You are on page 1of 133

UNIVERSITÉ JEAN MONNET OF SAINT-ETIENNE (FRANCE)

and

UNIVERSITÁ DEGLI STUDI OF PALERMO (ITALY)

Cotutelle Ph.D. Thesis


Giusy Origlio

Properties and Radiation Response


of Optical Fibers:
Role of Dopants

TUTORS:
Prof. Youcef Ouerdane
Prof. Marco Cannas

Ph.D. RAPPORTEURS:
Prof. Roberto Boscaino
Prof. Linard Skuja
To my new and to my old family
Contents

Contents 1

Introduction 1

I State of the art 3

1 The silica optical fibers 5

1.1 General structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.1.1 Light propagation in step-index optical fibers . . . . . . . . . . . . . . 6

1.1.2 Single-Mode and multi-mode optical fibers . . . . . . . . . . . . . . . . 8

1.1.3 Dispersion and losses in fibers . . . . . . . . . . . . . . . . . . . . . . . 11

1.1.4 Fiber Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2 Dopants in optical fibers 17

2.1 Germanium doped optical fibers . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.2 Fluorine doped optical fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3 Phosphorus doped optical fibers . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4 The optical fibers under irradiation exposure . . . . . . . . . . . . . . . . . . . 22

3 Point defects in optical fibers 25

3.1 Intrinsic point-defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.1.1 Oxygen Deficient Centers . . . . . . . . . . . . . . . . . . . . . . . . . 27


3.1.2 Oxygen associated hole centers . . . . . . . . . . . . . . . . . . . . . . 29

3.2 Extrinsic point-defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.2.1 Ge-related defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.2.2 P-related defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

II Materials and methods 37

4 The canonical samples 39

4.1 Tested optical preforms and fibers . . . . . . . . . . . . . . . . . . . . . . . . . 40

5 Experimental set-ups 47

5.1 Irradiations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

5.1.1 UV laser irradiations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

5.1.2 γ-ray and X-10 keV irradiations . . . . . . . . . . . . . . . . . . . . . . 48

5.2 Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5.3 Photoluminescence and Raman spectroscopy . . . . . . . . . . . . . . . . . . . 50

5.3.1 Photoluminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

5.3.2 Stationary and time resolved luminescence setup . . . . . . . . . . . . . 52

5.3.3 Photoluminescence under synchrotron radiation excitation . . . . . . . 54

5.3.4 Raman measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

5.3.5 Confocal Micro-spectroscopy setup . . . . . . . . . . . . . . . . . . . . 55

5.4 Electron Paramagnetic Resonance measurements . . . . . . . . . . . . . . . . . 57

III Ge-doped fibers and preforms 59

6 Measurements on non-irradiated samples 61

6.1 Discussion: the drawing effect . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

7 Effects of the UV and X-ray irradiation 69


7.1 EPR results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

7.2 Optical absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

7.3 Discussion: radiation effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

7.3.1 Localization of defect species . . . . . . . . . . . . . . . . . . . . . . . . 75

7.3.2 Generation processes of GECs defects . . . . . . . . . . . . . . . . . . . 77

7.3.3 The drawing effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

IV Influence of further dopants: fluorine and phosphorus 83

8 F-doped fibers and preforms 85

8.1 Raman results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

8.2 EPR measurements on irradiated samples . . . . . . . . . . . . . . . . . . . . 89

8.2.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

8.2.2 Discussion: generation of E0 centers . . . . . . . . . . . . . . . . . . . . 91

9 P-doped fibers and preforms 93

9.1 Optical activity of P-related point defects . . . . . . . . . . . . . . . . . . . . . 93

9.1.1 Absorption and photoluminescence analysis . . . . . . . . . . . . . . . 93

9.2 Discussion: luminescent P-defects structure . . . . . . . . . . . . . . . . . . . . 101

9.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

Conclusions 107

List of related papers 109

List of communications to congresses 111

Bibliography 113
Introduction

Today the circulatory system that sustains our communication society is made up by optical
fibers. These low-loss glass fibers facilitate worldwide broadband communication such as the
Internet. Light travels in thin guides of glass, and it carries almost all of the telephony and
data traffic in every direction. Text, music, images and video can be transferred around the
world in a fraction of second.

If unraveled, all of the glass fibers that wind around the globe would turn into a single
thread over one billion kilometers long-which is sufficient to encircle the globe more than 25000
times-and which is still increasing by thousands of kilometers every hour. Global communi-
cation, and in particular internet and long-distance telephony, is now based mainly on optical
fiber technology.

The main benefit resulting from the use of optical waves with respect to radio waves is
the high frequencies that allow high data transmission rate. Today, it is possible to transmit
several terabits per second in a single fiber and that represents an improvement by a factor
of one million to what could be obtained fifty years ago with radio signal transmission. The
number of optical fiber cables being installed all over the world is increasing rapidly. Fiber
optics is also important for a huge number of other applications in medicine, laser technology
and sensors.

In order to be developed and manufactured, the optical fiber needed modern glass tech-
nology. Furthermore, a reliable light source was also needed and this was provided by semicon-
ductor technology. Finally, a clever network needed to be assembled and extended, consisting
of transistors, amplifiers, switches, transmitters and receivers, as well as other units, all work-
ing together. This telecommunications revolution was made possible by the work of thousands
of scientists and inventors from all around the world. Even if the optical fibers have been so
intensively investigated over the years, the interest of the scientific community is still alive:
in fact the Nobel Price in Physics 2009 was awarded to C. K. Kao, whose discoveries have
paved the way for optical fiber modern technology. In 1966, Kao understood that it was not
imperfections in the fiber thread that was the main responsible for losses, instead it was the
glass that had to be purified, because of the presence of defects. He admitted that this would
be feasible but very difficult. The goal was to manufacture glass of a transparency that had
2 Introduction

never been attained before.

Even in the glass fiber of the highest purity, the signal, however slightly, is reduced along
the way and needs reinforcement when it is transmitted over longer distances. This task
previously required electronics, while it is nowadays performed by optical amplifiers. This has
allowed to overcome the unnecessary losses that occur in the transformation of light to and
from electronic signals.

Furthermore, choosing which fiber to use is subject to so many different technical consid-
erations, communication needs and costs, that it is not possible to speak of only one single
kind of fiber. The fibers are based on a complex interplay between size, material properties,
and wavelengths of light.

Following the interest in the field, this Thesis deals with the experimental study of the
spectroscopic properties of three types of prototype preforms and associated fibers. The sam-
ples have been designed and fabricated to investigate the role of germanium (Ge), fluorine
(F) and phosphorus (P) doping elements on the fiber attenuation and eventually on the ra-
diation sensitivity of silica-based glasses. We characterized the behaviors of these canonical
samples before, during and after irradiation through several spectroscopic techniques, to ob-
tain global information (electron paramagnetic resonance) or spatially-resolved information
(confocal microscopy, absorption and luminescence on preform).

The Thesis is organized in four parts. Part I, comprising Chapters from 1 to 3, deals
with the general system of optical fiber communication providing an extensive overview of
the history, construction, operation, and benefits of optical fiber, with particular emphasis on
the importance of the doping procedure to enhance the fiber characteristics. An overview of
the main intrinsic and extrinsic defects in silica is also presented. Part II includes Chapters
4 and 5 and it is devoted to the description of the prototype samples and of the adopted
experimental techniques. Part III, including Chapters 6 and 7, reports on the experiments
on Ge-doped samples and their main results. The results concerning F-doped and P-doped
samples are reported and discussed in Part IV, comprising Chapters 8 and 9. Finally, the
most relevant "conclusions" are summarized. A "list of the scientific papers" comprising the
results presented in this Thesis and a few others on closely related topics are reported at the
and of the manuscript, together with a list of communications to congresses.
Part I

State of the art


Chapter 1

The silica optical fibers

Optical fibers lie at the very heart of modern society, providing the information superhighways
required within our global communication systems. Fiber-optic communication is based on the
principle that light in a glass medium can carry more information over longer distances than
electrical signals can do in a copper or coaxial medium or radio frequencies through a wireless
medium. The purity of today’s glass fiber, combined with improved system electronics, enables
fiber to transmit digitized light signals hundreds of kilometers without amplification. With
few transmission losses, low interference, and high bandwidth potential, optical fiber is an
almost ideal transmission medium. Thanks to high transmission speed, low attenuation and
interference and the large bandwidth, optical fibers represent at now the major progress in
data transfer.

The purpose of this chapter is to introduce the main basic fiber features starting with a
description of some general properties of silica optical fibers and showing the reasons of the
massive scientific investment in optical fiber telecommunications technology nowadays.

1.1 General structure

An optical fiber is essentially a dielectric waveguide used for information transfer. The fiber
communication is based on the principle that the light in a glassy medium can carry more
information and at longer wavelengths in comparison to the electrical signal transferred by
classical cables. The usual telecommunication optical fibers are made of two cylindrical parts:
the interior part is called core, the exterior one is the cladding (Figure 1.1).

Core and cladding have different refractive index: the exterior part has a smaller refrac-
tive index than the inner one, allowing light reflection according to the classical geometrical
optics laws. The two fiber portions are generally made of the same glassy material where the
6 1. The silica optical fibers

cladding
jacket n2

buffer

core
n1

Figure 1.1: Structure of a classical optical fiber for telecommunications.

refractive indexes are varied and accurately controlled during the fiber fabrication through
dopant incorporation in the silica-based matrix. A classical fiber for telecommunication has
an exterior diameter of about 125 µm, while the core diameter varies from few µm to 60 µm,
depending on the network requirements a . Such a fiber could prove to be mechanically fragile,
so it is necessary to cover it with two protective coatings: a first vitreous buffer and a poly-
meric exterior jacket.

1.1.1 Light propagation in step-index optical fibers

The most basic function of a fiber is to guide light, i. e., to keep light concentrated over longer
propagation distances despite the natural tendency of light beams to diverge, and possibly
even under conditions of strong bending.

A very important concept in fiber optics is that of waveguide modes. These are field
configurations which maintain their intensity profile during propagation, apart from possible
power losses. Of highest interest are usually the guided modes, i. e. those modes which have
significant intensity only in or near the core. Depending on the fiber design and the optical
wavelength, some number of guided modes may exist, or only a single one, or even no guided
mode at all. A fiber with only one guided mode is called a single-mode fiber, and multi-mode
fibers support several guided modes (section 1.1.2).
a
We can also find multimode fibers with cores of 100, 200 µm, or more. Such fibers are used for peculiar
applications (like sensors) and they are not routinely used for telecommunication networks.
1.1. General structure 7

The propagation mechanism inside an optical fiber can be approximatively described by


geometrical optics principia. The used ray picture cannot be applied to fibers with a small
core or a small refractive index contrast between core and cladding: the approximation scale
improves with reduction in λ/r ratio, where r is the optical fiber core and λ is the light
wavelength propagating inside. The reason is that wave effects occur: a real beam has some
finite width, and the incident and reflected wave interfere with each other. Furthermore, the
optical field somewhat extends beyond the core/cladding interface. Therefore, the ray picture
is only a rough approximation for strongly guiding large core fibers, while a wave analysis
through Maxwell equations is required for the more general case.

The research activity described in this PhD Thesis is related to multi-mode optical fibers,
for which the ray optics approximation is effective: therefore only a description of light prop-
agation on fiber by means a geometrical approach is here supplied. The loss of generality
that such choice implies is partially balanced by the immediate physical interpretation of the
results and the simple visualization of the propagation processes.

It is common to explain the guiding effect as a result of total internal reflection. A light
beam S approaching the separation interface between two transparent and homogenous media,
with refractive index n1 and n2 respectively, is partially reflected and partially refracted. If
θ1 is the incident angle against the normal direction, the refractive beam will propagate in n2
medium in accordance with the Snell law:

n1 sin(θ1 ) = n2 sin(θ2 ) (1.1)

The same mechanism is involved in the optical fiber: the core refractive index n1 is greater
than the cladding one n2 and the refracted angle is greater than the incident one (θ1 <θ2 ).
For incidence angle values greater than θ1 =θc =arcsin(n2 /n1 ), called critical angle, there is
no refracted angle. This is the total internal reflection phenomenon which is at the basis of
the optical fibers working. All the rays propagating inside the core with an angle θ > θc ,
will be totally reflected and trapped inside the fiber. Typical values for the optical fibers are
n2 = 1.475, n1 = 1.5; θc = 79.5◦ .

It is possible to define an acceptance cone (Figure 1.2) containing all the rays propagating
inside the core through total internal reflection. The cone vertex is the center of the entry
face of the fiber and the vertex angle is called acceptance angle θA .
It is also possible to get a measure of the coupling efficiency between the source and the fiber
defining the so called numerical aperture (N.A.) defined byb :
q
N.A. = sin(θA ) = n21 − n22 (1.2)

b
If the core radius a is much larger than the wavelength λ, a geometrical-optics description for the propa-
gation of light is valid. However, when a is in the order of λ, a wave-propagation theory is needed.
8 1. The silica optical fibers

Ray outside the


acceptance cone

Ray lost into


θA the cladding

Acceptance cone

Figure 1.2: Acceptance cone in an classical optical fiber

1.1.2 Single-Mode and multi-mode optical fibers

As anticipated in section 1.1.1, based on the number of modes propagating through the
fiber, there are multi-mode and single mode fibers [1].

Figure 1.3: Optical fiber sizes

Multi-mode fibers

Multi-mode fiber was the first type of fiber to be commercialized. It has a much larger
core than single-mode fiber, allowing hundreds of modes of light to propagate through the
fiber simultaneously. Multi-mode fibers routinely used for telecommunication networks have
core sizes of 50 to 62.5 µm in diameter, while the overall diameter is about 125 to 200 µm
1.1. General structure 9

(Figure 1.3). Based on the refractive index profile we have two types of fibers: (a) Step index
fiber (b) Graded index fiber.

(a) Step index fiber : in the step index fiber, the refractive index of the core is uniform
throughout and undergoes an abrupt or step change at the core cladding boundary.
The light rays propagating through the fiber are in the form of meridional rays which
will cross the fiber axis during every reflection at the core cladding boundary and are
propagating in a zig-zag manner as shown in Figure 1.4a. When light is launched into a
multi-mode fiber, multiple guided modes can be excited, and at the fiber exit, there is
an intensity profile which arises from the interference of light in all these modes. In this
kind of fiber there is a considerable modal dispersion (section 1.1.3): even rays with the
same wavelength but emitted at a different incident angles (lower than the acceptance
angle) propagate with the same speed into the fiber but across different length paths.
They will arrive at the fiber end at distinct times, producing a temporal broadening of
the transmitted pulse.

(b) Graded index fiber : in the graded index fiber, the core refractive index is made to vary
in a parabolic manner so that the maximum value of refractive index is at the center of
the core. The light rays propagating through it are in the form of skew rays or helical
rays which will not cross the fiber axis at any time and are propagating around the fiber
axis in a helical or spiral way as shown in Figure 1.4b. In the case of multi-mode graded
index fiber, signal distortion is very low because of self-focusing effects. Here the light
rays travel at different speeds in different paths of the fiber because of the parabolic
variation of refractive index of the core. As a result, light rays near the outer edge travel
faster than the light rays near the center. In fact, the rays are continuously refocused as
they travel down the fiber and almost all of them reach the exit end of the fiber at the
same time due to the helical path of the light propagation.

Multi-mode fibers are strongly required when light from a source with poor spatial co-
herence has to be transported. As an example, the output of a high power diode bar contains
thousands of modes, and requires a correspondingly large number of fiber modes. Additionally,
the larger core diameter facilitates the use of lower-cost optical transmitters and connectors.

In most applications, the standard multi-mode graded index optical fibers have significant
performance advantages over conventional copper-based systems: they are very useful when
multimodal transmission is needed for relative long distances. However, performance require-
ments and cost restraints may prohibit the use of these fibers in certain applications. First
of all it is very complex and expensive to realize a graded-index fiber in which the refractive
index varies continuously during all the fabrication process. Sometimes fiber manufacturers
modify standard material composition and structural design to meet these additional require-
ments. The intent of each change is to increase performance and reduce cost. For instance it
10 1. The silica optical fibers

Figure 1.4: Different propagation modes in (a): multi-mode step index, (b): multi-mode graded-index
and (c): single-mode step-index fibers.

is possible to obtain the optimal characteristics of a graded-index fiber in the so called multi-
step index fiber. As its name indicates, the structure, showed in Figure 1.5, uses multiple step
indexes which approximate the parabolic curve of the refractive index profile. Although the

Figure 1.5: Refractive index profile in multi-step index fiber.

basic principle is the same as that of step index fiber because the index of refraction changes
in multiple steps, the locus of the light is shifted toward the center at the same time. In any
case, with enough number of steps, differences from a graded-index become small and they
could be neglected. So it is possible to reconcile the advantage of a little modal dispersion
with a more easy fiber production at reasonable prices.

Single-mode Fibers
1.1. General structure 11

In a single mode fiber, only one mode can propagate through its core (Figure 1.4c). The
single mode fiber has a smaller core diameter (10 µm, Figure 1.3) and the difference between
the refractive indices of the core and the cladding is very small. Its fabrication procedure
could be very difficult and the launching of light into single mode fibers is also hard. The
advantages of single mode optical fibers lie in the very low transmission loss and dispersion or
degradation, thus resulting very useful in long distance communication.

1.1.3 Dispersion and losses in fibers

Dispersion in the fiber means the broadening of the signal pulse width due to dependence of
the refractive index of the material of the fiber on the wavelength of the carrier. If we send
digitized signal pulses in the form of square pulses, they are converted into broadened gaussian
pulses due to dispersion. The dispersion leads to the distortion or degradation of the signal
quality at the output end due to overlapping of the pulses. There are two kinds of dispersion
mechanisms in the fiber: intramodal dispersion and intermodal dispersion.
The first one arises due to the dispersive properties of the optical fiber material (material
dispersion) and the guidance effects of the optical fiber (waveguide dispersion). Further it
increases with the increase in spectral width of the optical source.
Intermodal dispersion or multi-mode dispersion arises due to the variation of group velocity for
each mode at a single frequency. Different modes arrive at the exit end of the fiber at different
times. So there is multi-mode dispersion and hence there is broadening of the signal pulses.
The multi-mode step index fibers exhibit a large value of dispersion due to the enormous
amount of multi-mode dispersion which gives the greatest pulse broadening. At the same time
the multi-mode graded index fiber exhibits an overall dispersion which is 100 times lesser than
the multi-mode step index fiber’s dispersion. This is due to the shaping of the refractive index
profile in a parabolic manner. In the case of single mode step index fibers, they have only
intramodal dispersion.

Attenuation is the reduction of signal strength or light power over the length of the light-
carrying medium. Fiber attenuation is measured in decibels per kilometer (dB/km) and it is
a function of wavelength as shown in Figure 1.6.
Attenuation is caused by several different factors, but primarily diffusion (Raileigh scattering)
and absorption. It can be classified into two types: intrinsic and extrinsic losses generated by
several mechanisms:

• Tail of infrared (IR) absorption by Si-O coupling that it is present at higher wavelengths
around 1.4 µm to 1.6 µm.

• Tail of ultraviolet (UV) absorption due to electron transitions and present at lower
wavelengths near 0.8 µm. This produces a loss of 0.3 dB/km.
12 1. The silica optical fibers

Figure 1.6: Spectral attenuation of a silica optical fiber.

• Rayleigh scattering (Figure 1.7) originates from microscopic irregularities in the glass

Figure 1.7: Illustration of Rayleigh scattering effect.

structure; it is inversely proportional to λ4 and in many cases it can be expressed asc :


 4
0.85
αR [dB/km] = 1.7 (1.3)
λ[µm]
It produces high losses mainly in the ultraviolet region. In the wavelength region around
0.8 µm to 1 µm, it gives a loss of 0.6 dB/km.

• Absorption: conversion process of electromagnetic wave energy into other forms (i. e.
lattice vibration). Intrinsic silica glass absorption occurs in both ultraviolet and infrared
c
Equation 1.3 is sample-dependent: actually Rayleigh losses depend on the core composition. The formula
predicts 0.15 dB/km at 1.57 µm, while lower Rayleigh losses of 0.12 dB/km have been reported by Nagayama
et al. [2].
1.1. General structure 13

bands, in particular infrared absorption tail causes attenuation for the wavelengths longer
than 1.6 µm. Further attenuation is caused by light absorbed by residual species, such
as metals or OH ions, within the fiber core and inner cladding. In particular OH causes
the water peak region on the attenuation curve, typically around 1.4 µm. The removal of
OH ions is of primary interest to fiber manufacturers as this water peak has a broadening
effect and contributes to attenuation loss for nearby wavelengths. Figure 1.8 shows the
spectral attenuation of different material fibers.

