are chemical elements
deliberately added to the fiber material during its manufacturing process for the purpose of
modifying its optical properties. Some of the doping ions commonly used in optical fibers
are Erbium (Er) (in the range of λ = 1.5 µm), Thulium (Tm) (in the range of λ = 1.4 - 1.6
µm), Neodymium (Nd) (in the range of λ = 1.06 µm) and Ytterbium (Yb) (in the range of
λ = 1 - 1.1 µm).

Next, the attenuation spectrum of a conventional fiber is presented, where the contributions
of some losses present in optical fibers can be appreciated along a wavelength range of λ =
700 − 1800 nm.

Figure 1.9: Attenuation spectrum in a silica optical fiber in the infrared band. The
image was taken from [35]

Water absorption can cause attenuation of the optical signal, leading to a decrease in signal
strength and overall system efficiency. The absorption bands of water molecules in the
infrared region coincide with the transmission windows commonly in the infrared region.
This absorption leads to an increase in the fiber’s attenuation, causing signal loss therefore
high water absorption can limit the distance over which optical signals can be transmitted
effectively.

1.4.4 Curvature Loss


As discussed in section 1.2, from a geometric perspective, there is a fundamental requirement
in all waveguides to achieve light conduction: the total internal reflection condition (equation
(1.6)). Under this condition, if the angle is less than θmax , all the light will not propagate
along the waveguide. Upon reaching the core-cladding interface (as represented in the figure
(1.10)), as the angle of incidence increases, the total internal reflection intensifies, which can
cause losses of optical energy. If the condition is not satisfied, part of the light will be lost
and part will return to the core, causing notable losses in the bent portion of the fiber.

Figure 1.10: Schematic representation of the generation of losses due to bending in


the fiber.

1.5 Dispersion in Optical Fibers


The concept of dispersion in an optical fiber refers to the tendency of light to be scattered
and distorted as it passes through the fiber core. This phenomenon arises from the partic-
ular light-guiding characteristics inherent in optical fibers and the optical properties of the
materials used in their manufacture. Dispersion can occur for a number of reasons, all of
which contribute significantly to the total loss in the optical fiber in question.
In optical fibers, several types of dispersion can be identified. In general terms, dispersion in
optical fibers can be classified into the following categories:

• Multimodal Dispersion

• Material Dispersion
• Waveguide Dispersion

• Polarization Dispersion

1.5.1 Multimodal Dispersion


Optical fibers have the ability to propagate light in different modes. Each mode corresponds
to a propagation constant β and, therefore, to different group velocities. Due to the wave na-
ture of light, each mode experiences slight differences in propagation characteristics, including
path length and speed.

1.5.2 Material Dispersion


Material dispersion (Dm ) is the most influential dispersion component. It arises due to
variations in the speed of propagation of light in a specific material. This material scattering
can originate from the relationship between the refractive index of the material and the
wavelength, the intrinsic scattering properties of the material, as well as its absorption and
emission characteristics. Furthermore, it uniformly affects all wavelengths of light, regardless
of the refractive index profile of the fiber itself. Consequently, each wavelength experiences
a different refractive index. Mathematically, this is expressed as:
1 dnm
Dm = . (1.62)
c dλ
The refractive index nm , which is the refractive index of the material, is calculated using the
Sellmeier equation, which has the form

B1 λ2 B2 λ2 B3λ2
n2m (λ) =1+ 2 + + . (1.63)
λ − C1 λ2 − C2 λ2 − C3
The constants Bj and Cj (j = 1, 2, 3) are coefficients that are determined by the type
of material used in the optical fiber. Optical fibers are usually made with silica as the
predominant material. In the case of silica, the coefficients Bj and Cj acquire the following
values:

Material B1 B2 B3 C1 (10−3 µm2 ) C2 (10−2 µm2 ) C3 (µm2 )


Silica 0.696166300 0.407942600 0.897479400 4.67914826 1.3512031 97.9340025

Table 1.1: Sellmeier Coefficients for SiO2 (Silica)

1.5.3 Waveguide Dispersion


Even if we might assume that material dispersion is completely absent, delay distortion due
to finite frequency bandwidth would still persist because the speed of the group of modes in
a waveguide structure is often frequency dependent. This alteration in delay, caused by the
effect of bandwidth, is called waveguide dispersion.

The sum of the waveguide dispersion and the material dispersion is known as chromatic
dispersion. Chromatic dispersion is commonly quantified by the dispersion parameter (D) or
the dispersion coefficient. This parameter represents the rate of change of the group velocity
in relation to the wavelength. It is usually expressed in units of picoseconds per kilometer per
nanometer (ps/km/nm). Positive dispersion (D > 0) causes broadening of the pulse at longer
wavelengths, while negative dispersion (D < 0) causes broadening at shorter wavelengths.

Figure 1.11: Schematic representation of the phenomenon of chromatic dispersion.

1.5.4 Polarization Dispersion


Analogous to multimodal dispersion, the different modes that propagate through optical
fibers have different polarizations, which leads to light propagating at different speeds. This
results in light scattering, which can result in signal distortions and limit the performance of
optical communication systems.

Due to structural imperfections, optical fibers have a certain degree of birefringence, which
means that the refractive index varies according to the polarization of the light that propa-
gates through the fiber. When light traveling through the fiber has multiple states of polar-
ization (horizontal, vertical, diagonal, etc.), the various polarization components experience
varying effective refractive indices. As a consequence, these components propagate through
the fiber at different speeds, creating dispersion over distance and time.
Chapter 2

Photonic Crystal Fibers

In the previous chapter, the focus was on conventional fiber optics. However, there is a special
focus on another type of fiber known as Photonic Crystal Fibers or PCF for its acronym in
English. These fibers are especially noted for their ability to offer superior flexibility in
customizing optical properties. The photonic crystal structure can be designed and tailored
to suit the particular needs of an application, such as scattering manipulation, loss reduction,
or propagation characteristic optimization. This is especially desired in sensor applications.

PCFs are a type of optical fiber with microstructures that consist of a geometric arrangement
made of very small hollow tubes that run longitudinally throughout the fiber. Due to the
periodicity of these structures and their photonic properties, such as the photonic bandgap
effect, they are considered a type of photonic crystal, giving rise to the term ”Photonic Crystal
Fibers” or PCFs. In the late 1980s, the team led by Eli Yablonovitch [40, 41] successfully
described the phenomenon of the photonic bandgap effect. Shortly after, in 1991, Phillip
Russel developed the idea of these new fibers and named them ”Holey Fibers” [43]. Five
years later, Russell and his team developed and carried out the first experiamental tests
with a silica fiber, whose core was solid and was surrounded by air voids in its cladding.
These tests demonstrated the possibility of guiding light through fibers with photonic crystal
structures [44]. The fiber made by Russell’s team is shown in the Figure (2.2).

The PCF structure is defined by three characteristics, as shown in the figure (2.1): the
distance from the center of one air gap to another, known as ”pitch” (∧), the air gap diameter
(d). In solid core fibers, light is guided by a modified total internal reflection mechanism. In
the case of PCFs, the effective normalized frequency is defined as:

2π∧
q
Vef f = n20 − n21 , (2.1)
λ
where, again, n0 and n1 are the refractive indices of the core and cladding respectively. Within
the context of PCFs, the cutoff frequency, denoted as Vc , has a critical value that establishes
a limitation: below this value, only one mode can be guided, the so-called fundamental
mode. For PCFs, this critical value is set to Vc = 4.1 [42]. In other words, if the value of
(a) (b)

Figure 2.1: Schematic representation of a Photonic Crystal Fiber (PCF).

Vef f is consistently kept below Vc , thus ensuring that the PCF exhibits monomodal behavior
continuously throughout the spectrum.

Figure 2.2: Image of the first manufactured PCF reported by Philip Russell and his
team in 1996. Figure taken from [44]

2.1 Light-guiding mechanisms in PCFs


In contrast to conventional optical fibers, where electromagnetic modes are guided by total
internal reflection in the core region with a slightly higher refractive index than the cladding,
photonic crystal fibers (PCFs) have a periodic structure that gives rise to two different light
guide mechanisms:
• Guided by Modified Total Internal Reflection (mTIR).
Figure 2.3: Representative schematic of the mTIR mechanism in a PCF

• Guided by effective refractive index effect.

Figure 2.4: Representative schematic of the Bandgap Mechanism in a PCF

In this work, the mTIR light-guiding mechanism is of particular interest, since the develop-
ment and theoretical approaches used to describe the characteristics of the PCF in question
are based on the use of this mechanism.

2.1.1 Guided by Modified Total Internal Reflection (mTIR)


The first type of guidance in PCFs is very similar as in OFs, which is TIR. Due to the pho-
tonic crystal structure, an energy gap or photonic bandgap is generated, which prohibits the
propagation of light with certain wavelengths in the PCF core. However, near the photonic
bandgap, due to the interactions between light and the defects of the periodic structure of
the photonic crystal, a light mode is formed inside the fiber hollow core, which is propagated
due to the RIT phenomenon.

If we take into account a photonic crystal fiber, as illustrated in the figure (2.1a), in which
the core is solid, we can establish an equivalence with a conventional optical fiber. In making
this comparison, we consider that the core of the PCF corresponds to the region bounded
by the first ring of air holes in the photonic crystal structure. Therefore, the lining refers to
everything that is outside this region.
Figure 2.5: Schematic representation of the equivalence between a photonic crystal
fiber and a conventional optical fiber.

When establishing this equivalence, it is important to take into account that the refractive
index of a conventional optical fiber (n0 ) is equal to the refractive index of what we have
called the core of the photonic crystal fiber (nmat ). Thanks to this correspondence, we can
apply the theoretical foundations developed for optical fibers, as described in works such as
”Introduction to Fiber Optics” by Ajoy Ghatak [38]. Specifically, we focus on the case of
conventional optical fibers with a step-type refractive index profile.

2.2 Empirical Model for the Theoretical Design of a


PCF
The modeling and description of the characteristics of a PCF have traditionally been ap-
proached using numerical approximations (or specialized software based on such approxima-
tions), such as the finite difference method (FEM) [47, 48] or the multipole method. This
is due to the relative complexity of achieving a precise description of its characteristics in
an exact manner. Given the demanding demands on time and computational resources nec-
essary to carry out these numerical simulations, in 2004 Masanori Koshiba and Kunimasa
Saitoh published the article about applicability of classical optical fiber theories to photonic
crystal fibers [45]. In said article they demonstrated the possibility of using an analytical
approach based on relevant parameters for the characterization, design and manufacture of
conventional optical fibers, such as the normalized frequency (V ) or V the parameter, and
the normalized transverse attenuation constant, W or parameter . This approximation is ap-
plicable not only for conventional optical fibers, but also for photonic crystal fibers (PCFs),
since, as mentioned above, it is possible to establish an equivalence between both types of
fibers.

In the mentioned model, empirical relationships are established for V , W and other relevant
features, based on the fundamental geometric parameters of a PCF. These parameters include
the size of the air holes (d) and the distance between the centers of said holes or pitch (∧).

2.2.1 The V parameter


We take into account a fiber similar to the one represented in the figure (2.1 a), assuming
that the structure is made of silica, which has a refractive index of 1.45. Following what was
discussed in the previous section, we propose that V is expressed as follows:

π q √
V = 2 aef f n20 − n2f sm = U 2 + W 2 , (2.2)
λ

con

π q
U = 2 aef f n20 − n2ef f , (2.3)
λ q
π
W = 2 aef f n2ef f − n2f sm , (2.4)
λ

where n0 represents the refractive index of the core, nf sm is the refractive index of the
cladding, nef f indicates the effective index of the fundamental mode guided by the fiber, and
aef f denotes the effective radius of the core PCF core. Despite the fact that previously in
the literature Mortensen et al. [49] had proposed a definition of V as shown in the equation
(2.1), its applicability is difficult and therefore it was discarded as an expression for V . Using
a trial and error approach, Koshiba and Saitoh discovered that each point on a plot obtained
through the finite difference method (FEM) could be adequately modeled by the equation:

λ d A2
V , = A1 + . (2.5)
∧ ∧ 1 + A3 exp(A4 λ/∧)

Where the parameters Ai depend on some coefficients ai and bi and on the geometric param-
eters of the fiber d and ∧ as shown below

 d bi1  d bi2  d bi3


Ai = ai0 + ai1 + ai2 + ai3 , (2.6)
∧ ∧ ∧

Table 2.1 shows the values of the coefficients ai and bi


Coeficiente i=1 i=2 i=3 i=4
ai 0 0.54808 0.71041 0.16904 -1.52736
ai 1 5.00401 9.73491 1.85765 1.06745
ai 2 -10.43248 47.41496 18.966849 1.93229
ai 3 8.22992 -437.50962 -42.4318 3.89
bi 1 5 1.80 1.7 -0.84
bi 2 7 7.32 10 1.02
bi 3 9 22.80 14 13.40

Table 2.1: Values of the coefficients ai and bi for the equation (2.6).

By using the values presented in table 2.1 and substituting them into the equation (2.5),
we can calculate other parameters, such as the effective refractive index of the cladding
(nf sm ). In the work of Koshiba and Saitoh [45], it is reported that the values obtained for
nf sm using this approach differ by less than 0.25% compared to the results derived from the
finite difference method (FEM). Figure (2.6) shows the behavior of the parameter V obtained
through the empirical method proposed by Koshiba and Saitoh (solid line), as well as through
FEM (unfilled white dots). Similarly, the Figure (2.7) presents the behavior of the effective
cladding index

Figure 2.6: Graph of the parameter V as a function of λ/∧ using the empirical method
of Saitoh and Koshiba and FEM. This image was taken from [45]
Figure 2.7: Graph of the effective refractive index nf sm as a function of λ/∧ using
the empirical method of Saitoh and Koshiba and FEM. This image was taken from
[45]

2.2.2 The W Parameter

Carrying out the same procedure for the parameter V , it was possible to find an empirical
relationship for the parameter W of the form

λ d B2
W , = B1 + . (2.7)
∧ ∧ 1 + B3 exp(B4 λ/∧)

Where the parameters Bi depend on some coefficients ci and di and on the geometric param-
eters of the fiber d and ∧ as shown below

 d di1  d di2  d di3


Bi = ai0 + ci1 + ci2 + ci3 , (2.8)
∧ ∧ ∧

In the table 3 the values that the coefficients ai and bi are shown
Coeficiente i=1 i=2 i=3 i=4
ci 0 -0.0973 0.53193 0.24876 -5.29801
ci 1 -16.70566 6.70858 2.72423 0.05142
ci 2 67.13845 52.04855 13.28649 -5.18302
ci 3 -50.25518 -540.66947 -36.80372 2.7641
di 1 7 1.49 3.85 -2
di 2 9 6.58 10 0.41
di 3 10 24.80 15 6

Table 2.2: Values of the coefficients ci and di for the equation (2.8) .

Using these values, it is possible to calculate both the values of W (Figure 2.8) and determine
the values of nef f relative to λ/∧. It was observed that the maximum error for nef f using
this method only differs by 0.15% compared to the FEM (Figure 2.9).