Figure 1.8: Spectral attenuation of different material fibers.

For silica fiber, the lowest losses of about 0.12 dB/km can be obtained in the region around
1.55 µm [2]: at longer wavelengths, the attenuation increases. An optical signal transmitted
through fiber, could travel more than 100 km without regeneration or amplification.

Other attenuation mechanisms are due to macroscopic bends, occurring when installing
fibers, microscopic bends, due to local distortions of the fiber geometry, and nonlinear scatter-
ing.

Optical power propagating in a fiber decreases exponentially with distance:


P (z) = P0 exp(−α0 z) (1.4)
where P is the optical signal power and α0 is the attenuation coefficient [1/km].
Using a logarithmic scale we obtain:
log P (z) = −αz/10dB + log P0 (1.5)
where α is the logarithmic attenuation coefficient measured in [dB/km].

Overall optical fibers offer superior performances over other transmission media because
they combine high bandwidth with low attenuation. These properties allow the transmission
of signals over longer distances while using fewer regenerators or amplifiers, thus reducing cost
and improving signal reliability.
14 1. The silica optical fibers

1.1.4 Fiber Fabrication

The manufacture of an optical fiber takes place into two steps: the preform fabrication and
the drawing process. Preform is a cylinder of silica composition from 10 mm to some cm in
diameter and from 60 to 120 cm lengthd . It consists of a core surrounded by a cladding with
a desired refractive-index profile; in other words, this is a desired optical fiber, but on a much
larger scale. The main reason a preform is prepared is to have a drawable material that is clean,
low in OH concentration, low in metallic-ion contaminants, and inexpensive. Many techniques
have been developed to prepare these preforms. Some common commercially used methods
are Outside Vapor-Deposition (OVD), Modified Chemical Vapor Deposition (MCVD), Vapor
Phase Axial Deposition (AVD), and Plasma Chemical Vapor Deposition (PCVD) and Plasma
Modified Chemical Vapor Deposition (PMCVD) [3]. All these methods are based on thermal
chemical vapor reaction in which two gases, SiCl4 and O2 , are mixed at a high temperature
(>800 ◦ C) to produce silicon dioxide (SiO2 ):
SiCl4 + O2 → SiO2 + 2Cl2 (1.6)
Silicon dioxide, or pure silica, is usually obtained in the form of small particles (about 0.1 µm)
called soot. This soot is deposited on the target rod or tube layer upon layer and it forms a
homogeneous transparent cladding material. To change the value of the cladding’s refractive
index, some dopants are used. For example, fluorine (F) is used to decrease the cladding’s
refractive index in a depressed-cladding configuration. The soot for the core material is made
by mixing several gases which results in a mixture of SiO2 and of the core dopant. The degree
of doping is controlled by changing the amount of dopant gas added to the mixture. Since
deposition is made by the application of silica layers, the manufacturer can control the exact
amount of dopant added to each layer, thus controlling the refractive-index profile.

The different preparation methods differ mainly by the way the soot is deposited. The
preforms studied in this PhD thesis were all made by MCVD process which provided a simple
and straightforward means of manufacturing high-quality optical fibers.
This method was developed by Bell Laboratories [4]. The soot is deposited on internal wall of
the tube (Figure 1.9) and then vitrified by the traversing burner to provide a thin glass layer.
The procedure is repeated many times as the cladding and core layers are formed. When
the deposition is finished, the temperature of the burner is increased (≈1700 ◦ C) to collapse
the tube into a solid preform [5]. The entire process is highly automated and all process
parameters are precisely controlled.

Optical fibers are obtained by drawing from the preform at high temperature (≈2000 ◦ C).
The drawing process must be integrated with the coating process to avoid contamination of
fiber surface. These processes are shown schematically in Figure 1.10. The tip of the preform is
d
It remains difficult to have an idea of the maximum preform diameter as this is confidential for the fiber
manufacturers
1.1. General structure 15

Figure 1.9: Deposition by modified chemical vapor deposition (MCVD) process.

Figure 1.10: Optical Fiber Drawing Process

heated in a furnace to a molten state. Formed molten gob falls down under the force of gravity
while shrinking in diameter into a proper diameter strand. It is controlled continuously during
the drawing process. Diameter drift cannot exceed 0.1%. The strand is threaded through a
series coating applicators immediately after drawing. Liquid prepolymer coatings are cured
by thermal or ultraviolet apparatus. Dual coating, soft inner and hard outer, is needed to
protect against impact and crushing forces in both manufacturing process and installation.
16 1. The silica optical fibers

The fiber with coatings is pulled down and wound on a winding drum. The drawing process
must take place in controlled atmosphere, because air pollution influences fiber attenuation.
Both stages of fiber manufacturing are fully automated and are performed in a clean, climate-
controlled room. Obviously, the manufacturers use high-precision measuring equipment to
automatically control each step of the fabrication process. For example, preform analyzers
measure the critical characteristics of the optical-fiber preform. Also, specific measurement
systems control fiber geometry, the refractive-index profile, and the coating geometry.
Chapter 2

Dopants in optical fibers

As reported in section 1.1, fundamental condition to having light propagation in optical fibers
is the different refractive index between core and cladding. To realize an index variation in
a−SiO2 , dopants are usually added in the glass matrix.

Depending on the use and characteristics of the optical fibers, several elements can be
added to modify the fiber characteristics. GeO2 and P2 O5 are dopants commonly used for
doping the core region, raising the refractive index. On the other hand B2 O3 or F are dopants
chosen for the cladding region that in turn lower the refractive index (see Figure 2.1). Several

Figure 2.1: Refractive index as a function of dopant materials and their concentration (from ref. [6]).

dopants can be added in more special fibers to functionalize the glass, as rare-earth ions
(erbium, ytterbium [7]) for fiber-based amplifiers, or fluorine to improving the fiber radiation
hardness (see section 2.4).

The nature of the elements (impurities or dopants) contained in fibers deeply influences
18 2. Dopants in optical fibers

the optical properties of the fibers themselves: dopants can modify the fiber hardness under
radiation exposure or simply influence the drawing process.
The dopants need to have the following characteristics [8]:

• It is of high purity and easily available.

• It is easy to liquify.

• It differs from the transition metal in vapor pressure.

• It is easy to vitrify with silica and gives a proper refractive index.

• After being vitrified, its coefficient of thermal expansion is nearly equal to that of SiO2 .

• When vitrified, it has stable properties.

The following sections are devoted to a review of the influence of three particular dopants
often used in optical fiber technology: germanium, fluorine and phosphorus.

2.1 Germanium doped optical fibers

The addition of germanium in the silica matrix, disguised as GeO2 , allows to increase
the refractive index of the glass. This property is often used for the elaboration of the optical
fiber core and Ge has been the first traditional dopant used in fiber. Ge-presence does not
affect the fiber losses in the telecommunication windows (1300-1500 nm) (save for the increase
of Rayleigh scattering due to density fluctuations), but it can produce the apparition of new
energy levels within the silica band gap, thus leading to detrimental losses of part of the
transmitted signals into the fibers (see section 3.2.1 for details).

The scientific interest for germanosilicate glass increased even more after the experimental
discovery of the property of photosensitivity of this material. Photosensitivity of a medium is
defined as its capacity to have its refractive index permanently changed by a modification of
its physical or chemical properties through UV light exposure. Photosensitivity is a complex
phenomenon and it is not well understood yet because of the influence of many parameters:
fiber composition, fabrication process, operation wavelength and even light sources. Photo-
sensitivity was first observed in 1978 by Hill et al. [9] at the communication Research Centre
in Canada. The experiment consisted of injecting light from a single frequency Argon laser
(514 nm) into the core of a Ge-doped silica fiber. Hill observed that a fraction of the input
power was reflected by the fiber itself and this phenomenon was attributed to the formation
of a permanent index grating. Progress in optical fiber photosensitivity research developed
rapidly after the discovery of the possibility to write Fiber Bragg Gratings (FBG) into the
2.1. Germanium doped optical fibers 19

fiber illuminating the core from the fiber’s side with the interference pattern of two beams of
coherent UV radiation [10], as shown in Figure 2.2. Spectroscopic studies of Ge-doped fibers

Figure 2.2: Inscription of a Fiber Bragg Grating on the core of an optical fiber

before and after intense UV exposure have been interpreted as pointing to a color center model
for photosensitivity, in which a Ge-related defect optical activity (see section 3.2.1) at the ex-
posure wavelength (242 nm) is bleached [10, 11]. Nevertheless many studies provided some
additional clues about the microscopic mechanisms of photosensitivity, such as laser-induced
densification [12, 13, 14, 15].

Apart from the photoinduced change of the isotopic refractive index, it was discovered
in 1985 by Parent et al. that photoinduced birefringence could also be written into fibers by
polarized radiation [16]. Additionally, in 1986 Osterberg et al. discovered that a prolonged
exposure of an optical fiber to 1064 nm light from a Nd:YAG laser results in the generation of
second harmonic light at 532 nm [17,18]: optical nonlinearity was discovered in germanosilicate
glasses.

These features have a deep impact in several applications and therefore they stimulate
the strong interest in the study of Ge-doped amorphous silica (a−SiO2 ), to understand the
microscopic mechanisms at the basis of the properties of the material with the aim to control
and enhance them. Usually the largest part of scientific investigation on Ge-related glasses
for optical fibers consists in the direct study of bulk samples and the subsequent transfer of
information to the fibers [19, 20, 21]. However, this approach cannot take into account the
peculiarities implied in the fiber preparation procedure, such as the drawing process, which
can generate precursors and influence the defect generation [22] and the necessity of direct
studies on fibers samples strongly emerges.
20 2. Dopants in optical fibers

2.2 Fluorine doped optical fibers

Fluorine is an important dopant in optical fiber technology. In contrast to most of other


dopants, F decreases the refractive index of silica glass [23] (Figure 2.1). This property is of
great practical importance for designing optical fibers with an undoped high-purity silica core.
Such fibers exhibit the best performance in the UV and IR spectral regions and have a better
durability in environments with an increased level of ionizing radiation. So fluorine-doped silica
is a promising key material in optical fiber technology directed to applications requiring high
and stable transmission over a broad spectral range from infrared to ultraviolet. Indeed, recent
studies have shown that radiation toughness of silica samples is achieved by incorporating
Si − F groups (Figure 2.3) whose positive effect is assumed to be the reduction of defect
precursors [24, 25], such as strained bonds (≡ Si − O) from which is likely generated the pair
of silicon dangling (≡ Si• ) and oxygen dangling (≡ Si−O• ), where (≡) and (• ) indicate bonds
with three oxygen atoms and an unpaired electron, respectively (see section 3.1). Fluorine is a

Figure 2.3: Schematic illustration of fluorine incorporation in a−SiO2 matrix (from ref. [24]).

silica network modifier, because it considerably affects the viscosity and softening temperature
of silica glass [26,27]. Recently, thanks to high transparency in the vacuum ultraviolet (VUV)
range without any increase of optical defects [28,29,30] (Figure 2.4), the F-doping has received
a large attention through the application of silica glass as an optical material for projection
photolithography at 157 nm of F2 excimer laser [25, 31]. It is well known that the shape of
the fundamental absorption edge in the exponential (Urbach) region can yield information
on the disorder effects [32]. Skuja et al. [30] demonstrate that fluorine doping affects Urbach
VUV absorption edge by increasing its steepness. It is evident from Figure 2.4 that when F
was doped to 1 wt%, the transmittance at 157 nm increased to ∼80 % in comparison to non
doped glass. This gain in VUV transparency by F-doping is mainly due to a reduction of
concentration of strained bonds in the silica network. Because the Si − F bond is stronger
than the Si − O bond, a monotonic increase in the optical bandgap of fluorine doped SiO2
would be expected with increasing the F content [25, 31].

The replacement of a single bridging oxygen atom with a terminal fluorine, results in the
formation of SiO3\2 F tetrahedra, that produces a depolymerization of the silicate network and
2.3. Phosphorus doped optical fibers 21

Figure 2.4: VUV absorption spectra of SiO2 :F glasses as a function of F content. From ref. [25].

consequently the lowering the viscosity of silica [33].

The positive effects of F-doping seems to have at least 4 mechanisms: (a) by quenching of
distinct color centers absorbing in the edge region, (b) by reducing of the structural disorder
by breaking up the strained bonds in glass network, (c) by increasing of the band gap due to
the higher energy of Si-F bond as compared to Si-O bond, (d) by reducing the glass viscosity,
thus allowing to achieve more easily a lower fictive temperature of the glass. Actually, the
feasibility to exploit these properties in the fabrication of F-doped silica fibers is conditioned by
several queries including the dependence on F-concentration also on the basis of its influence
in modulating the silica refractive index and the role of F after drawing silica preform. Despite
the great importance assumed by silica glasses doped with F, their structure is not yet well
characterized. It is known, as above mentioned, that fluorine is incorporated in the glass matrix
essentially as Si-F bonds, taking the place of a bonding oxygen. For all glasses containing
≥1 wt% fluorine, a small fraction of the fluorine is bonded to silicon atoms containing four
bridging oxygen atoms, resulting in fivefold coordinated silicon of the type SiO4\2 F [27].

2.3 Phosphorus doped optical fibers

Phosphorus-doped a−SiO2 is a material of fundamental importance in fiber optics com-


munications and in microelectronics. First of all, the addition of P2 O5 allows to improve the
refractive index of the glass (Figure 2.1), for this reason phosphorus doping is often used in
optical fibers to achieve an optimal refractive index profile [6]. Phosphorus is indeed used in
optical fibers to ameliorate the internal glass structure [34] thanks to its ability in modifying
the viscosity of the core and cladding regions [34]. Phosphate glasses are potentially good
ultraviolet (UV) transmitting materials allowing the fabrication of thin glass films for appli-
22 2. Dopants in optical fibers

cation in microlithography and laser systems [34]. They are also good materials for highly
effective optical amplifiers, especially via co-doping with rare-earth elements [7,35], and phos-
phosilicate glasses are promising candidates as radiation sensors due to their closely linear
response to radiation dose [36]. It has been reported that strong photosensitive properties
can be induced in P-doped silica by hydrogen loading or high temperature treatment in a
hydrogen-oxygen flame [37, 38]. A thorough understanding of the microscopic arrangements
by which P impurities are incorporated in silica, as well as of the properties of the result-
ing P-related point defects, would be useful to optimize the performance of P-doped SiO2 in
applications (see section 3.2.2).

Due to the fact that phosphorus is first of all used as co-dopant in fiber technology, it
is particularly difficult to separate its contribution in optical fiber attenuation in the UV-
visible region [39]. Only a few papers in literature have investigated this issue, so that little
information is available at the moment on P-related point defects and on their generation and
transformation mechanisms [40, 34, 41]. The situation is very different from the case of other
dopants, like germanium for which, as seen in section 2.1, a vast amount of knowledge deriving
from experimental and theoretical work has been accumulated.

2.4 The optical fibers under irradiation exposure

The appearance of new radiative environments integrating silica components, such as optical
fibers [42, 10], requires their immunization under ionizing radiation. In particular two impor-
tant applications of optical fibers in the nuclear industry are related to plasma diagnostics
in fusion reactors [43] and transmission of signals from inaccessible parts of nuclear instal-
lations [44]: in this field the relevant doses are above 1 MGy [45]. When optical fibers are
subjected to radiation, whether it consists of high energy light, X-rays, γ-rays, neutrons or
high energy cosmic particles, their optical properties change due to the interaction of the ra-
diation in the fiber core and in the cladding material. The main effects result from electronic
processes: electrons are excited to leave their normal (bound) position, changing the physical
and chemical properties of the silica glass. The most obvious among the optical effects is the
radiation induced optical attenuation, that depends essentially from wavelength: this is in
general a not desirable effect because it causes degradation of the performance of the optical
fiber systems. Moreover, it is also possible to use the fiber response under radiation exposure
as a detector for radiation.

At the present time, there are several applications for the optical fibers under radiation
environments. As an example, fiber diagnostic and imaging are new interesting fields for the
development of the optical fibers technology under radiation exposure. Historically, the first
interest in fiber response under ionizing radiation comes from the military sphere. Moreover,
2.4. The optical fibers under irradiation exposure 23

due to the confidential nature of these information, very few literature data exist on this subject
[46]. In contrast the fiber applications in space environment or civil nuclear environment have
been largely investigated.

The interaction of the radiation with the fiber material is a complex process with quite
a number of dependencies on parameters related to the fibre fabrication process, operating
environment and radiation type. Exposure to radiation can induce stable alterations of the
material [47, 48], often related to point defects generation and conversion processes [21] (see
section 3).
24 2. Dopants in optical fibers
Chapter 3

Point defects in optical fibers

In 1966 Charles K. Kaoa and A. Hockham, two English researchers of the British Post Office,
found and demonstrated that the high-loss, till then observed in the existing optical fibers,
arose from impurities in the glass, rather than from an underlying problem with the technology
itself [49].

The presence of defects in optical fibers often causes the appearance of new energy levels
located inside the band gap of the dielectric [50, 51]. As a consequence, the glass absorbs a
more important part of the transmitted signal giving rise to an attenuation of the light guided
inside and consequently in a degradation of the fibers themselves.

The fiber radiation response depends on many intrinsic parameters: core and cladding
dopants, impurity content, strain [52, 22], which are generally not accessible for researchers.
Usually classical optical fibers for telecommunications are used in the IR, from 835 to 1600 nm,
where the optical transmission is the largest. Moreover new technological fields, like medical
application and plasma diagnostic, need the use of light guides in the visible and UV region
were optical transmission is affected by many losses [50, 53, 54, 55]. All these aspects moti-
vate the necessity to investigate the exact nature of point defects, checking their origin and
properties and so reducing degradation effects also in the UV-visible domain.

Point defects and their precursors in the amorphous silica network are introduced during
the fabrication process, through dopants and the interaction with ionizing radiation (high
energy, photons including UV laser irradiation, particles). Numerous publications are related
with defects in a−SiO2 [56, 57, 50, 58]. Most types of defects have optical absorption and
luminescence bands and could be detected by optical absorption (OA) in the visible, UV,
or infrared spectral range, Raman and photoluminescence (PL) spectroscopies. Detailed in-
formation and identification on the subset of paramagnetic defects is obtained by electron
paramagnetic resonance spectroscopy (EPR). Often, the combined use of different techniques
a
Nobel price in physics 2009
26 3. Point defects in optical fibers

allows to infer information not available by examining separately the results of single observa-
tions. The formation of paramagnetic point-defects in silica glass has been studied from two
points of view: transformation of diamagnetic precursors (also point defects) and the breaking
of intrinsic Si-O bonds.

Defects can be distinguished in intrinsic, when they are due to a variation of the basic
silica elements (silicon or oxygen) and extrinsic, if they are related to presence of impurities
in the silica matrix (H, Ge, P, etc.). Extrinsic defects due to the presence of impurities (Cl, H
and so on) are always present in variable concentrations in the material. On the other hand,
how above explained (see section 2), selected impurities can be deliberately added by doping
to induce many useful properties [59, 51, 60].

To provide a background for the presentation of the results, the following sections of this
chapter are devoted to review in more detail the current understanding of defects in a−SiO2 ,
particularly with regard to the generation and conversion of defects related to the optical
fibers.

3.1 Intrinsic point-defects

Amorphous silica is the principal building material for glassy fiber waveguides. The structure
of the glass network and point defects has been the subject of extensive studies through a large
variety of experimental techniques and theoretical modelling [61, 62, 63, 64]. An illustrative
picture of an amorphous silicon dioxide network is shown in Figure 3.1. The SiO2 network is

Si
O

Figure 3.1: structure of the amorphous silica, with Si atoms in grey and O atoms in black. The angle
α define the spatial configuration of two connected tethraedra.

built with SiO4 tetrahedra joined at the corners so that each Si-atom is bound to four O-atoms
and each O atom is the bridge between two Si-atoms. The angles defining the relative spatial
orientation of each pair of connected tethraedra are statistically distributed between 120◦ and
180◦ [65,66]. This description of the microscopic structure of amorphous silica is known as the
3.1. Intrinsic point-defects 27

Continuous Random Network (CRN) model, and is mainly based upon the evidences coming
from X-ray and neutron diffraction [62, 61, 64, 67]. The structural order in glass can generally
be divided into different stages or ranges [62]. The first stage is the tetrahedron structural unit
SiO4 followed by the interconnection of adjacent units. A third stage is the network topology
for describing the intermediate range order in shortest path ring structures. Finally, the long
range density fluctuations over several tens of are the fourth stage of structural order. A
point defect in the intrinsically disordered structure of silica can be defined as any deviation
from the perfect glass structure defined by the CRN model, provided that it is localized in a
region whose dimensions are comparable to the interatomic distance [62].

In the following the characterization of the main intrinsic defects are summarized.