Figure 2.8: Graph of the parameter W as a function of λ/∧ using the empirical
method of Saitoh and Koshiba and FEM. This image was taken from [45]
Figure 2.9: Graph of the effective refractive index nf sm as a function of λ/∧ using
the empirical method of Saitoh and Koshiba and FEM. This image was taken from
[45]

2.2.3 Monomodal and Multimodal Regime


As is common in conventional optical fibers, the cutoff frequency, which is the value of
the normalized frequency V that marks the beginning of the multimode regime, is set to
V = 2.405. Starting from this value for V and considering the equations (2.2) and (2.5),
it is possible to construct a graph that defines the regions or values for which the photonic
crystal fiber is multimodal or single-mode, depending on the geometric parameters of the
PCF (∧ and d) and the wavelength. In the previously cited work by Koshiba and Saitoh, it
was concluded that when the ratio d/∧ ≤ 0.4, the photonic glass fiber is considered infinitely
single-modal, regardless of the value of the wavelength λ of the light that is guided through
the fiber, as visualized in the figure (2.10). This figure not only presents theoretical results,
but also experimental results for an optical fiber with a stepped index profile.

2.3 PCF Transmission Characteristics


Based on the previous results obtained for nef f , nf sm , V and W , we can use them to determine
essential characteristics of the PCF, such as dispersion, mode diameter, divergence of the
guided beam in the fiber and the losses due to splices. In the previous section, we used the
equations (2.2), (2.4), (2.5) and (2.7) to examine the behavior of nef f and nf sm . Now we
can use the values of the obtained refractive indices to analyze the dispersion in the PCF.
Figure 2.10: Graph of the effective refractive index nf sm as a function of λ/∧ using
the empirical method of Saitoh and Koshiba and FEM. Image taken from [45]

Assuming that the dispersion follows the form:

λ d2 nef f
D=− + Dm , (2.9)
c dλ2

where c is the speed of light and Dm represents the material dispersion, defined by the
Sellmeier relation presented previously in the equation (1.63). In this equation, waveguide
dispersion and material dispersion are independent, allowing the results obtained in previous
sections to be used to analyze the behavior of D. This analysis is presented in the figure
(2.11), where the case for ∧ = 2 µm is illustrated.
Figure 2.11: Dispersion parameter of a photonic crystal fiber as a function of wave-
lenght for ∧ = 2 µm. Dashed curves, results of eq. (2.9) This image was taken from
[45]

Similarly, we can obtain the mode field diameter or MFD using the Marcouse formula (equa-
tion (1.58))

wef f 1.619 1.619


= 0.65 + 3/2 + 6 , (2.10)
aef f Vef f Vef f

In this relationship, wef f denotes half the fundamental mode diameter (MFD) and is known
as the effective modal point size. Figure 2.12 shows the results obtained for different PCFs. In
addition, a comparison with the results predicted by the theory of step-type refractive index
fiber optics (SIFA) is made, demonstrating that the proposed results reasonably correlate
with those obtained by the empirical approach.
Figure 2.12: Graph of the MFD as a function of the wavelength. Image taken from
[45]

Regarding the size of the modal point, the Gaussian beam propagation theory establishes
that the angular aperture of the beam (θ) is determined by the following expression:

−1
 λ 
θ = tan (2.11)
πwef f

The results of the equation (2.11), based on the SIFA, align approximately with the measure-
ments and the numerical results obtained through the use of the FEM, which are presented
below in the Figure (2.13):
Figure 2.13: Graph of the beam divergence as a function of the wavelength. This
image was taken from [45]

In the process of designing a PCF, we rely on empirical relationships for the V and W
parameters of PCFs, which depend exclusively on the diameter and spacing of the air holes.
We validated the precision of these expressions by comparing them with the results obtained
by the full vector finite element method (FEM). Through these empirical relationships, it
is possible to easily estimate the essential properties of PCPs without the need to resort to
detailed numerical calculations.
Chapter 3

Fabrication Process of the PCFs

There are several processes for the fabrication of photonic crystal fibers such as stack and
draw, 3D printing, sol-gel, etc. In the case of our work, the stack and draw method was
used, which in the past has given a good number of results [52], [51], [31].

In the manufacturing process of a solid core PCF using the stack and draw method, the
first step is to design what is known as preform, which consists of a structure made up of a
set of tubes and rods from pure silica with about 1 mm in outside diameters. These tubes
and rods are arranged in such a way that they have the desired geometry and number of air
channel rings of the photonic crystal to be used in the fiber, this is known as a stack. While
a core rod and tubes determine the hexagonal periodic structure of the photonic crystal, the
purpose of the rest rods in the stack design is to keep the structure fixed during stretching
of the preform, thus ensuring that it keeps the desired shape. A detailed description of this
method can be found in [50].

In the first stage of the stack and draw method, each of the tubs and rods that will be used
later in the stack is stretched in an optical fiber stretching tower. Below in the Figure (3.1)
there is a schematic of the manufacturing process for a PCF.
Figure 3.1: Schematic of the manufacturing process for a PCF.

It is relevant to note that there is the possibility of adjusting the dimensions of the external
diameter of the tubes and rods by varying the speed of extraction of the material from the
furnace by means of a convenient electronically controlled pulley system. It is also possible
to alter diameters of the tubes and rods by adjusting the temperature of the melting furnace,
which operates around 1950 degrees Celsius. This temperature is usually slightly lower than
that used in conventional optical fibers.

Once the stack is formed, it is stretched in the drawing tower. Initially, under the influence
of gravity, a glass drop is formed, which is carefully stretched downwards to obtain a fine
thread. This fine thread is pulled at considerable speed to further reduce its diameter.
The yarn then passes through a container containing polymer coating dye, and finally the
fiber undergoes a complex and rigorous UV curing process. This process involves the use
of a highly sophisticated array of ultraviolet lamps housed in a specialized structure with
advanced cooling systems. A schematic of the existing stretching tower at the Centro de

Investigaciones en Óptica A.C. is shown in Figure (3.2).


Figure 3.2: Diagram of the stretching tower that is located in the facilities of building
A of the CIO

Unlike conventional optical fibers that are manufactured with materials of different refractive
index (generally silica and some dopants), PCF are manufactured from a single material
(silica) depending on the desired application. However, the stretching of PCFs is more
difficult than that of conventional optical fibers. In a conventional fiber, the first step in
manufacturing the fiber begins with obtaining what is called a preform, which is a cylindrical
block of material that contains the necessary composition and refractive index configuration
(which is generally staggered). To create the desired fiber, the central core of the preform
has a slightly higher refractive index than the surrounding cladding. This refractive index
difference allows light to propagate through total internal reflection in the fiber core. The
preform is obtained using different deposition techniques, such as modified chemical vapor
deposition or (MCVD), with which low levels of impurities and high control of dopants for
the core are obtained.

3.1 Fabrication Process of Special PCF for Sensing Ap-


plications
For fiber optic sensors, the optical fibers used must have a strong evanescent field at some
part of their length. There are various methods to improve the evanescent field of optical
fibers, such as thinning, acid corrosion, polishing, etc. It is known that for solid core PCFs
it is possible to improve their evanescent field by decreasing the number of air channel rings
in their microstructured cladding. In the case of our PCF, we reduced the number of air
channel rings to two, thus generating a greater evanescent field. PCF also typically connects
with standard single-mode fibers in its applications as optical sensors. To have low loss in
such connections, solid core PCFs should have a core diameter of approximately 9 µm. A
limited number of silica tube dimensions are available in our laboratory. The most desirable
dimensions for our special PCFs for sensing applications are tubes with an inside diameter
of 9.5mm and an outside diameter of 19.9mm.

As mentioned in the previous section, the method used to manufacture our fibers is Stack
and Draw one. The design of a preform for the design of our PCFs is presented in Figure 3.3.
The preform is based on a silica liner tube with an inner diameter of 9.5mm and an outer
diameter of 19.9mm.

Figure 3.3: Schematic design of the preform used to Fabricate the PCF.

Two full hexagonal layers of pre-fabricated silica capillaries and additional rods were stacked
around a solid core. Some support rods were also added to minimize deformation of the
preform during PCF fabrication and also to prevent unwanted channels from appearing.
We fabricated the fibers from the aforementioned preform in a single pass, heating them
with a Centorr Vacuum Ind. furnace and using an improved conventional fiber draw tower
(Heathway Inc.). Using the same preform design shown in Figure 3.3, two PCFs (F1 and
F2) with outer diameters of 124 ± 1 µm were fabricated using different nitrogen pressures
inside the PCF preform ( 100 and 110 mBar, respectively) and different drawing speeds (25
and 32 m/s, respectively). Both PCFs were finally coated with a special polymer protective
coating, giving an outer diameter of 220 ± 2 µm. The fabricated PCF samples, with lengths
of about 300 m, were wound on spools of 16 cm in diameter.

3.1.1 Fabrication Results


Figure 3.4 shows light microscope images (obtained with a 40x objective) of the cut end
for the fabricated PCFs, with values of d/∧ = 0.36 (Figure 3.4a) and 0.43 (Figure 3.4b),
respectively; here, d is the diameter of the PCF air channels, and ∧ is the spacing of the air
channels.

(a) (b)

Figure 3.4: Cross-sectional images of the manufactured PCFs: (a) PCF (F1), (b)
PCF (F2).

To obtain the geometric dimensions of d and ∧ for the manufactured fibers, an atomic force
microscope was used (Digital Instruments, with a resolution of 50 nm for images of 30 µm
x 30 µm). 3 measurements were made for d and ∧ for each of the manufactured PCFs, and
then their averages were calculated. Using formulas from [4, 5], we also calculated mode field
diameters (MFDs) at the 1.31µm wavelength for the fabricated fibers. The results obtained
are shown in the following table.

Fibers davg (µm) ∧avg (µm) d/∧avg MFD (µm)


F1 3.31 9.10 0.36 6.70
F2 3.93 9.19 0.43 6.09

Table 3.1: Results of measurement of geometric dimensions and results of our MFD
calculations for fabricated PCFs.

Subsequently, measurements of the attenuation of the fabricated PCFs were carried out by
the use of the method known as cut-back, which is shown in the Figures (3.5a, 3.5b)
(a)

(b)

Figure 3.5: Diagram of the experimental setup used to carry out the attenuation
measurements of the fabricated PCFs.

We use a white light source (AQ-4305) and an optical spectrum analyzer (AQ 6315E) for our
attenuation measurements. The attenuation spectra for the fabricated fibers are presented
in Figure (3.6).

Figure 3.6: Attenuation spectra for the fabricated PCFs.

In this figure it can be seen that the attenuation spectra for the fabricated fibers have
three absorption peaks, located at 950 nm, 1244 nm and 1383 nm, due to the presence
of OH− functional groups. The fabricated fibers are single-mode in the 550 to 1700 nm
wavelength range, and the fibers have sufficiently high optical losses, more than 40dB/km.
The attenuation in the F1 fiber is greater than in the F2 fiber because the values of d and
λ, as can be seen in the table 3, are greater for F2 and therefore the amount d/∧ is smaller
for fiber F1; thus, it is seen that it is possible to change the attenuation of the fundamental
mode PCF in our fibers by varying d/∧. Estimations carried out in [55] demonstrated that
such fibers can be used for fiber optic sensors of pressure or refractive index, and to detect a
nanoscale adsorption layer of ammonia molecules deposited on a thin specialized absorbing
layer.

3.1.2 Theoretical features of the PCF given by the empirical model


of Koshiba and Saitoh

Based on the results previously obtained from the measurements carried out on both fibers,
as detailed in chapter 2.2, we can now extract fundamental characteristics of the fiber. Using
the equations (2.5, 2.7), we can determine the behavior of Vef f and Wef f . The results of these
calculations are presented in Figures (3.7, 3.8).

Figure 3.7: Normalized frequency curves for each of the manufactured fibers. The
gray dashed line represents the value of Vef f of the fiber regime (single-mode or multi-
mode).
Figure 3.8: Curves obtained by the empirical method of Koshiba and Saitoh of the
normalized attenuation constant for each of the fibers manufactured

In Figure 3.7, it can be seen that, in the case of the manufactured fibers, both maintain a
single-mode behavior throughout the range of wavelengths (from 0 to 15 µm). This feature
is of utmost importance, as our goal is to achieve effective coupling of the fundamental mode
with the cladding leakage modes.

Likewise, similar results were obtained for the effective refractive index behaviors in both the
core and the cladding. In Figures (3.9, 3.10), it is observed that the effective refractive index
is lower than the effective refractive index of the coating. This phenomenon causes a violation
of the total internal reflection (TIR) condition explained in chapter one, meaning that light
will not be reflected internally in the core. Instead, it will spread into the coating, escaping
the fiber. This leads to a significant improvement in the coupling of the fundamental mode
with the cladding modes.
Figure 3.9: Effective Refractive index of the PCF coating as a function of λ for fibers
F1 and F2.

Figure 3.10: Effective refractive index of the PCF as a function of λ for fibers F1 and
F2.

To calculate the dispersion in these fibers, we have used the results previously obtained for
nef f , thus allowing us to carry out the calculations of the equation (2.9). Next, we present
the results obtained for dispersion.

Figure 3.11: Dispersion Behavior of the PCFs as a function of λ for the manufactured
fibers F1 and F2.

It is notable that the dispersion value in the fabricated PCF is considerably lower compared
to other previously fabricated PCFs, as documented in [52], [51], [31]. However, it is relevant
to point out that in the manufactured PCF, it is not feasible to reach a zero dispersion point.
It is important to note that, depending on the intended application (detection), the absence
of a scattering zero point does not necessarily represent a drawback for its implementation.
Chapter 4

Conditioning and Evaluation of the


PCF for Sensing Application

In order to employ our PCF for sensing applications, it is important to perform conditioning
and evaluation of the characteristics of the fiber in question to ensure accurate and consistent
measurements of the parameters of interest, for sensing applications, it is imperative to find
a satisfactory splicing regime that allows for splicing lengths of our PCF either with sections
of the same PCF or with conventional fiber (SMF-28). This is due to the fact that in a
part of the fabricated PCF it is necessary to remove the coating and deposit a new one,
which must have certain necessary characteristics that allow the phenomenon of Lossy mode
Resonance or LMR [52]. By splicing the PCFs it is possible to increase the initial length
of the sensing system, that is to say, by using conventional fiber it is possible to transmit,
with low losses, a signal until reaching the sensing section where the signal can be later
modified and transmitted to another section of conventional fiber. In our case, splicing tests
were performed, using a Fujikura model fsm100p splicer or fusion splicer, using sections
of our PCF and a standard fiber (SMF-28), in order to verify that the properties of the
transmission spectrum remained consistent when performing a PCF-PCF or SMF-PCF-SMF
splice in such a way that a system of much longer length with conventional fibers could be
adapted. Additionally, to improve the sensitivity of our several PCF tapers were fabricated
on the PCF using a Vytran 3400 glass processor machine.