3.1.1 Oxygen Deficient Centers

Oxygen deficient centers (ODC) are formed when an oxygen is missing or removed for instance
by irradiation from its Si-bonding position.

The silicon dangling bond, or E 0 center, is the most widely investigated oxygen deficient
defect in a−SiO2 . It consist of a silicon atom with six electrons in three pairs and one unpaired
electron: ≡Si• , the symbol (≡) represents three bonds to oxygen atoms, (• ) represents one
unpaired electron (Figure 3.2 (B,C,D)). E 0 defect was observed for the first time in 1956 by

Figure 3.2: Oxygen deficient centers in silica (from [50]). (A): Relaxed oxygen vacancy (ODC(I)
center). (B,C): silicon dangling bond (E 0 center) relaxed into the plane of the neighboiing oxygens
(B) or relaxed towards neighboring bridging oxygen atom (C). (D): Surface-type SiE 0 center. (E):
Twofotdcoordinated Si atom (ODC(II) center).

R.A.Weeks in neutron-irradiated α-quartz (E10 ) and in silica [68] (Figure 3.2(B)). Thus far,
at least four different types of E 0 centers have been observed in silica: Eγ0 , Eδ0 [69, 70, 71, 72],
Eα0 [73], Eβ0 [69,59] (Figure 3.2(B,C,D)). They differ from each other in the second coordination
environment around the respective silicon atom and they are distinguishable, at least in prin-
ciple, either by their spectroscopic properties or on the basis of their generation mechanism.
Additionally there exists a surface type E 0 center [74] due to isolated silicon dangling bonds
(Figure 3.2(D)). E 0 center is found almost in every specimen exposed to radiation, but it can
28 3. Point defects in optical fibers

be also formed by the fibre drawing process [59]. The E 0 family of centers are paramagnetic
and give rise to strong EPR signals. The EPR spectrum of the E 0 center consists of a single
main resonance line (Figure 3.3) and of four hyperfine doublets with splitting of ∼40 mT
(strong hyperfine), ∼0.8 mT and ∼0.9 mT (weak hyperfine) and ∼0.05 mT (very weak hy-
perfine) [58, 75, 76]. An optical absorption band centered at 5.8 eV with FWHM=0.8 eV and

Figure 3.3: X-band electron paramagnetic resonance spectrum of E 0 center (from [75]).

oscillator strength f=0.14, well correlates with the growth of E 0 EPR signature [77, 78] and it
is usually ascribed to E 0 , even if the nature of the optical transition involved is still contro-
versial [60, 79]. The formation efficiency of E 0 defects strongly depends both on the content
of the hydroxyl radicals (OH) in the glass and and on the irradiation energy. It was shown
by Hanafusa [80] and Hibino [81] that E 0 defects also exist in non-irradiated optical fibers.
The strong tensions during the preform drawing process seem to be at the origin of the defect
formation. Also the drawing conditions, like temperature or drawing speed, influence the E 0
concentration in optical fibers: in particular it was shown that the E 0 concentration growth
as a function of the drawing temperature, follows the Arrhenius law [80, 81].

The neutral oxygen vacancy (ODC(I)) [82] consists of a bond between two Si atoms and
it is indicated as ≡ Si − Si ≡ (Figure 3.2(A)). This diamagnetic ODC is electrically neutral
and intrinsic to oxygen deficient silica since it disappears by oxidation. It gives rise to strong
optical absorption bands around 7.6 eV [82, 83]. By hydrogen loading of silica samples it was
shown that the oxygen vacancy converts itself to Si-H groups according to the reaction [50]:

≡ Si − Si ≡ +H2 −→≡ Si − H + H − Si ≡ (3.1)

The twofold coordinated Si (ODC(II)) [84] consists of a Si coordinated with two O atoms
and denoted by = Si•• where (•• ) represents two paired electrons in the same orbital (Fig-
ure 3.2(E)). This defect is also called silicon lone pair center or divalent Si. It shows a
relatively weak absorption band, called B2 band, with peak at 5.03 eV and FWHM 0.4 eV
and two photoluminescence emissions at 4.3 eV and 2.7 eV to singlet-singlet and triplet-singlet
transitions occurring in the same defect [50].
3.1. Intrinsic point-defects 29

3.1.2 Oxygen associated hole centers

The oxygen dangling bond or Non Bridging Oxygen Hole Center (NBOHC) (≡Si−O• ) [50, 85]
is shown in Figure 3.4(A). It is detectable by its characteristic EPR signal, as well as by

Figure 3.4: Oxygen excess-related color centers in a−SiO2 (from [50]). (A): Non Bridging Oxygen Hole
Center (NBOHC). (B): Peroxy radical (POR). (C): Peroxy bridge. (D): Interstitial oxygen molecule.
(E):Interstitial ozone molecule.

its optical activity, consisting in three absorption bands at 2.0 eV, 4.8 eV and 6.4-6.8 eV,
which excite a photoluminescence emission peaked at 1.9 eV. All these absorbtion bands, and
in particular the intense bands at 4.8 eV and 6.8 eV, make NBOHC the defect that more
influences the transmission of silica in the UV and VUV spectral ranges. There are several
formation mechanisms for the NBOHC. In optical fibers these centers could be created during
the drawing process [86] and their concentration grows in particular with the O2 [86] flux
and the drawing tension [87]. In bulk silica the NBOHC are usually created after energetic
radiation exposure (X, γ, UV) [85, 88].

The peroxy radical (POR), independently from its formation mechanism, consists in a
silicon atom linked to an oxygen molecule: ≡ Si − O − O• [58] (Figure3.4(B)). It has an
unpaired electron delocalized on two oxygen atoms that are not equivalent from a chemical
point of view: the electron spends 75% of the time on the more distant oxygen atom from the
silicon one. Several formation mechanisms were proposed in literature to explain the formation
of these defects. Like NBOHC and E 0 , the peroxy radical can be induced during the drawing
process and it can be revealed by EPR measurements at low temperature [86]. The following
reaction was proposed for the POR formation:
Si − O − O − Si −→ Si − O − O• + Si + e− (3.2)
Some authors supposed that the same mechanism could be responsible for the POR formation
under ionizing radiation [59, 89].

Other oxygen excess related defects are the peroxy bridge, the interstitial oxygen molecule
and the interstitial ozone molecules (Figure3.4(C, D, E)). The presence of O2 [50] in silica has
been inferred from out-gassing experiments, from reaction with H2 forming Si-OH groups and
from conversion from E 0 centers to POR centers.
30 3. Point defects in optical fibers

Finally, the self-trapped hole (STH) may be the first defect to form under the influence
of ionizing radiations. Its principal characteristic is the capture of a hole on a 2p orbital from
a doubly linked oxygen atom [90]. Two different STH species were identified: the STH1 and
the STH2 [91], consisting on a self trapped hole on one or two oxygen atoms respectively.

3.2 Extrinsic point-defects

Among the impurities present in silica fibers, hydrogen, germanium, phosphore and fluorine
are the most diffuse. How above explained, Ge, P and F are very important dopants in fiber
technologies. On the other hand several other impurities are often integrated in the silica
matrix whether for the difficulty in eliminated them during the preparation procedure or for
improving the fiber properties.

3.2.1 Ge-related defects

Germanium may be arranged within a−SiO2 in many different configurations, each of which
constitutes a specific point defect. Since Ge and Si are isoelectronic elements, it is qualitatively
expected that many Ge-related point defects are structurally identical to Si-related centers
apart from the substitution of Si with Ge [92]. Starting from the comparison between a Ge-
doped silica glass and a pure silica glass, it is possible to show that defects related to germanium
are predominant on the intrinsic ones [19]. This property implies an UV absorption from two
to three order od magnitude more intense in germanosilicate glasses, even before irradiation
exposure [93], as compared to pure silica.

Actually defects in germanium doped silica, and than in germanosilicate optical fibers,
are the main subject of many experimental and theoretical works in order to know the origin
of the photosensitivity in these glasses (see section 2.1). Several researchers have shown that
a contribution to the photosensitivity is due to the variation of the UV optical absorption
spectra associated with the so called Ge-lone pair center (GLPC ) [55,54,21]: a dicoordinate
germanium atom with a lone pairs (= Ge•• ) [94, 95]. This defect is characterized by an OA
band at 5.1 eV related to the optical transition S0 →S1 [54], from the ground state to the first
excited singlet state. It has been suggested that the bleaching after radiation exposure (UV
laser, γ-rays) of this OA band, often referred as B2β , associated with the growth of several new
absorption signals (Figure 3.5), is at the basis of the permanent glass refractive index change.
This observation clearly suggests that the defect responsible for the B2β band are converted
by UV radiation to other centers [55, 97, 96, 94, 95]. GLPC defect is also characterized by two
photoluminescence (PL) bands at 4.2 and 3.1 eV, related to the transitions from the excited
electronic states of singlet (S1 ) and triplet (T1 ), respectively, to the ground state (S0 ) [92, 98].
3.2. Extrinsic point-defects 31

Figure 3.5: Difference absorption spectrum, showing the bleaching of the 5.1 eV band and the growth
of two components at 4.5 eV and 5.8 eV. From Fujimaki et al. [96].

Figure 3.6: Evolution of PL signal associated with GLPC after UV pulsed laser irradiation. The
inset shows the OA spectra acquired before (dashed line) and after (solid line) the irradiation. Figure
adapted from Cannas and Origlio [21].

Then the energetic level scheme, pictured in Figure 3.7 and associated with GLPCs, consists
of a ground singlet S0 and the excited S1 and T1 states. The radiative decay channels from S1
and T1 are described by the rates KS and KT respectively, while the ISC process linking S1
and T1 is characterized by KISC b . Though the determination of the GLPC spatial distribution
in optical fibers is crucial to probe the silica refractive index variation, convenient experiments
have not been performed yet, the main obstacle being the small fiber dimensions. Experimental
literature data on the defect spatial distribution in fiber, exist mainly for elements of intrinsic
nature [99, 100].
b
other non-radiative channels can be neglected [92].
32 3. Point defects in optical fibers

S1 KISC

KS T1
OA 5.2 eV
PL 4.2 eV PL 3.1 eV
~ 10-8 s

S0
O

Figure 3.7: General scheme of GLPC diamagnetic defect. Solid arrows indicate the radiative transition
in absorption and luminescence. Dashed arrows indicate the Inter System Crossing (ISC) non-radiative
transition.

The most common Ge-related paramagnetic defects that are detected by EPR in irradiated
Ge-doped a−SiO2 are the GeE 0 center and the Germanium Electron Centers (GECs) Ge(1)
and Ge(2) (Figure 3.8).

(a) (b) (c)

Figure 3.8: Microscopic structures proposed by Neustrev [19] as models for (a): Ge(1), (b):Ge(2) and
(c): GeE0 defects.

The microscopic structures of GeE 0 and Ge(1) have been unambiguously identified by
EPR studies, further supported by theoretical calculations. The GeE 0 , which is observed also
in pure GeO2 , is structurally identical to the E 0 center apart from substitution of Si with Ge
(≡ Ge• ) [101, 102, 55] (Figure 3.8 (c)). This center was put in relationship with an absorption
band at 6.2 eV-6.4 eV [19, 103].
The Ge(1) consists in an electron trapped at the site of a substitutional 4-fold coordinated
Ge precursor (GeO•4 ) [19, 104] (Figure 3.8 (a)). An absorption band at 4.4 eV-4.6 eV has been
attributed to this center [19, 19, 105, 106].
Finally the structure of the defect responsible of the latter EPR signal, the Ge(2) center,
is still debated. Indeed, its structural model was first ascribed to a trapped electron center
at the site of a GeO4 unit, such as the Ge(1), on the basis of the similarities of their 73 Ge
hyperfine coupling constants, differing from Ge(1) for the number of Ge nearest neighbors
3.2. Extrinsic point-defects 33

ions [107]. According to this attribution, an absorption band at 5.8 eV was assigned to Ge(2)
center [54, 105]. However, subsequent studies, based on the defect annihilation, suggested an
alternative model for Ge(2): an ionized twofold coordinated Ge (= Ge• ) [19, 96]. Even the
circumstance that the g value of Ge(2) is smaller than 2.0023 does not permit its conclusive
assignment to a trapped electron center, because this line of reasoning is rigorously valid only
for very simple paramagnetic centers, and generally cannot be extended to point defects in
silica [108]. The EPR signals related to GeE0 and GECs centers are reported in Figure 3.9.

Figure 3.9: EPR signature of the GeE0 , Ge(1) and Ge(2) paramagnetic defects in germanosilicate
irradiated silica. From Fujimaki et al. [106].

3.2.2 P-related defects

Several literature papers are focused on the study and characterization of paramagnetic P-
related point defects generated by ionizing radiation [40, 109]. In a defect-free SiO2 glass each
oxygen would bridge two SiO4 tetrahedra. On the other hand, the ideal P2 O5 glass would be
characterized by one nonbridging and three bridging oxygens per phosphorus: any deviation
from this structure can be considered as a defect [40].

Most of the current understanding of P-related defects in SiO2 derives from electron
paramagnetic resonance (EPR) experiments on irradiated phosphosilicate glasses. Electron
paramagnetic resonance allowed to identify 4 main P-related paramagnetic point defects, re-
ferred to as P1, P2, P4 and Phosphorus Oxygen Hole Center (POHC ) centers [110, 40] that
were not observable before irradiation exposure [40]. Figure 3.10 shows the supposed struc-
tures for the above mentioned P-related defects, together with the precursors suggested by
Griscom et al. in ref. [40] for all these defect structures.
34 3. Point defects in optical fibers

Figure 3.10: Main phosphorus related paramagnetic defects induced by radiation in P-doped a−SiO2
silica. The supposed precursor structures are also showed (from ref. [40]).

In P4, P1 and P2, the unpaired electron is localized on the central P atom, bonded to
a different number of oxygen atoms, 2, 3, 4 respectively [40, 111, 112, 38, 109]. Hence, their
structure can be represented as [(O−)2 P • ]0 , [(O−)3 P • ]+ , and [(O−)2 P • (−O)2 ]0 respectivelyc .
The paramagnetic signal of POHC is ubiquitous in P2 O5 -containing glasses. In the simplest
model of this defect, the P atom is bonded to three bridging O atoms and to a fourth non-
bridging O which hosts the unpaired electron: [(O−)3 P − O• ]+ . However, this structure
(here referred to as l-POHC) has been argued to be stable only at low temperature, while
the room-temperature stable form of POHC (here referred to as r-POHC) was proposed to
feature an electron shared by two non-bridging oxygen atoms bonded to the same phosphorus
[(O−)2 P (−O)•2 ]0 [40]. l-POHC and r-POHC supposedly feature two slightly different EPR
signals. Figure 3.11 shows the EPR signature of the POHC center: arrows indicates the
metastable PHOC form. Figure 3.11 also shows in the central part of the EPR spectrum a
signal attributed by Griscom et al. [40] to the so called Si(E0 )(P), a species of Si(E0 ) center
with phosphorus next-nearest-neighbors.

After clarifying by EPR the microscopic structure of these defects, data obtained by
optical absorption (OA) studies of irradiated P-doped silica were interpreted by proposing
associations between some of the observed OA bands (Figure 3.12) and the paramagnetic
centers [40, 38]. In particular Griscom proposed the attributions presented in Figure 3.13 for
the OA bands of Figure 3.12, observed in the range 3÷6 eV. According to these assignments,
r-POHC centers absorb at 2.2, 2.5 and 5.3 eV, while l-POHC have an absorption band at
about 3.1 eV.

In contrast, much less is known about diamagnetic P-related centers in SiO2 . Based on
the results obtained by several independent experimental techniques, including Raman and
c
In P4, the P atom hosts an additional lone pair, not represented.
3.2. Extrinsic point-defects 35

Figure 3.11: EPR signature detected at low temperature (77 K) of the stable form of POHC defect.
Additional peaks indicated by the arrows are supposed to be due to the metastable POHC variant. In
the central part of the spectrum the lineshape related to Si(E0 )(P) is also visible (from ref. [40]).

Figure 3.12: Radiation induced optical absorption spectrum in a P-doped silica sample (from ref. [40]).

Infrared measurements, phosphosilicate glass is generally believed to consist of an intermixed


random network of [(O-)2 Si(-O)2 ]0 and [(O-)3 P=O]0 tetrahedra randomly bonded by sharing
O atoms, this being consistent with the fact that [(O-)3 P=O]0 is the basic building block of
pure stochiometric (P2 O5 ) phosphate glass [113, 114, 115, 116]. In this model each P atom is
bonded to 3 bridging O atoms and a single doubly-bond non-bridging O, and thus each site can
be argued to be a potential precursor for l-POHC via ionization of the non-bridging oxygen. In
contrast, r-POHC should be formed by ionization of a defective site [(O-)2 P(=O)2 ]− where the
P atom bonds two bridging oxygen atoms with single bonds and two more non-bridging oxygen
with double bonds. Finally, P4, P1, P2 centers are supposedly formed by irradiation via hole
or electron trapping on hypothetical diamagnetic precursor defects where P is 2-, 3-, and 4-fold
36 3. Point defects in optical fibers

Figure 3.13: Attribution of the main P-related absorption bands showed in Figure 3.12 in irradiated
phosphosilicate glasses. Also the average peak energy (E), the full with at half maximum (W ) and the
oscillator strength (f ) are reported. From [40].

coordinated respectively (Figure 3.10):d [(O-)2 P:]− , [(O-)3 P:]0 , [(O-)2 P(-O)2 ]+ [40,111,112,38].
In the intermixed random network model, also these sites should be considered as randomly
occurring point defects.

d
In the 2-fold coordinated precursor of P4 center, the P atom hosts an additional lone pair, not represented.
Part II

Materials and methods


Chapter 4

The canonical samples

This chapter is focused on the description of the particular specimens used to perform the
experiments discussed in the rest of the work. An important part of the interest in our approach
is based on the choice in the design of our samples that hereafter we will call canonical samples.
These samples have to be representative of commercial fibers that have already been tested
and will be used in future facilities [117]. They have also to offer an easier interpretation
of their responses thanks to their custom designs. Previously, two different studies used a
set of homemade samples to understand the influence of several process and composition
parameters on the radiation response of single-mode germanosilicate optical fibers at 1.3 and
1.55 µm [118, 22]. Due to the good knowledge of their sample characteristics, E.J. Friebele et
al. [118] were able to obtain statistically significant correlations between the γ-ray steady-state
Radiation Induced Attenuation (RIA) and some of the fabrication parameters. As an example,

they found that for doses of 2×103 rad at -35 C, the RIA level at the end of the irradiation
is correlated with the Ge content in the fiber core (for a Ge/F doped cladding). The second
study was devoted to the transient X-ray radiation response of Ge-doped fibers and showed

that, on lowering the standard preform deposition temperature from 2000 to 1600 C and
the drawing tension from 140 to 20 g, the induced losses slightly decrease wavelengths [119]
and the influence of various cladding co-dopants. However, these two studies were limited
by the difficulty to obtain the samples with the characteristics needed to get unambiguous
correlations. For the present work, we design the structures to overcome these difficulties.
First, a quantitative analysis of the influence of a dopant can only be achieved if all the
differently-doped glasses have been made with strictly identical processes. From a practical
point of view, due to the non-negligible influence of MCVD process parameters [119], this can
not be achieved by investigating several preforms and fibers. It must be done within a single
sample. Secondly, new spectroscopic techniques are now accessible, thus allowing to spatially
resolve the radiation-induced changes in the fiber with micrometer resolution. For example,
we show the efficiency of the confocal microscopy of luminescence (CML) to characterize the
40 4. The canonical samples

radiation-induced point defects in passive [100] or rare-earth doped [120] optical fibers. These
spatially-resolved techniques enable the characterization of new fiber designs (such as the
samples studied in this Thesis) that would have not been possible in the past.

4.1 Tested optical preforms and fibers

The purpose of this Thesis is to study the role of the dopants in the properties and in the
radiation response of the multi-mode optical fibers. To this aim the experiments were carried
out on three types of silica optical fiber and preform samples, doped with germanium, fluorine
and phosphorus respectively. These samples are prototype not commercialized yet: their future
use can be planned for specific fields, like space or nuclear power plants, where expositions
to high irradiation doses are expected. All preform and associated fiber samples were made
through the Modified Chemical Vapor Deposition process (see section 1.1.4) by iXFiber SAS
[121]. About 50 mm of each prototype preform has been kept for analysis whereas the other
part of the preform has been drawn to obtain several hundreds meters for each fiber. The
refractive index profiles for the three types of samples are presented in Figure 4.1. Standard
conditions of fiber manufacturing (preform deposition and drawing process) have been used
for these waveguides. To quantitatively investigate the influence of the dopants concentration,
the structure of each preform, and then of each fiber, has been designed with several steps
of concentration of one doping element, Ge or F or P, in the core. This particular sample
structure was thought and realized ad hoc with the purpose of studying the dopant influence,
thus maintaining the other fabrication parameters as fixed. The fiber and preform main
characteristics are listed in Table 4.1.