4.1 PCF-PCF, SMF-PCF-SMF splices


In order to obtain good splices between different fiber sections, a splicing machine Fujikura
model fsm100p was used. This equipment uses an electric arc to heat the fibers above the
melting point (in this case of silica) so that the ends of the fibers are softened and then, with
high-precision electric motors, pushes the fibers so that they stay together The arc continues
until both fibers are fused together strong enough to hold. Since the equipment is designed to
make splices, mostly in conventional optical fibers, different tests were carried out to obtain
a splice so that the PCF would maintain its guiding properties as well as possible. Due to the
microstructure presented in the fabricated PCFs, and the method used by the splicing team,
it was quite probable that the air gaps in the microstructure would collapse and therefore
the transmission properties generated by said microstructure would not be realized. In our
PCFs, the core diameter is similar to that of a conventional optical fiber (SMF-28). At the
first time, the equipment was calibrated to perform an optimal splicing conventional single-
mode optical fibers in such a way that said configuration would be used as a reference point
to start the splicing tests with our PCFs.

To perform PCF splicing tests, it is first necessary to perform some conditioning on the PCF
so that splicing is optimal. First, two independent sections of the PCF are taken, the first
one is connected to the light source, which, as mentioned in the previous chapter, is a white
light source (AQ-4305), and the second section is connected to the optical spectrum analyzer
(AQ 6315E). The ends of both fiber sections that are not connected to any of the equipment
undergo two procedures; The first one consists of removing the polymer coating that the
optical fiber has in such a way that it is exposed as shown in the figure (4.1). For this, a
splicer attachment was used which allows removing the fiber coating when heating blades are
more than 250◦ . Once the coating is removed, a cut is made in the fiber so that there are no
imperfections at the end of the fiber section.

Figure 4.1: PCF schematic drawing after coating removal and cutting.

Once both procedures have been carried out, both sections of the fiber are placed in holders
placed in the splicer in such a way that they are separated by a short distance, defined
directly by the equipment, as shown in the figure (4.2), in such a way that it is possible
to measure the transmission in the PCF before the splicing of both sections, an assembly
scheme to carry out this measurement can be seen below in the Figure (4.2).
Figure 4.2: Schematic design of the configuration to measure the transmission in
both sections of the PCF before making the splice.

The initial transmission spectrum for the fiber F1 with the length of 100 m before splicing
is shown in the Figure (4.3)

Figure 4.3: Graph of the initial transmission spectrum for the fiber F1 data obtained
using the optical spectrum analyzer.

As mentioned in the previous chapter, there are three absorption peaks, located at 950 nm,
1244 nm and 1383 nm, due to the presence of OH- functional groups. Once the reference
measurement for the transmission has been made (figure (4.3)), the splice was made by using
the equipment Fujikura model fsm 100p splicer. Different tests were carried out modifying
the characteristics of the splice, in particular the option STD (Standard) was modified, which
has the function of defining the power of the arc discharge with which the fibers will be fused,
and the fusion time held constant at 100 ms. Below are the experimental results for each of
the tests carried out for each of the cases from 0 STD to −500 STD.
(a) (b)

Figure 4.4: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.5: Results of the transmission spectrum measurements for before (blue line)
and after (orange line) splicing the two PCF sections.
(a) (b)

Figure 4.6: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.7: Results of the transmission spectrum measurements for before (blue line)
and after (orange line) splicing the two PCF sections.
(a) (b)

Figure 4.8: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.9: Results of the transmission spectrum measurements for before (blue line)
and after (orange line) splicing the two PCF sections.
(a) (b)

Figure 4.10: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.11: Results of the transmission spectrum measurements for before (blue
line) and after (orange line) splicing the two PCF sections.
(a) (b)

Figure 4.12: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.13: Results of the transmission spectrum measurements for before (blue
line) and after (orange line) splicing the two PCF sections.
(a) (b)

Figure 4.14: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.15: Results of the transmission spectrum measurements for before (blue
line) and after (orange line) splicing the two PCF sections.
(a) (b)

Figure 4.16: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.17: Results of the transmission spectrum measurements for before (blue
line) and after (orange line) splicing the two PCF sections.

It is possible to observe that by reducing the value of STD to −500 keeping the download
time constant (100 ms), a response quite similar to that obtained under the configuration
where the splicing is not yet performed is obtained. Said response has an average difference
of ≤ 1.3 dB, which is a low enough loss to be within the tolerance range, and therefore it
was decided that it was possible to use this configuration to make the connection between
PCF and SMF. For the case of splicing between SMF-28 and PCF, the following result was
obtained
(a) (b)

Figure 4.18: Photographs taken from the display of the splicing machine before (a)
and after splicing (b).

Figure 4.19: Results of the transmission spectrum measurements for before (blue
line) and after (orange line) splicing the two PCF sections.

As expected, the splice produced between conventional fiber and fabricated PCF is especially
stable under the configuration employed. Therefore, if this is the case, it is possible to splice
the PCF with conventional fiber so that the measurement can be performed remotely using
conventional fiber, which is generally more resistant than the manufactured PCF. Finally,
the results of the transmissions after splicing the fibers are presented, there the behavior of
the transmission is observed as the STD parameter was modified in the splicing equipment.
Figure 4.20: Results of the transmission spectrum measurements for before (blue
line) and after (orange line) splicing the two PCF sections.

4.2 PCF tapering

We tapered the fabricated PCF fibers to improve their evanescent fields. The tapering of
optical fibers consists of gradually reducing the diameter of a section of the fiber along its
length. In other words, the diameter of the fiber tapers in a progressive and controlled
manner, generating a conical shape as shown in the figure (4.21). As light travels along
the tapered fiber, the reduced diameter causes a change in light distribution, which in turn
affects the interaction of light with the materials surrounding the fiber.
Figure 4.21: Schematic representation of a tapered PCF.

To carry out the tapering of the fabricated PCF, a Vytran 3400 glass processor was used,
which uses a filament which is heated above 2500◦ in a uniform and stable manner. At the
initial, the investigated PCF with the length of about 1 m was taken. Then a PCF coating
as stripped out in the middle of the PCF for a length of about 3.5 cm. Once the stretch of
fiber is exposed, it is placed on a holder similar to those used in the splicing process and the
equipment is prepared to carry out the tapering. In the first place, a tension calibration is
carried out, applying around 20 g of tension to the fiber, in such a way that the fiber remains
taut so that it does not break when the tapering is carried out. This is possible by using
software-controlled movements of the movable bases where the fiber is positioned. As shown
in figure 4.21, there are control variables for weight loss such as waist length, transition size,
and waist diameter. Through software included with the equipment, it is possible to request
these variables, which allows great flexibility to carry out different tests. Before starting
the tapering of the PCF, a tapering test was performed with SMF-28 single-mode fiber to
receive a taper with the following characteristics: waist length LW = 30 mm, transition size
LT = 8.25 mm and waist diameter 44 µm. The above in order to obtain a reference given
that on the one hand our PCF has similar dimensions as the SMF-28 (in terms of the outside
diameter), and on the other hand the use of this equipment for PCFs is not common. But
the outside diameter (125 µm) of the SMF-28 fiber mentioned above are almost the same
one that the PCF reported in this work. So we used the SMF-28 tapering at the initial.