Using spatially-resolved techniques [100, 120, 122], we will then be able to study the fiber
or preform properties and the radiation response to the dopant concentration for samples with
strictly identical MCVD process parameters.

Dramatic importance in the sample fabrication had the choice of the dopant concentration
levels. Part of the dopant concentration values have been chosen to reproduce the classical
range of concentrations measured on commercial fibers (e.g.,from 2 to 12 wt.% for the Ge-
doped fibers). The other ones have been defined in relation with our ab initio calculations
conducted on a 108 atoms silica-based supercell [123].

At the fabrication stage, the obtained multi-step radial distribution of the dopant along
the fiber diameter can be roughly estimated through measurements of the fiber and preform
refractive- index profiles (Figure 4.1). A more accurate estimation of the concentration values
of the dopants is obtained by electron microprobe analysis (EMPA) which also allows to inspect
the impurities content inevitably present in the samples. Fiber and preforms are made up of
four cylindrical layers (core part, zones 1-4) of high pure synthetic silica differently doped,
Table 4.1: Parameters related to the canonical fiber and preform samples. ∆n refers to the refractive index change at λ=633 nm with respect to
a−SiO2 . Dopants and impurities average concentrations were evaluated by electron microprobe analysis.

(a)

Ge canonical samples P canonical samples


Zone <Ge> <Cl> ∆n Pref. <P> <Cl> ∆n Pref. Fiber
(wt%) (wt%) (×10−3 ) diam. (wt%) (wt%) (×10−3 ) diam. diam.
(mm) mm (µm)
Cladding 0 0.0 0.01 0.33 10.16 0.00 0.23 0.46 12.8 125.0
1 2.5 0.07 2.19 5.02 1.63 0.07 1.37 5.0 62.5
4.1. Tested optical preforms and fibers

2 4.5 0.10 4.33 4.24 3.17 0.06 2.58 4.2 52.8


Core 3 8.0 0.09 7.97 3.10 5.05 0.05 4.80 3.1 38.6
4 11 0.06 11.95 1.48 7.09 0.02 7.08 1.6 18.4

(b)

F canonical samples
Zone <F> <Cl> ∆n Pref. Fiber
−3
(wt%) (wt%) (×10 ) diam. diam.
(mm) (µm)
Coating 0 0.0 0.00 0.33 9.47 125.0
Cladding 1 2.5 0.08 -5.9 4.44 62.5
2 1.3 0.10 -4.1 3.98 52.8
Core 3 0.7 0.08 -2.1 3.10 38.6
4 0.2 0.00 -0.5 1.50 18.4
41
42 4. The canonical samples

1 2

1 0
G e
8

)
-3
∆n ( x 1 0
6

0 (a )
-4 -3 -2 -1 0 1 2 3 4

-1 F
)

-2
-3
∆n ( x 1 0

-3

-4

-5

-6 (b )
-3 -2 -1 0 1 2 3

5
P
)
-3

4
∆n ( x 1 0

0 (c )
-4 -3 -2 -1 0 1 2 3 4
R a d ia l d is ta n c e ( m m )

Figure 4.1: refractive index change (∆n) measured at λ=633 nm with respect to a−SiO2 in: (a)
germanium, (b) fluorine and (c) phosphorus canonical preform samples.

following a multiple step distribution. The layers were deposited in a tube of undoped fused
silica which forms the cladding (zone 0). The preform samples have a diameter of about 10 mm,
with a 5 mm inner doped region, and they were cut and polished into plates of approximately
1.5 mm thickness. The fiber/preform length ratio is about 4×103 and the fiber core diameter
is 62.5 µm.
4.1. Tested optical preforms and fibers 43

Ge-doped canonical samples

Ge-doping increases the refractive index of a−SiO2 , so germanium doping profile grows from
the boundaries to the center. The samples have been doped with several amounts of germanium

Figure 4.2: Microscopic preform view obtained with an optical microscope. Numbers from 0 to 4 refer
to the zones listed in Table 4.1 with different amounts of germanium.

from ∼2 wt.% in the exterior part (zone 1) and up to ∼11 wt.% in the inner-center part (zone
4), as shown by the microscopic vision in Figure 4.2.

Doped parts of this fiber contain typical levels of chlorine impurity (∼1200 part per million
(ppm)) and OH-groups (∼60 part per billion (ppb)). Figure 4.3 shows the Ge and Cl trend
inside the fiber and preform samples. The average Ge and Cl concentrations measured by
EMPA for each zone are given in Table 4.1.

The drawing speed was ∼ 40 m/min, the drawing tension ∼ 70 g and the temperature of
the furnace ∼ 1600◦ C. These fibers exhibit pre-irradiation optical characteristics at 1.55 µm
close to that of commercial fibers, that is 0.34 dB/km.

F-doped canonical samples

In the fluorine doped samples the core consists in three zones (zones 2 to 4) with three different
F-concentrations (see Figure 4.4).

The F-incorporation inside a−SiO2 decreases its refractive index, so F-doping profile has
an opposite trend with respect to Ge as shown in Figures 4.1 and 4.5. Optical cladding (zone
1) corresponds to the zone with the highest F-doping region (∼1.8 wt%). More details are
presented in Table 4.1.

The outer cladding (zone 0) is made of pure-silica. Only a little part of the signal is guided
in this part of the waveguide. As a consequence, its contribution to the global transmission can
be considered as negligible. Doped parts of this fiber contain very small amounts of chlorine
impurity (∼1000 ppm) and OH-groups (∼200 ppb) typical of MCVD glasses. In this fiber the
44 4. The canonical samples

D is ta n c e f r o m p re fo rm c e n te r (m m )
-4 -2 0 2 4
0 .5
1 2 4
0 .4
1 0
3 3
8 0 .3

C l (% w t)
w t)

6
(%

2 2
2

0 .2
G e O

4 1 1
0 .1
2
0 0
0 0 .0
-4 0 -3 0 -2 0 -1 0 0 1 0 2 0 3 0 4 0
D is ta n c e f r o m f i b e r c e n t e r ( µm )

Figure 4.3: Impurities trend inside canonical Ge-doped fibers (lower x scale) and preforms (upper
x scale). Empty squares represent the Ge content, full circles the Cl content. The estimation of the
concentration values of the elements is obtained by EMPA analysis. Numbers from 0 to 4 refer to the
zones listed in Table 4.1.

0
1 1
2
3

Figure 4.4: Microscopic preform view of the F-doped canonical sample. Numbers from 0 to 4 refer to
the zones listed in Table 4.1. The different sample zones are clearly visible.

attenuation at 1.55 µm is about 1.9 dB/km.

P-doped canonical samples

The P canonical samples are made up of an outer undoped high purity silica layer, and
four internal cylindrical layers (core part, zones 1-4) of highly pure synthetic silica doped
4.1. Tested optical preforms and fibers 45

D is ta n c e f r o m p re fo rm c e n te r (m m )
-4 -2 0 2 4
0 .5
F
0 .0 C l
4 0 .4
0 0
-0 .5 3 3

C l- c o n te n t ( w t% )
F -c o n te n t (w t% ) 0 .3

-1 .0
2 2 0 .2

1 1
-1 .5 0 .1

-2 .0 0 .0
-4 0 -3 0 -2 0 -1 0 0 1 0 2 0 3 0 4 0
D is ta n c e f r o m f i b e r c e n t e r ( µm )

Figure 4.5: Fluorine (empty squares) and chlorine content (full circles) in F-doped fibers (lower x
scale) and preforms (upper x scale). Impurities concentrations were evaluated by EMPA analysis.

with different P-amounts. Phosphorus doping profile grows from the boundaries to the center

(a )

0
1
2
3
4
-4 -3 -2 -1 0 1 2 3 4
µm

Figure 4.6: Enlarged view of the P-doped fiber canonical sample. Numbers from 0 to 4 refer to the
various sample zones listed in Table 4.1 with different P-amounts. The various zones are clearly visible.

(Figure 4.1) following the desired multiple step distribution, as shown in the microscopic image
of the fiber sample in Figure 4.7.

The core-cladding part does not contain a relevant concentration of extrinsic impurities,
except for chlorine that is present with a maximum concentration of ∼0.2 wt%. The preforms
had an initial diameter of 12.84 mm with a 5 mm doped region and they were subsequently
46 4. The canonical samples

D is ta n c e f r o m p re fo rm c e n te r (m m )
-3 -2 -1 0 1 2 3
8 0 .5
4
7 C l
P 0 .4
6
3 3

C l- c o n te n t ( w t% )
P -c o n te n t (w t% )

5
0 .3

4
2 2 0 .2
3

2 1 1
0 .1
1
0 0
0 0 .0

-3 0 -2 0 -1 0 0 1 0 2 0 3 0
D is ta n c e f r o m f i b e r c e n t e r ( µm )

Figure 4.7: P-content (empty squares) and Cl-content (full circles) obtained by electron microprobe
analysis at various distances from the fiber (lower x scale) and preform (upper x scale) center.

cut into 6×6×1.5 mm3 samples and polished. Phosphorus and chlorine concentrations were
checked by EMPA, giving the results shown in Table 4.1 and in Figure 4.7.

The spectral attenuation at 1.5 µm, obtained with the cutback method, is about 0.5 dB/km
for this fiber. The OH content is evaluated as sensibly inferior to 10 ppb.
Chapter 5

Experimental set-ups

This chapter is concerned with the description of the instruments and setups used to perform
the experiments discussed in the rest of the work.

Several irradiation sources were used to analyze the generation processes of defects both
on fibers and on preforms. Additionally, various spectroscopic techniques have been employed
to identify the precursor sites in non-irradiated samples or stable point defects in irradiated
samples. Some of these techniques have been applied both on fibers and on preform samples
allowing the study of the influence of the drawing effects on defects generation. Other tech-
niques could be applied only on preform (time resolved luminescence, absorption) or on fiber
samples (radiation induced attenuation), for the particular sample structure.

5.1 Irradiations

5.1.1 UV laser irradiations

UV exposures at 5 eV (248 nm) were carried out at room temperature with two distinct set-
ups:a pulsed KrF laser and a continuum (CW) Ar-laser and .
UV pulsed laser
Irradiation exposures with the high power KrF pulsed laser were performed at a repetition
rate of 10 Hz, a duration time of 30 ns for every pulse and a pulse energy varying from 100
to 400 mJ. Irradiations on fibers were conducted by moving the samples at a constant speed,
transversally to the UV laser. The energy of the laser pulses is measured with a pyroelectric
detector; the accuracy, taking into consideration the laser fluctuations, is ±10%.
Continuum UV laser
The setup used for the continuum UV exposures was developed at the Hubert Curien Labo-
ratory. Irradiations were performed using an UV Argon Laser, emitting at 244 nm (5.1 eV)
48 5. Experimental set-ups

and having a gaussian intensity profile. For preform samples the beam was unfocused on the
center of the preform thanks to a spherical lens. The laser energy was measured by a power
meter with an accuracy of ∼5%. For the fiber samples the irradiation system is more compli-
cated [117,124] and it is described in Figure 5.1. The fiber samples are mechanically uncoated

Figure 5.1: Schematic representation of the experimental setup used for the CW ultraviolet (244 nm,
5.1 eV) exposures of optical fibers.

to prevent the UV light absorption by their acrylic coatings. A variable fiber length (from few
millimeters to several meters) filed past the CW laser beam. To this purpose, the fiber was
maintained in a V-groove and was interdependent with a tended thread by a counterweight.
This thread was pulled by a rotating motor. The 244 nm light was focused by a spherical lens,
leading to a spot size of some mm diameter on the fiber. By an appropriate choice of the fiber
translation speed (0.2−0.6 cm/s) and of the laser power (5−100 mW), it was possible to vary
the fluence value. The uncertainties in the fluence evaluation were mainly attributed to the
mechanical part of this setup and they were estimated within ±20%.

5.1.2 γ-ray and X-10 keV irradiations

The samples were exposed to γ-rays produced by a 60 Co source available in the IGS-3 irradiator
of the Department of Nuclear Engineering of University of Palermo. γ-rays have energies
between 1.17 MeV and 1.33 MeV. Irradiation was performed at room temperature, in ordinary
atmosphere at a dose rate of 1.39 kGy/h.

The 10-keV X-ray irradiations of fiber and preform samples were performed at room
temperature using an ARACOR Semiconductor X-ray irradiator [125,126] at the French atomic
energy center (CEA). The irradiated zone is homogeneous over a diameter size of ∼2.5 cm; the
dose was varied from 50 Gy up to 2 MGy (two different dose rates: 10 Gy/s and 0.1 kGy/s)
in fibers and from 1 kGy up to 2 MGy (dose rate = 0.1 kGy/s) in preform samples.
5.2. Absorption 49

5.2 Absorption

How shown in the introduction section, point defects in a−SiO2 can introduce new electronic
levels inside the valence and the conduction band. So, using electromagnetic radiation at
energy lower than the gap, it is possible to induce transitions corresponding to the absorption
bands of these materials.

The optical absorption experiments were conducted sending electromagnetic radiation on


the sample, varying continuously the light energy during a fixed interval and than analyzing
the transmitted light. According to the Lambert-Beer law, the transmitted radiation intensity
IT is linked to the incident intensity I0 by the relation 5.1 [127]:

IT (~ω) = I0 (~ω) exp{−α(~ω) · d} (5.1)

where d is the sample thickness and α(~ω) is the optical absorption coefficient, measured in
cm−1 . Knowing the absorbing species concentration, N , the absorption coefficient can be
expressed by the following equation:

α(~ω) = N σ(~ω) (5.2)

where σ(~ω) is the absorption cross section [50].

Optical absorption spectra presented in the result section were performed with two dif-
ferent spectrometers: the JASCO V-560 and the AVANTES S2000 spectrophotometers.

The JASCO spectrophotometer

The JASCO V-560 spectrometer is a double beam spectrophotometer providing the measured
absorption values in optical density (O.D.), defined as:
 
I0 (λ)
O.D. = log10 (5.3)
I(λ)
The incident beam I0 is split thanks to a beam splitter and it is alternatively sent to the
sample and to the detector. In this way we can have the control, and than the correction, of
every source fluctuation. The source is made of two lamps: a deuterium lamp, working in the
ultraviolet (UV) region from 340 nm to 190 nm, and a Xenon lamp, operating in the visible
and infrared range (340÷2500 nm). The detector is a photomultiplier (PMT).

The AVANTES spectrophotometer

In the optical fiber AVANTES S2000 spectrophotometer the source is a deuterium lamp (D2 )
that injects light into an optical fiber that splits up in two channels, referred to as Master
50 5. Experimental set-ups

and Slave. The optical fibers are multimode pure silica core\F2 -doped silica cladding with
diameter of 200 µm. They are loaded with H2 to better resist to the prolonged exposure to
UV light without being deteriorated. The light carried by the master channel gets out of the
fiber and is used as the probe beam (PB). The PB is collimated by two lens and it is coupled
to another fiber that brings it to the detector. The two lenses are mounted on independent
micrometric positioning controls (xyz), which permit both to control the alignment of the PB
to the sample and to optimize the collection efficiency after the sample. The slave channel
passes through a variable attenuator, after which it goes to the detector. Since the slave
channel does not traverse the sample, it could be used to correct experimental data for the
temporal drift of the lamp. The detector consists in a 1200 lines/mm grating with blaze at
300 nm, dispersing on a 2048 channels Charge Coupled Device (CCD) array. The instrument
works in the 200 nm−500 nm range with a spectral resolution of 5 nm. Before acquiring
a spectrum it is necessary to obtain a dark reference signal D(λ) from the master channel
when the D2 lamp is disconnected from the fibers. If I0 (λ) and I(λ) are the signals acquired
respectively without and with the sample, the absorption profile of the specimen is given by:
 
I0 (λ) − D(λ)
O.D. = log10 (5.4)
I(λ) − D(λ)

Absorption in the VUV spectral range

VUV absorption spectra in the 6.0÷7.5 eV range were obtained using an ACTON SP-150
single-beam spectrophotometer, equipped with a 30 W D2 lamp and two 1200 lines/mm
monochromators, and working in N2 flux. The acquired spectra were corrected by subtracting
the contribution due to surface reflectance.

5.3 Photoluminescence and Raman spectroscopy

5.3.1 Photoluminescence

An optical property of great importance in the study of point defects in a−SiO2 is the pho-
toluminescence (PL), i.e. the process by which a system excited by light with wavelength λex
emits light at λem > λex while decaying back to its ground state [50, 128, 129]. To picture
the physical processes determining the PL emission band, let us consider the two-level system
of Figure 5.2. Due to the absorption process OA, a number N(λex ) of defects will be in the
excited state during exposure of the sample to the excitation light. Some of these defects can
decay radiatively with a rate kr , thus originating the photon emission or luminescence. The
remaining excited defects relaxes back to the ground state by a temperature dependent non-
radiative process, with a rate knr , in which the energy is dissipated by emission of phonons.
5.3. Photoluminescence and Raman spectroscopy 51

OA kNR
ZPL kR

Figure 5.2: Electronic-vibrational level scheme of a two levels point defect. The continuous arrow
oriented upward represents the absorption transitions OA from the ground state to the exited state.
The continuous arrow oriented downward represents spontaneous emission from the excited state to the
ground state. The dotted arrow represents the non-radiative decay process. The grey arrow indicates
the ZPL transition and the dashed arrow within the excited levels represents the internal relaxation
process.

In Figure 5.2 the transition at lowest energy (that is from (0, 0) to (1, 0)) is called the zero
phonon line (ZPL). The PL intensity is given by:

IP L (λex , λem ) = kr N (λex )LP L (λem ) (5.5)

where LP L (λem ) is the emission lineshape determined by homogeneous and inhomogeneous


contributions [51,50,15]. The variation rate of the excited state population, N , depends on the
absorption and decay processes, both radiative and non radiative, according to the following
equation:
dN (λex , T )
= I0 (λex ) [1 − exp {−α(λex )d}] − (kr + knr (T ))N (λex , T ) (5.6)
dt
Steady state luminescence experiments are performed when the system undergoes a continuous
excitation. In this case dN/dt = 0 and combining Equations 5.6 with 5.5 we get:

IP L (λex , λem , T ) = ηI0 (λex ) [1 − exp {−α(λex )d}] LP L (λem ) (5.7)


kr
where η = kr +k nr
is the luminescence quantum yield, defined as the ratio between emitted
photons and absorbed photons.

Two basic types of measurements are possible:

• measuring the intensity IP L (λex , λem ) as a function of λem for fixed λex we obtain the
shape and intensity of the band emitted by the center, that is the emission spectrum
(PL);
52 5. Experimental set-ups

• acquiring IP L (λex , λem ) as a function of λex for fixed λem , we have the excitation spectrum
(PLE), which represents a measurement of the efficiency of the emission process in
dependence of the excitation wavelength.

Differently from stationary PL measurements, in the time-resolved PL measurement the


time decay of the emitted light after an exciting light pulse is studied. After the excitation of
a point defect with a light pulse, that produces a population of N (0) of the excited state, the
light source is switched off (I0 = 0) and N (t) decays as:

N (t) = N (0)e−t/τ (5.8)

with τ = 1/(kr + knr). From equations 5.5 and 5.8 we obtain the luminescence timedecay:

IP L (λex , λem , T, t) = kr IP L (λex , λem )N (0)e−t/τ (5.9)

At sufficiently low temperatures, the non-radiative decay channels are usually quenched,
i.e. knr  kr , so that the measurement directly yields the radiative decay time τ = 1/kr . The
importance of the knowledge of τ relies in the possibility of calculating the oscillator strength
f of the center [50]:
1 m~2 c3
f= 2 (5.10)
E τ 2e2

Not all the point defects that feature a measurable absorption band decay by emitting
luminescence, but when this occurs, their study by PL spectroscopy has some important
advantages with respect to OA. In particular, PL is more selective, as it often allows to isolate
a center whose absorption band overlaps to those arising from other defects, based on the
different emission properties.

5.3.2 Stationary and time resolved luminescence setup

In this section we are going to describe the experimental setup used for acquiring the lumi-
nescence data of preform samples. As shown in the schematic representation of Figure 5.3
the equipment is mainly constituted by a laser source, a sample chamber, a dispersion system
and a detection one. The tunable laser (VIBRANT OPOTEK) [130] is an integrated system
that emits pulses of 5 ns duration with a maximum repetition rate of 10 Hz in the range
(210-2400) nm. The excitation beam with a spot size of ∼1mm2 hits the sample mounted
in the so-called 45◦ back-scattering geometry. The emitted light is collected by a lens and
then arrives at the detection system. The energy of the laser pulses is measured with a py-
roelectric detector capable of giving in output a short electric pulse for each laser shot (some
mus), whose amplitude is read by a digital meter. It is positioned before the sample holder
5.3. Photoluminescence and Raman spectroscopy 53

Xe-lamp

Tunable Laser Sample


OPOTEK

CCD+ Delay
Spectrograph
generator

Figure 5.3: Schematic representation of the equipment used for the luminescence measurements.

to measure the intensity of incident laser radiation. The accuracy of pulse energy measures,
taking into consideration the laser fluctuations, is ∼10%.

Time resolved luminescence spectra are performed with a detection system consisting of
a Spectrograph and an Intensified CCD Camera. This latter amplifies the input luminescence
signal: for each photon that strikes the photocathode surface many photons are produced.
Moreover, the possibility of varying the photocathode voltage allows to enable or to disable
the CCD: in the GATE ON mode the photocathode voltage is -200 V and the CCD sees
the light, in the GATE OFF mode the photocathode voltage is 0 and the CCD does not
see the light. This peculiarity permits the detection of time resolved luminescence spectra
synchronized with the laser excitation pulses. In fact, the CCD is triggered by an electronic
synchronization signal produced by the laser 60 ns before the pulse. The CCD can accumulate

(0.1-1)s
4 ns

τ’ Δt
IPL
IPL

λem λem

Figure 5.4: Diagram of the CCD timing.

in a time window defined by the width parameter ∆t (Gate Width) and by its delay τ (Gate
Delay) from the origin of the time scale. So, as shown in the diagram of Figure 5.4, the Gate
54 5. Experimental set-ups

Width determines the amplitude of the time window during which the CCD is enabled to
reveal the luminescence light (GATE ON mode); while the Gate Delay regulates the temporal
shift of the acquisition window with respect to the trigger signal.

In the next chapters all luminescence spectra are presented as a function of the energy E
instead of the wavelength λ, so they require a specific correction procedure: the CCD counts
are directly proportional to the luminescence spectral density dI/dλ that is the intensity
collected with a constant spectral bandwidth dλ. Since energy and wave number are linked by
the relationship E = hc/λ, the spectral density dI/dE with respect to E must be multiplied
for spectral dispersion λ2 because to a constant spectral bandwidth in λ corresponds a spectral
bandwidth which depends from the emission energy E.

5.3.3 Photoluminescence under synchrotron radiation excitation

Excitation (PLE) end emission spectra excited in the Vacuum UV (VUV) range on preform
samples were carried out under excitation by pulsed synchrotron radiation at the SUPER-
LUMI station on the I-beamline of HASYLAB, DESY (Hamburg) [131]. Measurements were
performed in the spectral range 4.5÷9.0 eV, with a pulse width of 130 ps, an interpulse of
500 ns, and a spectral width of 0.3 nm. The excitation beam is directed into the primary
monochromator with a 1200 lines/mm grating (Figure 5.5; the excitation wavelength can be

Figure 5.5: Schematic representation of the experimental station used for PL, time resolved PL and
PLE under synchrotron radiation (figure adapted from [131]).

varied from 310 nm to 50 nm (4.0 to 24.8 eV). The emitted light was spectrally dispersed by
a 300 grooves/mm grating blazed at 300 nm and acquired by a liquid nitrogen cooled CCD
camera (1100 Princeton Instruments) for PL spectra.

Luminescence spectra were corrected both for the spectral response and dispersion of the
detecting system, while excitation spectra were corrected for the spectral efficiency of the
5.3. Photoluminescence and Raman spectroscopy 55

exciting light, using a sodium salicylate sample as a reference. Excitation bandwidth was
0.3 nm, while emission bandwidth was 20 nm.

During laser-excited luminescence and measurements, we put the sample behind a prop-
erly built mask so as to allow spatial selection of the various preform zones (see Section 4.1).

In time resolved PL and in PL under synchrotron radiation, temperature dependencies


(from 10 to 300 K) were investigated using continuous flow helium cryostats.

5.3.4 Raman measurements

The Raman effect is the anelastic scattering of light (by a molecule or a point defect) due
to emission or absorption of a vibrational quantum [132, 127, 133]. If Ei is the energy of
the incident photons, scattering at Es < Ei implies the excitation of a vibrational mode of
energy ~ω = Ei − Es . A Raman spectrum consists of a plot of the scattered intensity as a
function of ω. This spectroscopy allows to probe the vibrational modes of a molecule or a
point defect, sometimes bearing some advantages with respect to common IR spectroscopy.
For example, it can happen that a vibrational mode is Raman-active but non IR-active, or
vice versa [132, 127]. Raman spectra presented in this PhD Thesis were all performed by the
microRaman spectrometer described here below.

5.3.5 Confocal Micro-spectroscopy setup

Confocal microscopy luminescence (CML) measurements and microRaman measurements were


performed with the LabRAM Aramis (Jobin-Yvon) integrated confocal microRaman system
whose scheme is presented in Figure 5.6.

The confocal microscope is coupled to a 460 mm focal length spectrograph equipped with
a four interchangeable gratings turret. The different excitation wavelengths are supplied by up
to one internal laser (He-Ne, 633 nm, 2.0 eV) and by two external lasers: a He-Cd laser working
at 442 nm (2.8 eV) and 325 nm (3.8 eV) and an Argon laser working at 488 nm (2.5 eV). Briefly
the principle of confocal microscopy consists of focusing the laser source through the objective
of the microscope and carrying out a spatial filtering of the signal coming from the illuminated
volume, by using a diaphragm of small diameter placed in the conjugated plane where the
magnified image of the sample is formed by the objective (see Figure 5.7).

On the incoming path, the laser beam is reflected towards the microscope by the means of a
special filter (holographic notch filter or dielectric edge filter) used in injection/rejection mode.
On the return path to the spectrograph, the Raman backscattered light is fully transmitted
56 5. Experimental set-ups

CCD detector Grating turret Laser HeNe External Lasers

Filter turret

Autofocus

Camera

White light source

Figure 5.6: Top view of the the LabRAM Aramis (Jobin-Yvon) integrated confocal microRaman
system with the optical path.

Figure 5.7: Scheme of micro-luminescence and micro-Raman analysis set-up.

through the filter towards the confocal slit-hole located at the entrance of the spectrograph.
The spectrograph disperses the multichromatic Raman signal onto the CCD multichannel
detector.

The optical drawer constitutes the coupling platform between the laser, the sampling
chamber (microscope or macro chamber) and the spectrograph. In order to reduce the laser
power at the sample, a density filters wheel driven by the software, can be used. During
our measurements we make sure that the power level of the probe light was reduced to few
5.4. Electron Paramagnetic Resonance measurements 57

hundreds of µW to avoid photobleaching effects.

For Raman measurements, the Raman signal is then collected by the same microscope
objective (backscattered configuration) and follows the return path to the spectrograph. The
raman filter filters out the backscattered laser light (Rayleigh scattering) whereas the Raman
signal is transmitted to the spectrograph entrance slit.

The sample can be translated, under computer control, with an accuracy of about 0.1 µm.
The excitation beam penetrates of a few micrometers into the sample. The spot diameter
focalized on the samples varies as a function of the microscope objective used. Using a ×50
objective, the spot size is some µm. With this spot size it is possible to inspect in detail not
only the preform samples, but also the fibers. This is a new tool of our study: previously the
largest part of scientific investigation on defects in optical fibers consisted in the direct study of
bulk samples and the subsequent transfer of information to the fibers [19,20,21]. However, this
approach cannot take into account the peculiarities implied in the fiber preparation procedure,
such as the drawing process, which can generate precursors and influence the defect generation
[22, 134, 135]. Micro-Raman and micro-luminescence investigations, allowing inspection of
defect directly inside optical fibers, permit us of overcoming this limit.

5.4 Electron Paramagnetic Resonance measurements

The Electron Paramagnetic Resonance (EPR), also referred to as Electron Spin Resonance
(ESR) spectroscopy measures the absorption of microwave radiation corresponding to the
energy splitting of an unpaired electron when it is placed in a strong magnetic field. EPR is
a spectroscopic technique that detects the presence of these unpaired electrons in a chemical
system [136, 137, 138]. This can yield meaningful structural and dynamic information, even
from ongoing chemical or physical processes (i.e. kinetics, etc.) without influencing the process
itself. In the absence of an external magnetic field the two possible electron spin states (spin
up and spin down) are degenerate. When an atom or molecule with an unpaired electron is
placed in a magnetic field, the spin of the unpaired electron can become aligned either in the
same direction (spin up) or in the opposite direction (spin down) of the applied field. These
two possible electron alignments have different energies (i.e. are no longer degenerate) and are
directly proportional to the applied magnetic field strength. This is called the Zeeman effect.

The EPR measurements reported in this Thesis were performed by a Bruker EMX spec-
trometer working at ω0 =9.8 GHz. In Figure 5.8 it is reported a simplified scheme of the
instrument. The microwave radiation travels down a waveguide to the sample, which is held
in place in a microwave cavity held between the poles of two magnets. A variable attenuator
permits to regulate the actual power Pi incident to the cavity from a maximum value 200 mW
down to 200 nW. In this way, Pi is usually chosen so as to avoid the saturation of the observed
58 5. Experimental set-ups

Source Detector

A/D
Attenuator
RC Filter
A/D Amplifier
Lock-in

Integrator
Modulator
Magnet
Resonant
cavity
EPR Signal

Figure 5.8: Schematic representation of the Bruker EMX spectrometer. The main sections are visible:
magnet, cavity, source, attenuator, modulation system and detector.

magnetic resonance transition. The microwaves arriving on the entrance of the cavity are
partially absorbed by the sample and partially reflected to join the revelation system. The
reflected power PR is measured by a detector that gives a current signal I proportional to
the square root of PR . Spectra are obtained by measuring the absorption of the microwave
radiation while scanning the magnetic-field strength. EPR spectra are usually displayed in
derivative form to improve the signal-to-noise ratio.

It is important to underline that the doubly-integrated intensity of the EPR spectrum is


proportional to the number N of paramagnetic centers. In this sense, if a reference sample is
available, EPR may be used to provide a measurement of the absolute concentration ρ = N/V
of every paramagnetic defect, where V is the volume of the sample. To this purpose, in this
work it was used a specimen where the absolute number of E 0 centers (purposely generated
by γ irradiation) was known by spin-echo measurements [139, 140]. The accuracy of the
absolute concentration measurements obtained by ESR, based on comparison with the spin-
echo reference sample, is estimated as 20%a .

a
this error never explicitly appears when reporting in the following the uncertainties on the concentration
measurements. The reason for this choice is that the uncertainty on the concentration of E 0 in the spin-
echo reference sample plays the role of a systematic error, affecting in the same way all the concentration
measurements here reported.
Part III

Ge-doped fibers and preforms


Chapter 6

Measurements on non-irradiated samples

As discussed in detail in the introductory chapters (section 2.1), identification of de-


fects responsible for the permanent change of the refractive index in Ge-doped glasses under
radiation exposure is a clue to make clear the microscopic origin of transparency loss and pho-
tosensitivity. Consequently, the study and the characterization of point defects in Ge-doped
a−SiO2 is a very effective subject to understand the fiber properties and to improve their
performances.

We have seen that germanium lone pair centers are the main responsible defects for
photosensitivity property. The GLPC optical activity variation under UV-radiation exposure
contributes to the permanent refractive index variation of the glass. So the exact knowledge
of the GLPCs concentration before radiation exposure and the modifications induced by the
drawing process, result to be a fundamental aspect in optical fibers technology. The study of
GLPCs variations in optical fibers can so be adopted as probe for testing the refractive index
modulation induced by external factors.

To this aim, this section is devoted to introduce the optical properties of the as-grown ma-
terials performing direct investigation of GLPCs inside our germanosilicate step-index canon-
ical fibers and preforms.

First of all we performed absorption measurements on preforms samples before irradiation


in order to investigate GLPC presence in the different sample zones. Figure 6.1 shows the OA
spectra detected by the use of the AVANTES spectrometer in the spectral range 2.5-5.5 eV;
in the figure the variation of the OA intensity as a function of the different sample zones,
that is as a function of the Ge-content, is also shown. For GeO2 >5% the OA signal saturates,
indicating defects concentration higher than our detection system (∼70 cm−1 ). As expected,
the OA band centered at 5.2 eV and related to GLPCs, grows with the Ge-content.

Though the determination of the GLPC spatial distribution in optical fibers is crucial to
62 6. Measurements on non-irradiated samples

9 0

) [G e O ] (w t% )
-1

2
A b s o r p tio n C o e ffic ie n t ( c m

6 0 1 1
8 G L P C
5
3
0
3 0

2 .5 3 .0 3 .5 4 .0 4 .5 5 .0 5 .5
E n e rg y (e V )

Figure 6.1: OA spectra detected in pristine Ge-doped canonical sample. The arrow specifies the GeO2
content from the exterior to the inner sample part. The OA band at ∼5.2 eV is due to GLPC centers
whose microscopic structure is showed in the inset. For [GeO2 ]>5 wt% the OA signal saturates.

probe the silica refractive index variation, convenient experiments have not been performed
yet, the main obstacle being the small fiber dimensions. So far, studies dealing with GLPCs
were extracted from bulk samples and than transferred to fibers [19, 20, 21], thus remaining
affected by an intrinsic deficiency caused by the information lack related to the drawing pro-
cess. Nevertheless the use of Confocal Microscopy Luminescence (see section 5.3.5) allows to
examine the variation of the PL bands linked to GLPCs even on micrometric scale.

Than we performed direct investigation of GLPCs inside germanosilicate step-index op-


tical fibers by using CML technique with the purpose of examining their radial distribution
along the core. The comparison with the corresponding preforms permits to recognize the
actual role of the drawing process in modifying the defect formation, and consequently to
control the fiber photosensitivity.
The recognition of GLPC concentration in preform samples is not possible via OA B2β band
due to our measurement conditions: in the central sample region the absorption coefficient is
too high and cause the saturation of the detected signal (see Figure 6.1), thus preventing a
careful evaluation of defects concentration.
There exists opinion that there could be two very closely spaced optical bands at about 5.2 eV
due to Ge-related oxygen-deficient centers, and that only one of them could give rise for
63

PL [55]. Otherwise, using the linear correlation between B2β band and the GLPC PL activity,
we assumed that the GLPC absorption band was unique. This assumption could systemat-
ically affect the absolute evaluation of GLPC concentration, nonetheless, for our purposes,
the relative concentration variation is still meaningful. In principle this problem could be
overcome estimating concentration by the singlet-triplet GLPC absorption band at ∼3.8 eV.
Nevertheless, this OA band is much weaker than the singlet-singlet one (about 104 times) and
it is not detectable in our samples.

GLPCs in fibers and preforms were revealed by the He-Cd ion laser excitation line of
the LabRam Aramis spectrometer (photon energy 3.8 eV, power 0.5 mW). We used a ×40
objective and a diaphragm diameter of 100 µm. The excitation beam penetrates of a few µm
into the sample as to be sure that the collected signal was not due to abnormal surface defect
concentration detection.
GLPCs were checked monitoring the intensity variation of the 3.1 eV phosphorescence band
by direct excitation S0 →T1 (Eex = 3.8 eV) (see GLPCs levels scheme in the left inset of
Figure 6.2). As an example, in Figure 6.2 OA at 5.2 eV and PL at 3.1 eV (right inset)
detected in zone 2 of preform sample ([GeO2 ]∼5 wt%) are depicted. The use of a laser probe

5 0
S 1
P L ( a r b . u n its )

T 1
O A
E E X
P L
4 0 5 .2 e V
)

3 .8 e V 3 .1 e V
-1

S 0
A b s o r p tio n C o e ffic ie n t ( c m

2 .7 3 .0 3 .3
3 0 E n e rg y (e V )

2 0

1 0

4 .6 4 .8 5 .0 5 .2 5 .4 5 .6 5 .8
E n e rg y (e V )

Figure 6.2: OA spectrum detected in zone 2 ([GeO2 ]∼5 wt%) of Ge-doped pristine preform. The
right inset shows the PL spectrum al 3.1 eV obtained exciting at 3.8 eV in the same preform zone. In
the left inset transitions giving rise to the observed bands are schematically depicted.

at 3.8 eV has a double advantage:


64 6. Measurements on non-irradiated samples

• First of all we have seen that the germanosilicate exposition to UV laser induces the
bleaching of the optical activity related to GLPCs [21] (see section 3.2.1). Consequently
exciting the GLPC activity by the use of a laser beam at the energy of 5.2 eV, that is
by the excitation S0 →S1 , could produce an intensity diminution of the 3.1 and 4.2 eV
PL bands not related to an actual defect concentration decrease in pristine samples, but
rather to an undesirable sample irradiation. The result will be an invasive measurement
that alters the evaluation of defects concentration. in contrast, as shown in Figure 6.1,
our samples are almost transparent in the excitation spectral range from 2.5 to 4 eV,
making sure that the use of a laser probe at 3.8 eV results in a non invasive analysis.

• In the second place, it is worth to note that in our samples the absorption at ∼5 eV is very
high: comparing Figure 6.1 we can see that at 5.2 eV the absorption coefficient exceeds
70 cm−1 in the central core region. As a consequence, the luminescence signal obtained
exciting in the peak of the OA band, will not be proportional to GLPC concentration, due
to the fact that the sample will be not uniformly illuminated during PL measurements.

GLPC-related emissions at ∼ 3.1 eV detected in the fiber layers with different Ge-content
are shown in Figure 6.3. CML spectra were recorded from 370 nm (3.4 eV) to 500 nm (2.5 eV)

Figure 6.3: 3D plot of the CML band related to the T1 →S0 GLPC transition, measured at different
distances from the fiber core center and obtained under excitation at 3.8 eV in the fiber sample. The
PL intensity decreases with the distance from the center, i.e. with the GeO2 content.
6.1. Discussion: the drawing effect 65

with a spatial resolution of about 1 µm. Qualitatively, we observe that the PL intensity
decreases with the distance from the center, i.e. with the Ge content.

6.1 Discussion: the drawing effect

To perform a quantitative calculation of defects concentration, [GLPC], we convert the PL


intensity PL(GLPC) to an absolute concentration measurement thanks to the evaluation of
the oscillator strength, f , for the GLPC centers. Starting from Smakula’s equation [50], the
5.2 eV absorption intensity in the preform is determined by its linear correlation with the PL
bands and its oscillator strength, estimated to be f = 0.07 [21] from the radiative decay time
of the 4.2 eV (τ ' 7.8 ns) transition, measured at T=10 K under synchrotron radiation [141]:

[GLP C] = P L(GLP C) × 3.62 × 1013 [cm−3 ] (6.1)

The calculation giving rise to Equation 6.1 was performed for preform samples and than
transferred to fibersa .

GLPC-related emissions at ∼ 3.1 eV detected in the fiber layers with different Ge-content
are shown in Figure 6.4. To draw the GLPC radial distribution in the fiber sample, the
side (a) of Figure 6.5 shows their concentration measured at various distances from the fiber
center. For sake of clarity in the lower part of Figure 6.5 (side (b)) the microscopic fiber view
is sketched, so as to localize the various zones where these defects are detected.

To point out the influence of drawing effects on the defect generation, Figure 6.5 shows
the comparison of [GLPC] as a function of [Ge], as measured in preform and in fiber. In the
preform such a dependence has an almost linear trend over the whole range with a coefficient
βpref ' 2×10−3 . In contrast, in the fiber is observed a discontinuity at [Ge]≈ 14×1020 cm−3
(8 wt%): over the lower range the experimental points overlap with those of the preform,
whereas at larger Ge concentrations the ratio [GLPC]/[Ge]=βf iber exceeds βpref up to about
five times (βf iber ' 9×10−3 ) for [Ge] ' 2×1021 cm−3 (11 wt%).

The interpretation of this result could be related to the differences between radial stress
of the fiber and of the preform. Both materials are characterized by a thermoelastic stress
that increases linearly with the Ge concentration [142, 143, 144]: as this kind of stress is
linked to the temperature expansion coefficient, it results in a positive tensile stress [143].
This means that highly doped core regions, as those investigated in our experiment, are more
likely to exhibit tensile core stresses. On the other side, the drawing process can introduce
an additional negative compressive stress in the fiber samples [145]. The interplay between
these two contributions influences therefore the net fiber stress, which could result to be either
a
Our experimental set-up does not allow to perform spatial resolved OA measurements on fibers.
66 6. Measurements on non-irradiated samples

Distance from center (µm)


-6 0 -4 0 -2 0 0 2 0 4 0 6 0

(a )
1 5
)
-3
c m

1 0
1 8
[G L P C ] (1 0

(b )

-4 0 -2 0 0 2 0 4 0

Figure 6.4: (a): GLPC concentration obtained by a transversal mapping of the fiber sample at various
distances from the core center. (b): microscopic view of the fiber cross section. The different layers are
distinguishable. The x scales in the upper and lower figures coincide.

negative or positive in core zones with a different Ge content [145, 143], differently from the
preform where the radial stress is always positive. We suggest that a sudden change of the
stress in the core inner part causes the observed discontinuity of GLPC concentration, due to
a new defect generation mechanism.