In order to perform the tapering of the SMF-28 fiber, the power that would be used to heat
the graphite filament was calibrated. It was determined that the power necessary to obtain
a needed waist diameter (the thinnest part of the taper) of about 44 µm [55] is around
50 − 54 watts and the translation speed was initially defined at 2.5 units/s, said speed comes
in units not known and only determined as a/s where a unit is the minimum movement
that can be made when moving. When carrying out several tests with SMF-28, it was
determined that an appropriate speed to obtain a waist diameter equal to that requested was
1 per second, for which reason, once the tests with SMF-28 were finished, the investigated
PCF was used. Around 15 tests were carried out with the PCF where each test, due to
the difference of the PCF and SMF-28 fiber, the values had to be modified to carry out
the tapering of the PCF fibers. When using a speed of 1 units/s some fibers broke before
finishing the tapering process and in other cases a fairly low waist diameter was obtained
(up to 31 µm), for which reason it was decided for increasing the speed up to 1.7 units/s,
where the best results were obtained in PCF tapering. Although the Vytran 3400 equipment

has an integrated microscope that allows measurements to be made at the time of tapering,
however, it was found that these measurements are not precise enough because they depend
on an estimate, which is made merely for conventional fibers. Therefore, it was decided to use
non-contact measurement equipment of the CIO drawing tower to compare the measurement
in the tapering equipment. Overall, the Vytran 3400 measurement was found to differ by
about ±5 µm. The measurement results of some of the tapers obtained for the investigated
PCF by the usage of the CIO drawing tower equipment are shown in the figure (4.22):

Figure 4.22: Profile of the PCF taper #9 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.4 units/s, time
0.1 ms, diameter of waist requested is 44 µm
Figure 4.23: Profile of the PCF taper #11 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.4 units/s, time
0.1ms, diameter of waist requested is 44.2 µm

Figure 4.24: Profile of the PCF taper #13 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.5 units/s, time
0.1 ms, diameter of waist requested is 42 µm.
Figure 4.25: Profile of the PCF taper #16 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.6 units/s, time
0.1 ms, diameter of waist requested is 45.5 µm

Figure 4.26: Profile of the PCF taper #21 measured by using the CIO drawing tower.
Parameters used to obtain the taper profile: Power 55 watts, Speed 1.7 units/s, time
0.1 ms, diameter of waist requested is 45 µm

It is possible to observe that there is a discrepancy between the diameter of the PCF taper
waist measured by the Vytran 3400 equipment and the diameter measured by the CIO
drawing tower equipment, as can be seen in each of the Figures (4.22), (4.23), (4.24), (4.25).
By modifying the working speed of the tapering equipment, it is possible to obtain a smoother
taper profile. This can be seen by comparing the results obtained from Figures (4.20) and
(4.26). In the first one, the speed of work has a value of 1.4 un/s and in the second, the
speed has a value of 1.7 un/s. The target taper waist diameter for our PCF was 44 µm
[55], and was easily obtained for a conventional SMF-28 fiber. However, the tests with our
PCF showed that due to the fundamental difference of its internal structure, the obtaining
of said taper waist diameter was not as precise when using the parameters of the tapering
equipment. With which, by modifying the speed parameter of the fiber tapering process, it
was possible to obtain a result quite close to the desired one of 44 µm.
Chapter 5

Refractive Index Sensing at the use of


the Fabricated PCF

5.1 Principle of Detection: Lossy Mode Resonance


It is possible to carry out structural modifications to the OFs in such a way that part of the
light that is guided in said structures can interact with the environment external to it. By
removing part of the coating from an OF and covering said part with a coating whose main
characteristic is to have a complex refractive index (IR), that is, a coating with losses (or
Lossy) in the OF, results in the generation of several attenuation bands in the transmission
spectrum, which is described as the coupling between OF modes and the lossy modes of the
film deposited on the fiber [56]. In general, the modes described above are the guided modes
in OF that have a complex effective index, also called long-range modes or Leaky modes
[32,33]. While these modes propagate through the fiber, they come across a section of said
fiber on which a layer with suitable characteristics has been deposited, that is, its refractive
index has a complex value, then certain attenuation bands are produced due to nature with
losses. A schematic representation of the situation is presented in Figure 5.1.
Depending on the type of material from which the special coating is created, the mecha-
nisms by which leaked mode resonances are generated can be classified into the following
categories: Surface plasmon resonance (SPR) and lossy mode resonance ( LMR). For the
SPR phenomenon to occur, it must be true that the real part of the dielectric constant of
the special coating (ϵ′ ) must have a negative value and of greater magnitude than its own
imaginary part, and it must also be true the above is compared to the permittivity of the
detection analyte (ϵs ). On the other hand, the LMR phenomenon occurs when the real part
of the dielectric constant of the special coating (ϵ′ ) is positive and greater than its imaginary
part. Below is a table where the permittivity conditions of the coating are expressed to
obtain one or another phenomenon.
Generally, the effect of the LMR phenomenon on the light transmission of an OF can be
observed as absorption peaks as a function of the wavelength, which is predetermined by
the characteristics of the coating, among which is its thickness, and additionally the refrac-
tive index of the medium outside the coating [58]. By changing any of the aforementioned
Figure 5.1: Schematic representation of the outcome produced by the LMR phe-
nomenon on an optical fiber

Type of Resonance Conditions (ϵ′ = ϵ′r + iϵ′i )


Lossy Mode Resonance (LMR) ϵ′r < 0, |ϵ′r | > |ϵ′i |, |ϵ′ | > |ϵs |
Surface Plasmon Resonance (SPR) ϵ′r > 0, |ϵ′r | > |ϵ′i |, |ϵ′ | > |ϵs |

Table 5.1: Conditions on the dielectric constant of the special coating that is used to
generate some of the resonance phenomena (SPR and LMR). The variables ϵ′r and ϵ′r
are the real and imaginary parts of the dielectric constant, respectively.

characteristics it is possible to observe a change in the absorption peaks in the transmission


as a function of the wavelength [59]. These changes provide us with information about the
variable of interest that we want to measure in the sensor and therefore represent one of the
most important aspects of the work.

5.2 Refractive Index Sensing Trials


To test the possibility of using the developed fibers for refractive index sensing, we im-
plemented a simple light transmission measuring set-up consisting of a white light source
(AQ-4305) and an optical spectrum analyzer (AQ-6315E). The investigated PCFs (F1) with
a full length of approximately 1 m were used. The protective coating on the pieces of the
investigated F1 fibers were deleted in the length of around 3.5 cm and then PCF tapers with
a taper waist diameter of about 44 µm [55] were fabricated. The waist regions of the tapered
fibers were sufficient in length (3 centimeters) to obtain the adiabatic criteria in accordance
with [62]. As it was shown earlier, we achieved also a properly optimized taper temperature
and pull rate. Under these conditions, the air channel structure of the PCFs after the taper-
ing process was preserved. 6 layers of polyethyleneimine (RI = 1.5290) coatings with a total
thickness t of about 1.2 µm were applied on a surface of the fiber taper cladding. Figure (5.2)
(a) (b)

Figure 5.2: Transmission spectra for the investigated F1 fiber taper # 21 with the
polyethyleneimine coating in air: (a) for wavelengths of 550-1650 nm, (b) for wave-
lengths of 1180-1210 nm.

presents transmission spectrum for the investigated F1 fiber taper #21 (See Figure (4.26))
with a taper waist diameter of 45 µm and with the mentioned above polyethyleneimine (PEI)
coating, when the sensing part of the fiber was placed in air.

We can see that high loss-bands exist in Figure (5.2a). It was shown in [63] that such ta-
pers can be used for creation of band-rejection filters. Theoretical estimates also made in
[55] showed that such fiber tapers can be used for refractive index or pressure fiber optic
sensors, and to sense a nanoscale adsorption layer of ammonia molecules deposited on a thin
specialized absorbing coating. The transmission spectra for the F1 fiber taper # 21, which
sensitive region was placed in water (RI = 1.333), and glycerin-water solutions with 20%,
40%, and 60% v/v of glycerin concentrations are presented in Figure (5.3). The glycerin
concentrations in water correspond to RI values at room temperature 1.362, 1.392 and 1.420,
respectively [64]. It is clear that a sensing effect can be seen in Figure (5.3)) for the in-
vestigated F1 fiber taper, which is appeared under action of refractive index changes in the
surrounding medium. The initial peak at 1188 nm was shifted to longer wavelengths when
RI of a surrounding medium was increased.
Figure 5.3: Transmission spectrum for the investigated F1 fiber taper # 21 with the
polyethyleneimine coating: in water, 20% glycerin, 40% glycerin, 60% glycerin for
wavelengths of 1180-1210 nm.