This result is relevant for its connection with the refractive index change in the core zones.
As the drawing process dramatically modifies the distribution of UV optically defects inside
the fiber, the photosensitivity properties cannot be deduced simply by the fiber composition
or by measurements on preform samples. From a handy point of view, it seems possible to
6.1. Discussion: the drawing effect 67

2 0

P re fo rm
1 6
F ib e r

)
-3
c m
1 8
[G L P C ] (1 0 1 2

2 3 4 5 6 7 8 9 1 0 1 1 1 2
[G e O 2
] (w t% )

Figure 6.5: GLPC concentration trend detected in the different sample layers as a function of Ge
content both in fiber (empty squares) and preform (full circles).

Figure 6.6: Left: Assial stress as a function of GeO2 content in a graded index Ge-doped fiber. Right:
Stress profile of a graded index Ge-doped fiber. Figures from Lee et al. [142].

modulate the fiber photosensitivity during the fabrication process by a systematic control of
the fiber drawing conditions.

In conclusion, our experimental results point out that the CML is a powerful and not
invasive tool to probe the defect spatial distribution directly in optical fibers. We have demon-
68 6. Measurements on non-irradiated samples

strated that the GLPC radial distribution is different in fibers and preforms. This difference
can change the photosensitivity of the fiber regarding to the preform sensitivity. A possible
explanation for this effect is the changes in the mechanical stress introduced by the drawing
process. Such a stress can lead to an enhancement of the defect generation in some specific
part of the fiber. Our study shows that the drawing process has to be considered to explain the
defect generation mechanisms at the origin of the glass photosensitivity. More detailed analy-
sis with fibers designed with variable process parameters and stress measurements have to be
done to control this phenomena in order to enhance the photosensitivity of germanosilicate
fibers or to reduce their degradation under radiative environments.
Chapter 7

Effects of the UV and X-ray irradiation

As pointed out in section 3.2.1 the identifications of centers causing the degradation of the fiber
properties under radiation exposure is one of the main subject in the current study of optical
fibers response. Many studies, both experimental and computational, have investigated the
generation and conversion processes of Ge-related point defects under radiation exposure. The
research interest is mainly motivated by the practical importance of these materials in photonic
applications, from standard optical fibers as waveguides for telecommunications to non-linear
optical devices. This multi-use of Ge-doped silica is strongly influenced by its response to
radiation; for instance, it is known that radiation exposure induces Ge-related defects that
are cause of attenuation for the waveguides, thus leading to detrimental losses of part of the
transmitted signals. The identification of the specific defects, sensitive to radiation, is therefore
of crucial importance to make clear the microscopic origin of technologically relevant processes.
Nerveless the current understanding of many important aspects is not at all complete, and
several relevant questions are still debated.

In the following we present the results of a series of experiments on the effects of the
radiation exposure on canonical Ge-doped fibers and preforms [126, 123, 146, 122] to get a
deeper clarification of the microscopic origin of the photoinduced structural changes observed
in these materials.

7.1 EPR results

One of the main techniques employed to investigate the effects of irradiation on Ge-doped
canonical samples is the electron paramagnetic resonance (see section 5.4).

After X-ray exposure we observe the presence of paramagnetic centers, as an example


Figure 7.1 shows the EPR spectra detected in fiber and preform samples after 20 kGy X-ray
70 7. Effects of the UV and X-ray irradiation

irradiation.

g = 2 .0 0 7

2 4 X d o s e 2 0 k G y
P re fo rm
F ib e r
1 6

g = 1 .9 9 3
-8
g = 1 .9 9 9

3 4 6 3 4 7 3 4 8 3 4 9
M a g n e t ic F ie ld ( m T )

Figure 7.1: EPR spectra in fiber (open circles) and preform (full line) Ge-doped samples after a X-dose
of 20 kGy.

We can note that in both samples the irradiation induces the same defect species but
with a different efficiency: defect concentration in fiber is higher than in preform. It is also
apparent that these EPR spectra are composite signals, resulting from the overlap of different
contributions. As shown in Figure 7.1 we can distinguish the g values of three germanium
related defects: 2.007, 1.999, 1.993. To obtain information on the concentration of every defects
specie contributing to the total EPR lineshape, we performed a deconvolution of the measured
spectra. We found that the EPR spectra can be least-square-fitted, after any irradiation, by
a linear combination of the lineshapes associated with the following Ge-related paramagnetic
point defects, already described in section 3.2.1: (i) Ge(1) (GeO4• ), (ii) GeE0 (≡ Ge• ) and
(iii) Ge(2), whose structure is still questioned between models consisting either in a trapped
electron center or in a hole center (see section 3.2.1 for details). As an example, Figure 7.2
shows the fitting result on fiber (panel (a)) and preform (panel(b)), after a X-ray dose of
20 kGy. Red solid line plots the best fitting function. The normalized EPR lineshapes of
these three defects, used in the best fitting procedure, are shown separately in panel (c) of
the Figure 7.2. Ge(1) and GeE0 lineshapes are experimental curves measured in γ-irradiated
Ge-doped samples [147], where it has been possible to single out the two defects. The Ge(2)
7.1. EPR results 71

3 0

(a ) (c )
2 0

1 0 G e (1 )
E P R In te n s ity ( a r b . u n its )

G e E ’
(b )
2 0

1 0

G e (2 )
0

3 4 6 3 4 7 3 4 8 3 4 6 3 4 7 3 4 8 3 4 9
M a g n e tic F ie ld ( m T )

Figure 7.2: EPR signal detected in (a): fiber and (b): preform samples after 20 kGy X-ray exposure.
Solid red lines plot the best fitting functions obtained as a superposition of the three lineshapes related to
GECs and GeE0 centers. Panel (c) reports the three base lineshapes used for the best fitting procedure.

lineshape is a curve taken from literature, obtained by Friebele et al. [101] by a simulation
procedure aimed to fit EPR lineshapes observed in γ-irradiated multimode Ge-doped fibers.
We point out that fitting an EPR signal with a linear combination of single-defect lineshapes
is founded on the assumption that the centers are sufficiently far from each other that their
lineshape is not influenced by mutual interactions. This condition is usually satisfied at defects
concentrations up to 1018 cm−3 .

Figure 7.3 shows the changes in EPR spectra after different X doses in fibers and in
preforms. Qualitatively we can deduce that the EPR signal intensity increases with increasing
the X-ray dose both in fibers and in preforms, nevertheless the lineshape does not significantly
change.

We show that after UV pulsed laser, continuum UV laser irradiations or γ exposure the
EPR spectra show the appearance of the same defect species with similar trend, even if with
different intensity. As an example in Figure 7.4 the comparison between a spectrum obtained
in preform sample after a X-dose of 2 MGy and the signal detected in gamma irradiated
preform sample (total dose=91 kGy) is reported. In the rest of this section, only results
72 7. Effects of the UV and X-ray irradiation

3 0 3 0
(a ) (b )
P re fo rm F ib e r
X d o s e (k G y )
2 0 2 0 0 0 2 0
X d o s e (k G y )
2 0 0 0 2 0
2 0 5
1 0 2 1 0
2

0 0

-1 0 -1 0
3 4 6 3 4 7 3 4 8 3 4 9 3 4 6 3 4 7 3 4 8 3 4 9
M a g n e tic F ie ld ( m T )

Figure 7.3: EPR signal intensity as a function of the total deposited X-dose in (a): preform and (b):
fiber canonical Ge-doped samples.

obtained after X-ray irradiations will be discussed.

From the best fit coefficients appearing in the linear combination of the GEC defects
(Figure 7.2(c)), we are able to evaluate the contribution of the three defect species to the
overall signal measured upon exposure at each total deposited X-dose. We point out that
EPR measurements give results on the whole sample, regardless the different sample layers
with various Ge amount. The contribution of each paramagnetic specie, obtained with the
above described procedure as a function of the X-ray dose is reported in Figure 7.5. It is
evident that the GECs grow with the same trend on increasing the X-dose, Ge(2) being the
defect induced in largest concentration over the considered dose range. Both in fiber and in
preform Ge(1) and Ge(2) grow with the X irradiation dose till a saturation value, in contrast
GeE0 centers do not reach a saturation value, at least at the considered doses.

7.2 Optical absorption

Due to the reduced fiber dimensions, absorption measurements in the UV-visible region were
performed only on preform samples. Results will be than extended to fibers. In our specific
case, this experimental procedure is supported by results showed in section 7.3.3, in which
the effects of the drawing process are illustrated: the results clearly show that only at low
7.2. Optical absorption 73

X ir r a d ia te d ( 2 M G y )
γir r a d ia te d ( 9 1 k G y )

In te n s ity ( A r b . U n its )
E P R

3 4 6 3 4 7 3 4 8 3 4 9
M a g n e t ic f ie ld ( m T )

Figure 7.4: EPR spectra detected in preform samples after 10 keV-X-ray exposure (total deposited
dose 2 MGy, full line) and after γ exposure (total dose=91 kGy, open circles).

irradiation dose the drawing procedure can influence defect generation.

The transmission losses from 2.5 eV to 5.5 eV of the preform sample after an X dose of
2 MGy, detected with the AVANTES spectrometer, are shown in the panel (a) of Figure 7.6.
As the beam spot size diameter is smaller than the different zones dimensions, we are able to
selectively investigate the distinct sample layers. The spatial resolution of this measurement
is obviously limited by the diameter of the beam spot, that is about 0.5 mm.

Through this procedure we can easily obtain the OA spectra as a function of the average
GeO2 content, as shown in Figure 7.6(a). The OA spectra referred to the different sample
zones evidence the wing of an induced band centered at energies higher than 4.5 eV, more and
more intense with increasing the Ge-content. We notice that the detection of the whole band
in the heavy doped zones is limited by the presence of the OA band centered at 5.1 eV and
related to the GLPCs, whose amplitude is higher than our detection limit (∼50 cm−1 ).

To single out the different OA components, the spectra have been fitted by Gaussian
curves. Panel (b) of Figure 7.6 shows the analysis of the spectrum detected in the preform
zone containing 5 wt.% of GeO2 ; it is accounted for by three bands: the first centered at
(4.6±0.1) eV, FWHM of (1.8±0.2) eV, the second centered at (5.14±0.02) eV, FWHM of
74 7. Effects of the UV and X-ray irradiation

2
1 0
(a )
1
1 0

0
1 0
In te n s ity ( a r b . u n its )

-1
1 0
G e (1 )
G e E ’
2 G e (2 )
1 0
(b )
1
E P R

1 0

0
1 0

-1
1 0
0 1 2 3
1 0 1 0 1 0 1 0
X d o s e (k G y )

Figure 7.5: Integrated EPR intensity related to GECs and GeE0 centers as a function of the X-ray
dose detected in (a): fiber and (b): preform samples.

(0.50±0.02) eV and the last centered outside the investigated range (∼6 eV) to take into
account higher energies contributions.

In regard to the absorption tail centered at ∼4.5 eV, Figure 7.7 shows its dependence
from the X total deposited dose. The graph illustrated the OA increasing, from 2.7 to 4.8 eV,
in the central part of the preform sample, containing the higher GeO2 -content (zone 4 in
Figure 4.2 and in Table 4.1, GeO2 ∼11%). Similar trends are observed in the other sample
zones. From the Figure 7.7 we can qualitatively conclude that, with respect to pristine sample,
the radiation exposure causes an increasing in transmission losses at about 4.5 eV.

7.3 Discussion: radiation effects

The use of several spectroscopic techniques allows from one side to identify defect species
causing radiation losses in the considered spectral range and on the other side to localize
7.3. Discussion: radiation effects 75

5 0
6 0
(a ) (b )
4 0
5 0
) [G e O 2
] (w t% )
[G e O ] = 5 w t%
-1

2
4 0 1 1 3 0
A b s . C o e ff. (c m

8
3 0
5 2 0
2 0
3
0 1 0
1 0

0 0

3 4 5 3 4 5
E n e rg y (e V )

Figure 7.6: (a): Transmission losses detected in the various preform zones after 2 MGy X-ray exposure.
The arrow specifies the GeO2 content from the exterior to the inner sample part. (b): Absorption
spectrum detected in the zone 2 (5 wt.% of GeO2 ) superimposed to the best fit curve, dashed lines
represent the Gaussian components.

them inside the fiber sample in the considered spectral range. This result is a very original
and interesting feature in the study of the optimization of fiber manufacture and it will be
discussed in the following two points:

• localization of defect species

• generation of GECs defects

7.3.1 Localization of defect species

In the previous section we have shown the EPR results on preform and fiber Ge-doped canon-
ical samples after radiation exposure. The results allow to identify various defect species
present in the samples, nevertheless this measurement procedure does not permit to localize
the defects in the various sample layers. We have seen in Section 4, that our specific canonical
samples were made with concentric layers containing different amount of germanium. We
can reasonably expect that the defect distribution along the sample layers is not uniform:
Ge-related species will be concentrated in the central sample zone with the grater Ge-content
(<[GeO2 ]>∼11%).

In the case of absorption measurements, this assertion is immediately confirmed by the


76 7. Effects of the UV and X-ray irradiation

3 0

Z o n e 4
2 5
[G e O 2
] ~ 1 1 w t%

2 0
)
-1

X -d o s e (k G y )
A b s . C o e ff. (c m

1 5 2 0 0 0
2 0
2
1 0 0

2 .7 3 .0 3 .3 3 .6 3 .9 4 .2 4 .5 4 .8
E n e rg y (e V )

Figure 7.7: Transmission losses from 2.7 to 4.8 eV, detected in the central preform zone (zone 4,
[GeO2 ]∼11%) as a function of the X-ray exposure. The arrow specifies the X total deposited dose.

radiation induced losses trend illustrated in Figure 7.6. While the absorption band centered
at 5.1 eV is undoubtedly attributed to GLPC centers, defects causing losses in the region
3÷4.5 eV are not unambiguously identified. Neustruev [19] and Fujimaki et al. [106] tentatively
attributed the absorption at ∼4.5 eV to Ge(1) centers. As shown in section 7.1, our samples
reveal the presence of the Ge(1) EPR signature after radiation exposure (see Figure 7.2).
Plotting the absorption variation at 4 eV detected in the central sample layer with respect
the Ge(1) EPR intensity, we obtain the graph illustrates in Figure 7.8 (lower x-scale). It is
apparent that Ge(1) EPR signal is proportional to OA absorption around 4 eV. The linear
correlation is confirmed by the best fit function represented as a full line in Figure 7.8. This
confirms that Ge(1) is the main responsible defect for the transmission losses at 4 eV.

So, while EPR measurements do not allow to spatially resolve defect concentration, by
relating Ge(1) OA band to their EPR signal, we can locate Ge(1) presence in the central part
of the sample (the more doped part). Ge(1) concentration can be calculated considering the
central part of the sample (x-upper scale in Figure 7.8).

It is also possible to measure the Ge(1) cross section, σ, at 4 eV , defined as [50]:

σ4eV = ∆α4eV /N (7.1)

where ∆α is the variation of the absorption coefficient compared to pristine sample and N
7.3. Discussion: radiation effects 77

1 8 -3
∆[ G e ( 1 ) ] ( 1 0 c m )
-2 0 2 4 6 8 1 0 1 2
1 2

1 1

1 0

8
)
-1

7
(c me V

6
∆α4

2
-5 0 5 1 0 1 5 2 0 2 5
G e (1 ) E P R In t e n s it y ( A r b . U n it s )

Figure 7.8: Correlation between the EPR intensity of the Ge(1) signal (lower x scale) or the Ge(1)
concentration (upper x scale) and the OA absorption variation at 4 eV. Full line represents the linear
best fit curve.

represents the defect concentration.


To get the σ4eV value for Ge(1) centers, we can evaluate the slope of the correlation curve of
Figure 7.8, between Ge(1) concentration and the OA absorption variation at 4 eV. By a linear
fit of the type:
[Ge(1)] = A + σ · ∆α4eV (7.2)

we find that the cross section is σ4eV ≈6×1019 cm2 . Hence it so possible to demonstrate that
Ge(1) centers are responsible for the observed transmission losses and that they are mainly
localized in the central part of the samples where the larger germanium content favors the
electron trapping on GeO4 sites.

7.3.2 Generation processes of GECs defects

In section 7.3.1 we have seen that Ge(1) defects are concentrated in central sample zone, with
the higher Ge-content. Hence, starting from the global EPR measurements, we are able to
evaluate their effective concentration in the different preform zones. The result is illustrated
in Figure 7.9. Comparing Figures 7.5 and 7.9, we can observe that the Ge(1) concentration,
78 7. Effects of the UV and X-ray irradiation

1 0 0 0
F ib e r
1 0 0

1 0
)
3
s p in /c m

0 .1
1 8
D e fe c ts c o n c e n tr a tio n ( 1 0

0 .0 1 G e (1 )
1 1 0 1 0 0 1 0 0 0 G e E ’
1 0 0 G e (2 )
P re fo rm
1 0

0 .1

0 .0 1
1 1 0 1 0 0 1 0 0 0
X d o s e (k G y )

Figure 7.9: Concentration of the GECs centers as a function of the X-ray dose detected in (a): fiber
and (b): preform samples. Defect concentrations are evaluated as concentrated in the central sample
zones (<[GeO2 ]>∼10%).

both in fibers than in preforms, is larger than the value expected considering only the EPR
intensity trend. In particular, after an X-ray exposure of 2 MGy the Ge(1) concentration is
about 1019 cm−3 in the inner core fiber part, that is about two order of magnitude larger than
the value obtained considering the defects as equally distributed on the whole sample. For
Ge(2) the defect concentration in fibers at the saturation value is ∼1×1019 cm−3 . For the
considered total deposited doses, the GeE0 centers do not reach a saturation value: after a
total X deposited dose of 2 MGy, [GeE0 ]'4×1017 cm−3 , both in fibers and in preforms.

We note that the Ge(1) localization and the calculation of their concentration can be
indirectly applied to fiber samples. This outcome is crucial to find out the radiation effect
in modulating the optical properties of step index fibers, both the light transmission and the
refractive index change.

Regarding GECs generation mechanism, many authors have supposed the following reac-
7.3. Discussion: radiation effects 79

tion as the basic creation process of Ge(1) and Ge(2) centers in Ge-doped silica after UV or
gamma exposure:
= Ge•• + GeO4 + hν −→ (= Ge• )+ + (GeO4 )• (7.3)
A germanium lone pair center (= Ge•• ) and a 4-coordinate germanium (GeO4 ) generate, after
radiation exposure, an ionized GLPC ((= Ge• )+ ), that is a Ge(1) center, and a hole center on
a 4-coordinate germanium ((GeO4 )• ).

Evidently Equation 7.3, implies that the structure for Ge(2) center is a hole center and
not a trapped electron center. This also means that, to confirm the generation mechanism
proposed in Equation 7.3, a direct proportionality between Ge(1) and Ge(2) concentration has
to exist.

According to our measurements, the dependence of Ge(1) concentration from Ge(2), both
evaluated from EPR measurements, has the trend shown in Figure 7.10. We point out that,

1 0 0

9 0 F ib e r
P re fo rm
8 0

7 0
)
3

6 0
s p in /c m

5 0
1 8

4 0
[G e (2 )](1 0

3 0

2 0

1 0

0
0 2 4 6 8 1 0 1 2 1 4 1 6
1 8 3
[G e (1 )](1 0 s p in / c m )

Figure 7.10: Correlation between Ge(1) and Ge(2) center in fiber (open circles) and preform (full
squares) Ge-doped canonical sample after irradiation exposure.

to obtain the graph in Figure 7.10, we have considered also the Ge(2) centers as concentrated
in the central sample zone, even if rigorously we have demonstrated this assertion only for
Ge(1) centers. We justify this hypothesis on the basis of Equation 7.3: if Ge(1) and Ge(2)
generation mechanisms coincide and if we find that the Ge(1) centers are only present in the
central part of our canonical samples, this must be true also for Ge(2).
80 7. Effects of the UV and X-ray irradiation

7.3.3 The drawing effect

Up to now, only few studies investigated the influence of the drawing process on the fiber
radiation response by comparing the behaviors of fiber and preform. Most of them have
characterized the effect of varying the drawing tension, speed or temperature on the radiation
response of optical fibers [119, 118, 148]. In our study, all our canonical samples have been
made with standard MCVD drawing conditions. By comparing the concentrations of room-
temperature stable paramagnetic defects at the same doses in a preform and its corresponding
fiber, we can estimate the global influence of the drawing process on the glass sensitivity. The
obtained results are illustrated in Figure 7.11 for the Ge canonical samples.

1 0 G e (1 )
G e E ’
G e (2 )
[F ib e r ]/[P r e fo r m ]

n e g a tiv e in flu e n c e
o f d r a w in g p r o c e s s

p o s itiv e in flu e n c e

1 1 0 1 0 0 1 0 0 0
X -D o s e (k G y )

Figure 7.11: Defect concentration ratio on fibers and preforms in Ge-doped samples at different doses.

Our results clearly show that the influence of the drawing process on the generation of Ge-
related centers is dose-dependent. The dose-dependence is observed separately for the three
paramagnetic defects: Ge(1), Ge(2) and GeE0 . The impact of the drawing process is strongly
negative at low doses and it becomes nearly negligible at higher doses: at 2 MGy this ratio is
about 1 and it becomes 10 for a dose of 2 kGy. These results show that at lower doses (space,
military applications) the drawing process governs the generation of paramagnetic defects in
Ge-doped glass whereas the glass composition seems to be the most influential parameter for
high-dose applications (nuclear power plants, high energy physics).

A possible explanation for this dose dependence is that drawing strongly increases the
7.3. Discussion: radiation effects 81

number of defect precursors such as GeODC(II) and the so called neutral oxygen vacancy
(NOV) (≡ Ge − T ≡; T =Si or Ge) [19, 106] which can turn into Ge(1), Ge(2) or GeE0 under
irradiation following the process described by Equation 7.3 and by the following relation:

≡ Ge − T ≡ +hν −→= Ge• ++ T ≡ +− e (7.4)

where ≡ Ge−T ≡ is the NOV, T is either Ge or Si, and = Ge• is the GeE0 center. At low doses,
the contribution of defects generated from these precursor sites to the total concentration of
defects is predominant whereas it may become less important at higher doses due to defect
generation via other mechanisms. As above shown, our CML measurements on the Ge-doped
fiber with UV excitation provided evidence for the presence of GeODC(II) in both pristine
(see Section 6) and irradiated samples. By the way, Ge(1) and Ge(2) defects may be created
from the preexisting and X-ray radiation-induced GeODC(II). We have previously shown
that the Ge(1) and Ge(2) concentrations saturate at higher doses (>20 kGy) whereas GeE0
concentrations continuously increase up to 2 MGy (Figure 7.9) providing evidence for several
generation mechanisms for this defect.

Additional tests have to be done to fully understand the drawing influence on the fiber
radiation response. Complementary EPR measurements have to be done at lower doses to
investigate this effect for low-dose environments. Furthermore, different fiber samples have to
be drawn from the same Ge-doped preform to determine the most favorable drawing conditions
for this kind of optical fiber.
82 7. Effects of the UV and X-ray irradiation
Part IV

Influence of further dopants: fluorine and


phosphorus
Chapter 8

F-doped fibers and preforms

In this chapter we will analyze the fluorine role in changing the structure and the radiation
effects in the optical fibers. As shown in Section 2.2, there are several plausible reasons for
justifying the addition of fluorine to the a−SiO2 matrix:

• F substitutes to OH groups or Cl, always present in optical fiber structure, without


inducing any optical absorption in the transparent region of a silica glass.

• Si − F bond is stronger than Si − O, resulting in widening the optical band gap.

• Formation of Si − F bonds makes the glass network to be more stable, i.e., it decreases
structural disorder, which shift the vacuum-ultraviolet silica fundamental absorption
edge (Urbach edge) to shorter wavelengths [30].

• F favours structural relaxations, and makes it easy to fabricate silica glasses with lower
fictive temperatures (Tf ), which is defined as "the temperature at which the glass would
find itself in equilibrium if suddenly brought to that temperature from its given state"
[149].

The first part of the Chapter is devoted to the materials characterization by a micro
Raman analysis. Successively the fluorine ability of improving the material radiation hardness
is investigated. Specifically, after fiber and preform exposition to ionizing radiations of different
nature, the samples are spectroscopically analyzed by electron paramagnetic resonance (EPR)
to reveal the possible presence of point defects thus evaluating the fiber radiation resistance.
86 8. F-doped fibers and preforms

8.1 Raman results

It is known that fluorine is incorporated into the silica glass structure as SiF with the silicon
atom bonded to three oxygen atoms in the network [150]. Raman spectroscopy is an useful tool
to determine the presence of SiF linkages in silica glass, since they give rise to a spectroscopic
line around 945 cm−1 . In the following we take advantage of the spatial resolution of the
confocal micro-spectrometer elsewhere described (Section 5.3.5) to check F-presence in our
samples.

In Figure 8.1 we show the Raman spectrum of the zone 1 (see Table 4.1(b)), obtained
in the spectral region (150-1100) cm−1 . It is clearly visible the peak centered at 945 cm−1
R a m a n In te n s ity ( a r b . u n its )

S i- F
p e a k

2 0 0 4 0 0 6 0 0 8 0 0 1 0 0 0
-1
R a m a n S h if t ( c m )

Figure 8.1: Preform cladding (zone 1) Raman spectrum in the spectral range (100-1100) cm−1 . The
peak at 950 cm−1 is related to Si − F stretching mode.

corresponding to the Si − F stretching vibration mode of a SiO3 F tetrahedron [151]. The


variation of this F-related line in the other zones is plotted in Figure 8.2, all spectra being
normalized to the intensity of the peak around 800 cm−1 which is ascribed to the stretching
mode of Si − O bond [152]. Table 8.1 gives the nominal average F-concentrations for each
zone, as evaluated by electron microprobe analysis, and the ratio of the Raman 945 cm−1 band
and the 800 cm−1 reference band, for the various layers. As known from literature [27], for
low F content (inferior to 2 wt%) only Si − F links are present; this agrees with our results
that point out a correlation between F content and the Si − F Raman signal.
8.1. Raman results 87

R a m a n In te n s ity ( a r b . u n its ) D is ta n c e f r o m c e n te r
2 .3 5 m m
(z o n e 0 )
2 .1 5 m m
(z o n e 1 )
1 .6 m m
(z o n e 2 )
1 m m
(z o n e 3 )
0 .4 5 m m
(z o n e 4 )

2 0 0 4 0 0 6 0 0 8 0 0 1 0 0 0 1 2 0 0
-1
R a m a n S h if t ( c m )

Figure 8.2: Raman spectra detected in the different sample zones listed in Table 4.1. The variation of
the Si − F band (950 cm−1 ) at different distances from the preform center is clearly visible. For sake
of comparison, spectra are shifted and they are normalized to the 800 cm−1 band.

All the other peaks are intrinsic features associated with the glassy silica matrix [153].
The peaks at 495 cm−1 and 606 cm−1 in Figures 8.1 and 8.2, superposed to the larger band at
440 cm−1 , are the so-called defect bands D1 and D2. The D2 and the D1 bands are attributed
to the three- and the four-membered ring structures, respectively [154]. In these small ring
structures the Si − O − Si bond angle (130.5◦ for the three-membered ring and 160.5◦ for
the four-membered ring) [25] is significantly different from the stablest angle (150◦ ) [154,155].
As a consequence, these ring structures are composed of heavily strained S − O − Si bonds.
According to estimates by Galeener [154], the atomic fraction of these rings is of the order of
1%.

It has been demonstrated that strained Si − O − Si bonds, which are created on den-
sification of SiO2 glass by high pressure [156], cause the absorption edge shift to a longer
wavelength [25]. The densified SiO2 glasses are extremely sensitive to defect formation by ir-
radiation. The formation of strained three- and four-membered ring structures has been found
to be suppressed effectively by doping of a small amount of F because fluorine is incorporated
in the form of Si−F bonds that terminate in a continuous silica network structure [25]. This is
evident from the Raman spectra depicted in Figure 8.3, where a comparison (without vertical
shift) is reported between the F-free zone (zone 0 in Figure 4.4) and the 11 wt% F-doped zone
88 8. F-doped fibers and preforms

Table 8.1: Fluorine average concentration and ratio of the Raman bands at 945 cm−1 and at 800 cm−1
band for the different preform zones. N.D. stands for not detected.

I(945 cm−1
Zone <F> I(800 cm−1
(wt%)

Coating 0 0.0 N.D.


Cladding 1 2.5 0.14±0.03
2 1.3 0.12±0.02
Core 3 0.7 0.09±0.02
4 0.2 0.03±0.01

D 1
D is ta n c e f r o m c e n te r :
2 .1 5 m m
R a m a n In te n s ity ( a r b . u n its )

(z o n e 1 ; [F ]= 1 .8 w t% )
2 .3 5 m m
(z o n e 0 ; [F ]= 0 w t% )

D 2

2 0 0 4 0 0 6 0 0 8 0 0 1 0 0 0
-1
R a m a n S h if t ( c m )

Figure 8.3: Raman spectra of the F-free and the F-doped (11 wt%) zones of the pristine canonical
preform sample. D1 and D2 defect bands are highlighted.

(zone 4 in Figure 4.4). The main difference between the two spectra, apart from the presence
of the 945 cm−1 Si − F band, is the intensity reduction of the D1 and D2 bands. The role
of F is similar to that of OH groups in this respect. However, laser damage is drastically
enhanced for OH doping because OH groups have strong optical absorption at a wavelength of
157 nm. The breaking of the continuous silica network structure by F doping largely facilitates
structural relaxation in the process of cooling from the melt: F doping may be regarded as
chemical annealing [25].
8.2. EPR measurements on irradiated samples 89

8.2 EPR measurements on irradiated samples

As pointed out in Section 2.2, recent studies have shown that radiation toughness of silica
samples is achieved by incorporating Si−F groups whose positive effect is assumed to be the
reduction of defect precursors [150,27,24], such as strained bonds (Si − −O − −Si) from which
is likely generated the pair of silicon dangling (≡ Si• ) and oxygen dangling (≡ Si − O• ), where
(≡) and (• ) indicate bonds with three oxygen atoms and an unpaired electron, respectively.
These two paramagnetic defects, that as shown in the introduction section are also named
E0 center and non bridging oxygen hole center (NBOHC) (see Section 2.2), are indeed one of
the main causes of transparency loss due to their absorption bands peaked in visible and UV
spectral range [157]. To this aim the role of fluorine doping in the response to UV pulsed laser
and γ radiation of silica preforms and fibers was studied using EPR.

8.2.1 Results

Figure 8.4 shows the EPR spectra of preform samples after exposure to UV- (8×103 J/cm2 ;
panel (a)) and γ-rays (91 kGy; panel (b)); we observe that neither preforms nor fibers show
EPR signals before exposure. Regardless the kind of irradiation, two identical signals are
observed in our samples. The central part of the spectra shows the typical line shape of the
E0 centers, whose concentration is measured to be (5.5±0.8)×1015 cm−3 in the UV irradiated
sample, and (1.8±0.4)×1016 cm−3 in the γ irradiated one. In the spectra is also observed a
structured signal extended over 10 mT, which is ascribed to the [AlO4 ]0 center, whose structure
is shown in Figure 8.5. This defect, consisting in a substitutional Al atom [159, 158], has been
extensively studied in many experimental [160, 161] and theoretical [162, 104] studies. The
calculated concentrations of these defects are (3.9±0.7)×1015 cm−3 and (6.9±0.4)×1015 cm−3
after UV- and γ irradiation, respectively.

The presence of [AlO4 ]0 defect indicates the existence of extrinsic impurities in preform
samples before exposure. Though the EPR spectra do not allow to spatially resolve the
distribution of paramagnetic defects, as the sample core-cladding part is made of high pure
F-doped silica, we can suppose that [AlO4 ]0 center are concentrated in the external layer of
non doped silica, containing ∼10 ppm of Al impurities.

Similar irradiation treatments were performed on fiber samples; in this case, the E0 centers
could be observed only by second harmonic mode, whereas detection of [AlO4 ]0 EPR signal
is prevented by our detection limit. Figure 8.6 shows the E0 -related second-harmonic signal,
both in UV (Figure 8.6(a)) and γ (Figure 8.6(b)) irradiated samples. UV irradiations were
conducted at different amounts of energy fluence, from 10 J/cm2 to 132 J/cm2 . As clearly
shown in Figure 8.7, E0 concentration grows with the UV fluence up to 55 J/cm2 , after that
90 8. F-doped fibers and preforms

(a )
s ig n a l ( a r b . u n its )

3 3 8 3 4 0 3 4 2 3 4 4 3 4 5 .0 3 4 5 .3 3 4 6 3 4 8 3 5 0 3 5 2

(b )
E S R

3 3 8 3 4 0 3 4 2 3 4 4 3 4 5 .0 3 4 5 .3 3 4 6 3 4 8 3 5 0 3 5 2
M a g n e tic F ie ld ( m T )

Figure 8.4: EPR first harmonic spectra in UV (8×103 J/cm2 ) (a) and γ (91 kGy) (b) irradiated
preforms. The central enlarged zone shows the typical signal of E0 centers. Parts (a) and (b) have the
same scale.

Figure 8.5: Microscopic structure of the [AlO4 ]0 center. Figure adapted from ref. [158].

it saturates at a value of 8×1015 cm−3 . In γ irradiated fiber, after a dose of 91 kGy, E0


concentration becomes (2.3±0.2)×1016 cm−3 .
8.2. EPR measurements on irradiated samples 91

4 0 (a ) U V F lu e n c e
2
(J /c m )

1 3 2
2 0 5 5
3 7
s ig n a l ( a r b . u n its )

1 0

(b )
8 0
E S R

6 0

4 0

2 0

0
3 4 5 .5 3 4 5 .6 3 4 5 .7 3 4 5 .8 3 4 5 .9
M a g n e tic F ie ld ( m T )

Figure 8.6: EPR second harmonic spectra in fiber samples after irradiation. (a): variation of the
signal due to E0 centers with the UV fluence on fibers.(B): E0 signal in fibers after a γ dose of 91 kGy.

8.2.2 Discussion: generation of E0 centers

The above reported results point out two aspects, common both to preform and fibers and to
UV- and γ-irradiation:

1. The generation of E0 centers is due to the conversion of pre-existing precursors such


as oxygen vacancy or extrinsic Si-H or Si-F bonds. This is consistent with the growth
curve observed under UV irradiation, that manifests a saturating tendency due to the
exhaustion of precursors; hence, the larger E0 concentration induced by γ radiation
suggests the presence of different kind of precursor activated by one or the other radiation
source. On the other side, the absence of NBOHC, ensured by EPR spectra within a limit
of ∼1016 cm−3 , and by luminescence measurements on preform within ∼1015 cm−3 [157],
indicates that the radiolysis of strained Si-O bonds is an unlikely process in our F-doped
silica samples.
92 8. F-doped fibers and preforms

9
)
-3

6
c m
1 5
[ E ’] ( 1 0

0 2 0 4 0 6 0 8 0 1 0 0 1 2 0 1 4 0
2
F lu e n c e ( J / c m )

Figure 8.7: E0 concentration as a function of the energy fluence on UV irradiated fibers.

2. The drawing process does not weaken significantly the radiation toughness of the fiber
samples. In fact, after γ-irradiation the ratio between the number of E0 defects in
fiber (Figure 8.4(b)) and in preform (Figure 8.6(b)), both irradiated with 91 kGy, is
[E0 ]f iber /[E0 ]pref ∼
=1.2. An analogous comparison can be made for UV irradiated samples.
Considering the saturation value of E0 concentration on UV irradiated fibers, we can
evaluate the ratio between this concentration and that found in preforms after an UV
fluence of 8×103 J/cm2 : [E0 ]f iber_satur /[E0 ]pref ∼
=1.5. Hence, we can conclude that about
0
the same amount of E centers are generated in fiber and preforms both under UV and
γ radiation.

These points qualitatively evidence the role of F-doping in governing the response to
radiation of preform and fiber. In fact, irradiation of undoped silica, bulk [24] or fiber [163,164],
produces the generation of the pair NBOHC and E0 center by the rupture of strained Si-
O bonds, whereas the formation of NBOHC is inhibited in F-doped samples. This finding
suggests that Si-F groups reduce the presence of strained bonds, this role upholds both in
bulk and fiber, thus being crucial to improve the quality of optical fibers designed for visible
and UV transmission.
Chapter 9

P-doped fibers and preforms

As shown in the introduction sections (see Sections 2.3 and 3.2.2), despite the importance of
P-doping in fiber technology, its real influence in defect generation, and consequently in light
propagation attenuation, is still unknown. Not only the consequence of the radiation exposure
in terms of attenuation, but also the structure of P-related defects in pristine P-doped a−SiO2
is still subject of study and discussion.

In this chapter we report experimental studies on phosphorous-related point defects in


amorphous silica, based on photoluminescence, absorption, and electron spin resonance mea-
surements carried out on P-doped canonical fibers and preforms (see Section 4.1).

9.1 Optical activity of P-related point defects

From the experimental point of view, the optical properties of diamagnetic P-related
centers in silica are scarcely known at the moment, since only very little data exist on optical
absorption and luminescence of as-grown phosphosilicate glasses, expect for the basic evidence
that no strong UV absorption bands are generally induced in these materials just by P-
doping [38, 165].

The purpose of this section is to contribute to a better understanding of these topics, by


reporting data obtained by absorption and photoluminescence measurements in the UV and
vacuum UV spectral ranges on P-doped optical fiber preforms.

9.1.1 Absorption and photoluminescence analysis

Figure 9.1 shows the OA spectrum of the central zone of the canonical sample (zone 4, maxi-
mum P-content '7 wt%, see Figure 4.6) in the UV and in the VUV spectral range, detected
94 9. P-doped fibers and preforms

with the JASCO V-560 and the ACTON SP-150 spectrometers respectively.

6 0

5 0
1 .0

4 0
0 .5
)
-1

3 0
α(c m

0 .0
3 4 5 6
2 0

1 0

0
3 4 5 6 7
E n e rg y (e V )

Figure 9.1: Absorption spectrum of the P-doped preform in the UV-VUV spectral range (2.5-7.5 eV).
The inset shows the enlargement of the OA spectrum from 3 to 6 eV.

It is evident the presence of a strong absorption band in the VUV range centered at about
6.9 eV, an absorption tail for E>7eV, and a shoulder in the UV region, due to a band centered
at about 4.8 eV (inset of Figure 9.1).

By performing a PL emission measurement at room temperature under synchrotron exci-


tation at the energy corresponding to the UV band (4.8 eV), we detected a broad asymmetric
luminescence signal centered at 3.0 eV, reported in Figure 9.2(a). The PLE spectrum of this
signal, measured with emission at Eem =3 eV, is reported in Figure 9.2(b) and features two
components: the first one is centered at 4.7 eV (with FWHM ∼0.7 eV) and the second one at
6.4 eV (with FWHM ∼0.6 eV). The shape of the emission band turns out to be approximately
the same when excited at 4.8 eV (full squares) or 6.4 eV (open circles) by synchrotron radia-
tion, or upon laser excitation at 4.8 eV (solid line), as shown by the comparison of the three
signals in Figure 9.2(a).

We studied the dependence of the 3.0 eV luminescence signal intensity on P concentration,


by moving the laser excitation spot across the different preform zones via a micropositioning
stage. The spatial resolution of this measurement is limited by the diameter of the laser
spot, partialized by an iris, that is about 0.5 mm. As shown in Figure 9.3(a), we found
9.1. Optical activity of P-related point defects 95

E (e V )
4 5
E X
4 .8 ( la s e r e x c .)
(a ) (b ) 1 .0
4 .8

P L E In te n s ity ( a r b . u n its )
P L In te n s ity ( a r b . u n its )

6 .4 0 .8
3 0
0 .6

1 5 0 .4

0 .2
0
0 .0
2 .4 2 .7 3 .0 3 .3 4 5 6 7 8
E n e rg y (e V ) E n e rg y (e V )

Figure 9.2: (a): Emission spectra measured in the central sample zone (zone 4), containing ∼7 wt%
of phosphorus at 300 K, obtained exciting at 4.8 eV (full squares) and 6.4 eV (open circles) under
synchrotron radiation and at 4.8 eV under laser excitation with WT =20 ms and TD =5 ns (solid line).
(b): Excitation spectrum monitored at Eem =3 eV detected at room temperature under synchrotron
radiation in the range 4-8.3 eV.

that the 3.0 eV PL signal is observed only in the P-doped region of the sample, and rapidly
disappears when moving away from the center of the perform. By comparing with the spatial
dependence of P concentration, we see that the luminescent region is somewhat narrower than
the doped region. Indeed, the 3.0 eV PL is mainly localized on zones III and IV of the preform.
Figure 9.3(b) shows the PL intensity on the peak of the 3.0 eV band as a function of P-content.
It appears that the luminescence signal shows up only when P concentration overcomes a ∼4%
threshold.
96 9. P-doped fibers and preforms

8
1 .4 4 P -c o n te n t
(a ) P L I n te n s ity
1 .2

P L In te n s ity a t 3 e V ( a r b . u n its )
1 .0 6
3 3

P c o n te n t (w t% )
0 .8
2 2
0 .6
4
0 .4 1 1
0 .2

0 .0 2

-2 -1 0 1 2
D is t a n c e f r o m c e n te r (m m )

1 4

1 2
P L In te n s ity a t 3 e V ( a r b . u n its )

(b )
1 0

1 2 3 4 5 6 7 8
P c o n te n t (w t% )

Figure 9.3: (a): PL intensity at 3 eV measured at room temperature under laser excitation at 4.8 eV
with WT = 20 ms and TD =5 ns (full circles, left vertical scale) and phosphorus content (open circles,
right vertical scale) as a function of the distance from the preform center. Vertical lines refer to the
different sample zones (from 1 to 4) as listed in Table 4.1 and in Figure 4.6. (b): PL intensity at 3.0 eV
as a function of the phosphorus content.