The taper was subjected also to different temperatures between 25 and 60◦ C. In that range
of temperatures, the investigated peak (1188 nm) did not suffer any shift. We did not carry
out measurements at higher temperatures, because a melting point for PEI is around 67◦ C.
Also, we did not make measurements below 25◦ C, because of technical limitations.
Chapter 6

Conclusions

In Conclusion, we were able to fabricate 2 Photonic crystal fibers with a periodic structure of
only 2 hexagonal air channel rings with the use of the Stack and Draw method. It was shown
with our measurements and theoretical simulations that for the fabricated fibers (F1 and
F2) are single-mode in the 550 to 1700 nm wavelength range as was expected by using the
Koshiba and Saitoh empirical model, and the fibers have sufficiently high optical losses, more
than 40dB/km as it can be seen in Figure (3.6). The values of d and ∧ in each of these fibers
were crucial to determine which of the two fibers was going to be more suitable for the sensing
application performance. This is because the values of d and ∧ also determine quantity d/∧
which at the same time determines propagation characteristics of the Fiber. That quantity
is greater for F2 therefore makes F1 Fiber more suitable for sensing application as can be
seen in the attenuation spectrum in Figure (3.6). The latter causes that the attenuation of
the F2 fiber be lower than F1 fiber which proves that it is possible to change the attenuation
of the fundamental mode PCF in our fibers by varying d/∧.

The theoretical modeling by the use of the empirical model showed that the fiber F1 was
a better option to be selected for sensing applications, and thus was subjected to a series
of tests which include splicing and tapering. During the mentioned tests we made enough
samples for comparison in order to obtain the best fit for splicing and tapering of the fiber F1,
these results are presented in chapter 4. During the splicing tests we obtained a good enough
regime of work for the PCF-PCF, SMF-PCF, SMF-PCF-SMF splices as showed in Figure
(4.22). In this regime we obtained losses caused by the splicing of the fiber for about 1 dB
in comparison with the measured transmission spectra of the PCF F1 alone. As for the PCF
tapering tests, we obtained taper waists of about 44 µm with lengths of 3 cm, we achieved
also a properly optimized taper temperature and pull rate. Under these conditions, the air
hole structure of the PCFs after the tapering process was preserved, and thus reaching the
adiabatic criteria. Out of the 21 tapers that were made, we implemented the tapered fiber
#21 (which can be seen in the Figure (4.26)) because of its waist diameter and the smooth
of its waist was the best compared with the other tapers.

Finally, to test the possibility of using the developed fibers for refractive index sensing, we
implemented a simple light transmission measuring set-up consisting of a white light source
(AQ-4305) and an optical spectrum analyzer (AQ-6315E). 6 layers of polyethyleneimine (PEI)
with a total thickness t of about 1.2 µm were applied on a surface of the fiber taper cladding.
The transmission spectrum, when the sensing part of the fiber was placed in air is showed
in Figure (5.2) which shows high loss-bands demonstrating the LMR phenomena on the
PCF. Measurements with glycerin-water solutions with 20%, 40%, and 60% v/v of glycerin
concentrations were carried out in order to test the fiber sensor. It is seen a sensing effect in
Figure (5.3)) for the investigated F1 fiber taper, which is appeared under action of refractive
index changes in the surrounding medium. The initial peak at 1188 nm was shifted to longer
wavelengths when RI of a surrounding medium was increased. These measurements then
proof the viability of the fabricated PCF sensor.
Bibliography

[1] D. Hutt, K. SNELL, and P. BÉLANGER, Alexander Graham Bell’s PHOTOPHONE,


Optics & Photonics News, Vol. 4, 20-25, (1993).

[2] J. M. Senior , Optical Fiber Communications Principles and Practice, Third Edition.
England: Pearson Education Limited, 2009

[3] K. Swarnjeet & K. Malik, Review Of Optical Fiber Communication


System-Introduction And Applications, Webology, Vol. 18, 5. Available:
https://www.webology.org/abstract.php?id=1500

[4] J. Hecht, City of Light: The Story of Fiber Optics, New York: OXFORD UNIVERSITY
PRESS, 2004

[5] T, Maiman, Stimulated Optical Radiation in Ruby. Nature , Vol. 187, 493–494, 1960.

[6] Kapany NS. Fiber optics. Sci. Am. 203, 72-81, 1960

[7] R. Bates, Optical Switching and Networking Handbook : McGraw-Hill, 2001

[8] K. Kao & G. Hockham, Dielectric-fibre surface waveguides for optical frequencies, IEE
Proc. J. Opt. electron., vol. 113(3), 191–198, 1965.

[9] A. R. Goodwin, J. F. Peters, M. Pion and W. O. Bourne, ‘GaAs lasers with consistently
low degradation rates at room temperature’, Appl. Phys. Lett., 30(2), pp. 110–113, 1977

[10] N. Kapany, Y. Chigusa, Ultra-low-loss (0.1484 dB/km) pure silica core fibre and exten-
sion of transmission distance, Electron. Lett. 38 (20) (2002) 116 1169.
[11] R. Maurer Low-Loss Optical Fiber for Telecommunications, Corning Glass Works, New
York, assignee. Patent 3,785,716. 15 Jan. 1974.

[12] E. Yablonovitch, Inhibited Spontaneous Emission in Solid-State Physics and Electronics,


Phys. Rev. Lett. 58, 2059, 1987

[13] S. John, Strong Localization of Photons in Certain Disordered Dielectric Superlattices,


Phys. Rev. Lett. 58, 2486, 1987

[14] E. Yablonovitch and T. J. Gmitter, Photonic Band Structure: The Face-Centered-Cubic


Case, Phys. Rev. Lett. 63, 1950, 1989

[15] P. St J. Russell, Photonic crystal fibers, Science 299, 358–362, 2003

[16] J. C. Knight, T. A. Birks, P. St J. Russell, and D. M. Atkin, Pure Silica Single-Mode


Fiber with Hexagonal Photonic Crystal Cladding, Opt. Soc. of America, San Jose, Cali-
fornia, 1996

[17] A. M. Pinto & M López-Amo, Photonic Crystal Fibers for Sensing Applications, Journal
of sensors 598178, 21, 2012

[18] P. R. Bigot & P. Roy, Fibres a cristal photonique: 10 ans d’existence et un vaste champ
d’applications, Images de la physique 71–80, 2007

[19] S. A. Cerqueira, Recent progress and novel applications of photonic crystal fibers, Reports
on Progress in Physics, vol. 73, 2, 2010

[20] C. Nylander, B. Liedberg, T. Lind, Gas detection by means of surface plasmon resonance,
Sensors and Actuators, vol. 3, 79-88, 1982

[21] Y. Wang, et al., Fiber-optic surface plasmon resonance sensor with multi-alternating
metal layers for biological measurement,Photonic Sens, vol. 3, 202–207, 2013

[22] C. Justino, A. Duarte, & T. Rocha-Santos, Recent Progress in Biosensors for Environ-
mental Monitoring: A Review, Photonic Sens, vol. 17, 12, 2017
[23] C. Ribaut, et al., Small biomolecule immunosensing with plasmonic optical fiber grating
sensor, Photonic Sens, vol. 77, 315-322, 2016

[24] Tombelli, S., Minunni, M. & Mascini M., Analytical applications of aptamers, Biosens.
Bioelectron., vol. 20, 2424–2434, 205

[25] S. Hinman, K. S. McKeating, Q. Cheng, Surface Plasmon Resonance: Material and


Interface Design for Universal Accessibility, Anal Chem., vol. 90, 19-39, 2018

[26] O. Fuentes, et al. Simultaneous Generation ofSurface Plasmon and Lossy ModeReso-
nances in the Same PlanarPlatform, Sensors, vol. 22, 1505, 2022

[27] I. Del Villar, et al. Optical sensors based on lossy-mode resonances., Sens. Actuators B
Chem. vol. 240, 174-185, 2017

[28] A. Ozcariz, C. Ruiz-Zamarreño, F.J. Arregui, A Comprehensive Review: Materials for


the fabrication of optical fiber refractometers based on lossy mode resonance, Sensors,
vol. 20(7), 1-23, 2020

[29] V.P. Minkovich, V. Kir’yanov, A.B. Sotsky, L.I Sotskaya, Large-mode-area holey fibers
with a few air channels in cladding: modeling and experimental investigation of the
modal properties., J. Opt. Soc. Am. B, vol. 21, 1161–1169, 2004

[30] T.P White, R. C. McPhedran, C. M. De Sterke and M. Steel Confinement losses in


microstructured optical fibers. Opt.Lett., vol. 26, 1660–1662, 2001

[31] M. Salazar, DISEÑO, FABRICACIÓN Y CARACTERIZACIÓN DE UNA SERIE DE


FIBRAS DE CRISTAL FOTÓNICO ALTAMENTE NO-LINEALES PARA APLICA-
CIONES EN ÓPTICA CUÁNTICA, Master’s dissertation, CIO, León Guanajuato,
2019.

[32] B. E. A. Saleh, M. C. Teich Optical Fibers Fundamentals of Photonics, Third Edition,


Jhon Willey & Sons, inc., United States, 2019, Pag. 391-430.

[33] D. Marcuse, Loss analysis of single-mode fiber splices, Bell Syst. tech. j. , vol. 56 703-718,
1977
[34] H. Murata Handbook of Optical Fibers and Cables, Second Edition, CRC Press, 1996.

[35] M. Ding, D. Fan, Y. Luo, and GD. Peng, Basics of Optical FiberMeasurements, Second
Edition, CRC Press, 1996.

[36] E. G. Neuman, Single-Mode Fibers fundamentals, Berlin Heidelberg, Springer, 1988.

[37] A. Mendez & T.F Morse, Specialty Optical Fibers Handbook, Elsevier Academic Press,
2007.

[38] A. Ghatak, K. Thyagarajan, An Introduction to Fiber Optics, New Delhi, Cambridge


University Press, 1998.

[39] K. Okamoto, Fundamentals of Optical Waveguides, Third Edition, Ibaraki, Japan, El-
sevier Academic Press, 2022.

[40] E. Yablonovitch Photonic band structure: observation of an energy gap for light in 3-D
periodic dielectric structures, in Annual Meeting Optical Society of America, 1988, New
Orleans

[41] E. Yablonovitch, Photonic band-gap structures, J. Opt. Soc. Am. B, vol. 10 283-295,
1993

[42] T. A. D. Birks et al., The analogy between photonic crystal fibers and step index fibers,
in OFC/IOOC, 1999, San Diego

[43] P. St.J. Russell, Photonic-Crystal Fibers: A Historical Account, J. Lightwave Technol.,


vol. 24 4729-4749, 2006

[44] J. C. Knight, T. A. Birks, P. St. J. Russell & D. M. Atkin, All-silica single-mode optical
fiber with photonic crystal cladding, Opt. Lett. . 21 1547-1549, 1996

[45] M. Koshiba & K. Saitoh Applicability of classical optical fiber theories to holey fibers,
Opt. Lett., vol. 29 1739-1741, 2004
[46] S. Obayya, B. Rahman & K. Grattan Accurate finite element modal solution of photonic
crystal fibres, Opt. Lett., vol. 152 241 – 246, 2005

[47] M. Koshiba, Full-vector analysis of photonic crystal fibers using the finite element
method, IEICE Trans. Electron., vol. 85 881 – 888, 2002

[48] K. Saitoh & M. Koshiba, Full-vectorial imaginary-distance beam propagation method


based on finite element scheme: Application to photonic crystal fibers, IEEE J. Quantum
Electron., vol. 38 927-933, 2002

[49] M. D. Nielsen & N. Mortensen Photonic crystal fiber design based on the V-parameter,
Opt. Express, vol. 11 2762–2768, 2003

[50] N.A. Mortensen, J.R. Folkenberg, M.D. Nielsen, & K.P. Hansen, Modal cutoff and the
V parameter in photonic crystal fibers, Opt. Lett., vol. 28 1879-1881, 2003

[51] C. Porras, Design, Fabrication and characterization of Nonlinear Photonic Crystal Fibers
for Near Infrared Wavelength Supercontinuum Light Sources, Master’s dissertation, CIO,
León Guanajuato, 2017

[52] M. Pereira, Design, Fabrication and Investigation of Special Microstructured Fibers,


Ph.D Dissertation, CIO, León Guanajuato, 2016.

[53] J. Hecht, City of Light: The Story of Fiber Optics, New York: OXFORD UNIVERSITY
PRESS, 2004

[54] N. Kapany, Y. Chigusa, Ultra-low-loss (0.1484 dB/km) pure silica core fibre and exten-
sion of transmission distance, Electron. Lett. 38 (20) (2002) 116 1169.

[55] M. Salazar, V.P. Minkovich, B. Sotsky, A. Shilov, L. Sotskaya, E. Chudakov, Lossy mode
resonances in photonic crystal fibers. J. Eur. Opt. Soc.- Rapid Publ., 24, 1-12 2021

[56] F.J. Arregui et al. Fiber-optic lossy mode resonance sensors. Procedia Eng., 87, 3–8
2014
[57] I. Del Villar et al. Lossy mode resonance generation with indium-tin-oxidecoated optical
fibers for sensing applications. J. Light. Technol., 28, 111–117 2010

[58] Q. Wang, W.M. Zhao A comprehensive review of lossy mode resonance-based fiber optic
sensors. Micromachines, 100, 47-67 2018

[59] Vikas, S.K. Mishra, A.K. Mishra, P. Saccomandi, R.K. Verma, Recent Advances in Lossy
Mode Resonance-Based Fiber Optic Sensors: A Review. Opt Lasers Eng, 13, 1921 2022

[60] A. Ozcariz, C.R. Zamarreno, F.J. Arregui, A comprehensive review: Materials for the
fabrication of optical fiber refractometers based on lossy mode resonance. Sensors, 20,
1972 2020

[61] E.I. Golant, A.B. Pashkovskii, K.M. Golant, Lossy mode resonance in an etched-out
optical fiber taper covered by a thin ITO layer. Appl. Opt., 59, 9254–9258 2020

[62] J.D. Love, W.M. Henry, W.J. Stewart, R.J. Black, S. Lacroix, and F. Gonthier, Tapered
single-mode fibres and devices. 1. Adiabacity criteria. IEE J, 138, 343-354 1991

[63] G.A. Cárdenas-Sevilla, D. Monzón-Hernández, V.P. Minkovich, Demonstration of an all-


fiber band-ejection filter based on a tapered photonic crystal fiber. Appl. Phys. Express,
5, 1-3 2012

[64] P.R. Cooper, Refractive index measurements of liquids used in conjunction with optical
fibers. Appl. Opt., 22, 3070-3072 1983

You might also like