We performed time-resolved emission measurements (Figure 9.4) on the 3.0 eV band, by


acquiring at room temperature several emission spectra upon laser excitation, with WT = 500 µs
and TD going from 1 µs to 25 ms. These data allow to study the decay kinetics of the PL
signal at several spectral positions within the emission band. Since the decay turns out to
be single-exponential at any fixed emission energy, the lifetimes were obtained by a fitting
procedure with a single exponential function of time-resolved PL data at several spectral po-
sitions, as shown in Figure 9.5, where full lines represent the best fit curves. We found a slight
dispersion of the lifetime inside the emission band: τ varies from 6.9 ms at Eem = 2.7 eV to
9.1. Optical activity of P-related point defects 97

5.1 ms at Eem = 3.2 eV. Such a dispersion of the lifetime corresponds to a progressive red shift
of the emission peak of Figure 9.4 during the decay, due to the low energy tail of the signal
decaying slower than the right tail.

Figure 9.4: Time resolved PL spectra measured under laser excitation at 4.8 eV with WT = 500 µs
and TD from 1 µs to 25 ms.

We also studied the dependence on temperature of the luminescence signal measured by


exciting at 4.8 eV in the central sample region. Figure 9.6 shows the PL spectra under laser
excitation at 4.8 eV at different temperatures. Notwithstanding a certain degree of scattering
of data points, from this investigation we can clearly see that the emission intensity excited
at 4.8 eV is poorly dependent on temperature in the range 10-300 K. (inset of Figure 9.6);
also the peak position and width do not depend significantly on temperature. The lifetime
τ measured on the peak, at 3 eV, is independent of temperature as well within experimental
accuracy, as shown in Figure 9.7 (a), where several decay curves at various temperatures are
reported. Performing the same investigation on the left tail of the band, a small variation of
the τ value with the temperature appears (full circles and triangles in Figure 9.7(b)). This
may suggest the existence of another small component centered at energies < 2.8 eV. This
secondary effect needs a specific investigation but it will not be analyzed in this Thesis. It is
worth noting, however, that the presence of this component may contribute to the observed
asymmetry of the emission band.
98 9. P-doped fibers and preforms

6
e E e x = 4 .8 e V
T = 3 0 0 K
4
e
P L In te n s ity ( a r b . u n its )
2
E e m
= 2 .7 e V
e τ= 6 .9 m s

e
0 E e m
= 2 .9 e V
τ= 5 .7 m s
-2
e E = 3 .0 e V
e m
τ= 5 .2 m s
-4
e
E e m
= 3 .2 e V
-6
τ= 5 .1 m s
e
0 5 1 0 1 5 2 0 2 5 3 0 3 5
D e c a y T im e ( m s )

Figure 9.5: Decay curves detected at room temperature at different emission energies (Eem ) under
laser excitation at Eex =4.8 eV (WT = 500 µs and TD from 1 µs to 25 ms). For viewing purposes,
the initial values of the decay curves are arbitrarily scaled. Full lines represent the best fit curves by a
single exponential equation.

3 0
5 0
3 0 0 K
1 5
1 6 0 K
P L In te n s ity ( a r b . u n its )

4 0 4 0 K

0
0 1 0 0 2 0 0 3 0 0
3 0 T (K )

2 0

1 0

0
2 .1 2 .4 2 .7 3 .0 3 .3 3 .6
E n e rg y (e V )

Figure 9.6: PL bands under laser excitation at 4.8 eV in the central sample zone detected at various
temperatures. The inset shows the temperature dependence of the integrated PL intensity.
9.1. Optical activity of P-related point defects 99

3
1 0
E E X
= 4 .8 e V

1 0
2 E E M
= 3 e V

P L In te n s ity ( a r b . u n its )
1 0
1 T = 1 0 K
τ = 5 .1 m s

0 T = 7 0 K
1 0
τ = 5 .3 m s

-1 T = 1 3 0 K
1 0
τ = 5 .3 m s

1 0
-2 (a ) T = 3 0 0 K
τ = 5 .2 m s
0 5 1 0 1 5 2 0 2 5 3 0
D e c a y T im e ( m s )

7 .0
P e a k p o s itio n
2 .7 e V
2 .9 e V
6 .5 3 .0 e V
3 .2 e V

6 .0
τ(m s )

5 .5

5 .0

(b )
4 .5
0 5 0 1 0 0 1 5 0 2 0 0 2 5 0 3 0 0
T (K )

Figure 9.7: (a): Decay curves detected at Eem =3 eV at different temperatures under laser excitation
at Eex =4.8 eV (WT = 500 µs and TD from 1 µs to 25 ms). For viewing purposes, the initial values of
the decay curves are arbitrarily scaled. Full lines represent the best fit curves by a single exponential
equation. (b): Temperature lifetimes dependency at different energetic positions inside the 3.0 eV-PL
band at Eex = 4.8 eV under laser excitation.

Additionally, we performed EPR measurements both on pristine and X-ray irradiated


preform samples aiming to explore the local arrangements of P atoms in the SiO2 matrix, as
well as the radiochemical processes activated by X irradiation. The pristine samples did not
show any detectable paramagnetic defects. After a total deposited 10 keV-X-dose of 20 kGy,
at a dose rate = 0.1 kGy/s, the EPR spectrum shows a structured signal extended over 10 mT,
and shown in Figure 9.8-(a). The measurement was performed with a modulation amplitude
of 0.1mT and a non-saturating microwave power of 0.19 mW. By comparison with literature,
most of the observed signal can be ascribed to POHC-type defects [40]. Specifically, the
100 9. P-doped fibers and preforms

1
(a )
0

-1

-2 (1 )
s ig n a l ( a r b . u n its )

(2 )

-3
3 3 8 3 4 0 3 4 2 3 4 4 3 4 6 3 4 8 3 5 0
2

(b )
E S R

1 1 2 m T
1

-1

-2
2 5 0 3 0 0 3 5 0 4 0 0
M a g n e tic F ie ld ( m T )

Figure 9.8: Electron spin resonance spectrum detected in P-doped preform after a total deposited
X-dose of 20 kGy. The signal in panel (a) was acquired by using a modulation amplitude of 0.1 mT
and a 0.19 mW power. (1) and (2) indicate characteristic features of r-POHC and l-POHC respectively.
The spectrum in panel (b) was acquired by using a modulation amplitude of 0.5 mT and a 1.9 mW
power. We verified both values of power to be below the saturation threshold of the respective signals.

negative peaks at 341.9 mT and 346.5 mT (marked as (1)) are associated to r-POHC, while
the two additional peaks at 342.8 mT and 347.3 mT (marked as (2)) are characteristic features
of l-POHC. The positive portions for B<341.5 mT and for 345.5mT<B<346.2mT are due to
a combination of l-POHC and r-POHC. Both signals are actually hyperfine doublets with
∼5mT splitting, due to the interaction with the 31 P nucleus (100% natural abundance). It
is worth noting that the observation of l-POHC contrasts with previous suggestions of this
center being metastable at room temperature [40]. Finally, the central part of the spectrum
9.2. Discussion: luminescent P-defects structure 101

also shows an additional structure around 344 mT that is similar to a signal attributed to
the so called Si(E0 )(P) center [40], a type of Si(E0 ) center with phosphorus next-nearest-
neighbors (see section 3.2.2). Since the contribution of the latter defect to the overall signal is
minor, the overall concentration of POHCs can be estimated in good approximation by double
integration of the whole spectrum in Figure 9.8-(a) and comparison with a reference sample.
In this way we get: [POHC]=(1.2±0.1)×1017 cm−3 . a By scanning a wider magnetic field range
(Figure 9.8-(b)), we observe a hyperfine doublet with 112 mT separation which is consistent
with the spectral features of the P2 defect, i.e. a 4-fold coordinated P trapping an unpaired
electron [40]. The substantially larger hyperfine splitting as compared to POHC is due to the
much stronger localization of the unpaired electron on the P nucleus. This measurement was
carried out with a modulation amplitude of 0.5 mT and a non-saturating microwave power of
1.9 mW. The concentration of P2 centers can be estimated to be [P2]=(1.4±0.1)×1017 cm−3
and thus is consistent with [POHC] within experimental error. Finally, by PL measurements in
this sample we observed a 42% reduction of the intensity of the native PL band at 3.0 eV. b The
same measurements were performed on another sample irradiated with a lower (2 kGy) dose,
where we found similar results: the emission intensity was reduced of 30%, while from EPR
measurements we get: [POHC]=[P2]=(8.5±0.8)×1016 cm−3 . In neither sample we observed
the characteristic hyperfine doublets of P1 or P4 centers [40, 111]. This suggests that in these
materials the 2-fold and 3-fold coordinated configurations of P impurities are either present
in much lower concentration or are much less sensitive to radiation than the tetrahedrical
[(O-)3 P=O]0 and [(O-)2 P(-O)2 ]+ configurations acting as precursors for POHC and P2.

9.2 Discussion: luminescent P-defects structure

Data in Figures 9.2-9.7 demonstrate the existence in as-grown P-doped silica of an emission
signal peaked at 3.0 eV and featuring two excitation channels at 4.8 eV and 6.4 eV. The lowest
energy excitation is detectable as well as a weak absorption band. The spatial dependence of
the signal intensity in Figure 9.3 strongly suggests this emission to be associated to a P-related
defect. The fact that the spatial extension of the luminescent region is narrower than that
of the P-doped zone implies that the concentration of the specific center responsible for this
emission is not strictly linearly correlated with the overall P concentration (see Figure 9.3(b)).
a
This estimate is based on the assumption that P-related paramagnetic centers are uniformly distributed in
all the P-doped volume. If one assumes instead, following the results of Figure 9.3-(b), that the paramagnetic
centers are located in zones III and IV only (as the luminescence signal is), a correction factor must be applied,
leading to a higher estimate: (3.2±0.3)×1017 cm−3 . Under this hypothesis, all the other concentration values
reported later on must be scaled for the same factor.
b
The bleaching of the PL signal is accompanied by the appearance of new absorption signals which will be
discussed in the next section. It is worth noting that the induced absorption at 4.8 eV (∼0.1 optical density)
is not sufficiently intense to compromise the measurement of luminescence intensity.
102 9. P-doped fibers and preforms

Specifically, data in Figure 9.3 suggest the emitting center to be formed in detectable concen-
trations only when the overall P content overcomes a threshold of about 4%.

The main spectroscopic features of this PL signal are its long lifetime in the ms range and
its weak dependence on temperature. The first result suggests the emission to originate from
a spin-forbidden transition from an excited triplet state (T1 ). At the same time, the tempera-
ture independence implies the absence of efficient non-radiative decay channels from T1 to the
ground state, so that the ∼6 ms lifetime has to be interpreted as a purely radiative decay life-
time. Finally, the dispersion of the radiative decay lifetime within the emission band suggests
strong inhomogeneity effects to affect the overall width of the band, as recently pointed out for
other defects embedded in an amorphous matrix [166]. Data analysis aimed to quantitatively
estimate the degree of inhomogeneity of the band is in progress, being complicated by the
possible presence of another weaker signal on the left side of the band.

The first excitation peak at 4.8 eV detected by the PLE measurement is likely to be
related to the 4.8 eV band found in the OA spectrum of the defect (Figure 9.1). However, it
can be argued that the 3.0 eV luminescent band cannot be the inverse transition of the 4.8 eV
absorption.

To demonstrate this assertion, we evaluate the oscillator strength f of the 4.8 eV peak as
if the 3.0 eV was its inverse transition. Using Smakula’s equation [50] we evaluate the product
N f , where N is the defects concentration :
 2  2
E0  mc 
e E0
Nf = n αmax Γ∆ ≈n αmax Γ∆ × 9.111 × 1015 [eV −1 cm−2 ] (9.1)
Eef f 2π 2 e2 ~ Eef f
n is the refractive index of our samples, (Eef f /E0 )2 is the effective field correction [50], αmax
is the amplitude and ∆ is the full with at half maximum (FWHM) of OA band, Γ is a numeric
coefficient which depends on the bandshape and me , c, ~ have their usual meaning.
If we consider that Γ ≈ 1.0645 for a gaussian shape, and n(E0 /Eef f )2 for silica is close to
unity throughout the IR to near-UV spectral range, eq. 9.1 becomes:

N f ≈ αmax ∆ × 9.699 × 1015 [eV −1 cm−2 ] (9.2)


To evaluate the product αmax ∆, we performed a gaussian fit of the OA peak at ∼4.8 eV of the
inset of Figure 9.1, imposing the same FWHM of the 4.8eV PLE band (0.7 eV) (Figure 9.2(b)),
and finding αmax = 0.10 cm−1 . From this calculation we obtain:

αmax ∆ ' 0.07eV cm−1 (9.3)

N f = 6.79 × 1014 cm−3 (9.4)


The oscillator strength is related to the radiative decay time τ of the inverse emission by [50]:
 2
1 E0 1 −8 1 1
f≈ × 2.305 × 10 ≈ × × 2.305 × 10−8 = 1.18 × 10−7 (9.5)
(~ω)2 τ Eef f n (~ω)2 τ n2
9.2. Discussion: luminescent P-defects structure 103

where n = 1.466 and ~ω represents the zero phonon line energy position, roughly half way
between absorption and emission: ~ω = 3.7 eV.
From eq. 9.4 and 9.5, with τ =6 ms, we finally obtain:

N ' 6 × 1021 cm−3 (9.6)

This concentration value is higher than the maximum phosphorus concentration in our samples
(7.09 wt% in zone 4 corresponds to [P] = 3.0×1021 cm−3 ). Hence, we consider really improbable
for the 3.0 eV PL band to be the inverse transition of the 4.8 eV OA. On the contrary, assuming
a non-degenerate singlet ground state S0 for the defect, the 4.8 eV band must arise from an
allowed S0 →S1 transition. Also, the second PLE peak at 6.4 eV, whose intensity is comparable
to the 4.8 eV component, must be associated to a S0 →S2 transition to an upper excited
state. Given the absence of any other emission signals upon 4.8 eV or 6.4 eV excitations,
the decay from these two excited levels (S1 , S2 ) to the emitting 3.0 eV state (T1 ) must occur
by a very efficient intersystem crossing (ISC) process active at all temperatures. Based on
these considerations, we can propose for the P-related luminescent defect the level scheme in
Figure 9.9.

Figure 9.9: Electronic levels scheme of the diamagnetic P-related defect supposed to be at the origin
of the observed optical activity. Solid arrows indicate the radiative transitions in absorption and lumi-
nescence. Dashed arrows indicate the Inter System Crossing (ISC) non-radiative transitions from S1 or
S2 to T1 .

Contrastingly, the 6.9 eV absorption band is due to another diamagnetic P-related defect,
unrelated to the emitting center. As a matter of fact, the 6.4 eV peak in the PLE signal occurs
in a spectral region where the absorption is almost zero (compare Figure 9.1 and Figure 9.2),
104 9. P-doped fibers and preforms

while the excitation of the sample at 6.9 eV did not result in any measurable emission at any
temperature. The 6.4 eV OA band which, based on PLE data, one may expect to find in the
OA spectrum with comparable intensity to the 4.8 eV, is likely to be buried under the intense
6.9 eV peak.

Indications on the microscopic structure of the luminescent P-related defect can be drawn
from the explorative EPR measurements described in the previous section. Several evidences
in literature lead to the assumption that in phosphosilicate glass most of P atoms are arranged
as [(O−)3 P = O]0 tetrahedra [115,114,116,113]. The other possible defective arrangements of
P in silica have been proposed based on EPR data, and are expected to be present only in minor
concentrations. EPR data on each of the two irradiated preforms reveal POHC and P2 defects
induced by X-rays in about the same concentration. This finding is consistent with a scheme
in which the formation of the two paramagnetic defects is correlated, and occurs by trapping
on Si-substitutional P centers [(O−)2 P (−O)2 ]+ of the electron made available by ionization
of [(O−)3 P = O]0 c . Similar correlated POHC-P2 formation under laser irradiation has been
observed in a recent study on P-doped SiO2 glass [38]. In this context, the concurrent obser-
vation of a partial bleaching of the 3.0 eV emission band upon irradiation, leads to tentatively
identify the luminescent center with one of the two precursors, i.e. the most likely microscopic
structure of the emitting diamagnetic defect is either [(O−)3 P = O]0 or [(O−)2 P (−O)2 ]+ .
A 42% reduction of the PL accompanied by the formation of (1.2±0.1)×1017 cm−3 param-
agnetic defects implies a concentration of (2.8±0.3)×1017 cm−3 for the luminescent center in
the unirradiated sample d . Comparing with [P]=3.0×1021 cm−3 in zone 4, we see that the the
[(O−)3 P = O]0 model is unlikely, since literature data strongly suggest that this structure
should be present in concentrations much higher than 1017 cm−3 , and possibly close to the
total P content. This conclusion is consistent also with the lack of a strict linear correlation
between the overall concentration of P and that of the luminescent defect (Figure 9.3(b)).
Hence, present data suggest the diamagnetic Si-substitutional P impurity, [(O−)2 P (−O)2 ]+ ,
as a tentative microscopic model of the luminescent center. Our model of the emitting de-
fect and of the conversion process activated by X radiation is represented in Figure 9.10. By
comparing the concentration inferred above with the intensity of the 4.8 eV band, one can
estimate the oscillator strength f∼2×10−3 of the S0 ->S1 transition. Computational studies
may help now to find out if these optical properties are consistent with those expected for a
Si-substitutional P impurity.

c
The observation of both l-POHC and r-POHC, according to Griscom’s model, may slightly modify this
interpretation by suggesting that a significant fraction of ionized P sites feature two doubly-bonded non-
bridging O atoms instead of one.
d
This line of reasoning still holds if one uses the higher estimate of the concentration of paramagnetic
defects obtained by assuming them to be present only in zones III and IV
9.3. Conclusions 105

o e- o
+
o P o o P o

o o

Luminescent
Center

o o
+
o P o o P o

o o

POHC P2

Figure 9.10: Graphical representation of the photochemical processes activated by X-rays in the P-
doped preform sample, and of the model proposed (upper right structure) for the luminescent defect
responsible for the 3.0 eV emission band.

9.3 Conclusions

By luminescence measurements on step-index P-doped SiO2 fiber preforms, carried out under
excitation by laser and synchrotron UV and VUV light, we detect an optical activity consisting
in a long-lived (∼6 ms radiative lifetime) and temperature-independent luminescence emission
peaked at 3.0 eV, featuring two excitation bands centered at 4.8 eV and 6.4 eV. The 4.8 eV
transition can be also revealed as a weak band in the optical absorption spectrum of the
preform. The spatial profile of the PL intensity on the preform is consistent with that of
P doping, thus allowing the attribution of this optical activity to a P-related point defect.
The detailed study of the spectroscopic features of the defect allows to propose a scheme of its
electronic transitions, comprising two singlet (S1 , S2 ) and one triplet (T1 ) excited levels, where
the long-lived emission is due to the spin-forbidden transition to the ground state from the
excited triplet populated by very efficient non-radiative decay from S1 and S2 . PL and EPR
measurements on X-irradiated samples allow to propose a microscopic model for the defect
106 9. P-doped fibers and preforms

consisting in a 4-coordinated diamagnetic P impurity substitutional to Si atoms in the SiO2


matrix.

We point out that the acquired data will be used for future multiscale simulation besides
giving important contribution for the study of all silica P-doped glasses.
Conclusions

In this Thesis we have reported an experimental study on three types of canonical fibers
and their related preforms doped with germanium, fluorine and phosphorus.

Results on Ge-doped samples could be grouped in two main points: i) the effect of the
drawing process and ii) the influence of the radiation exposure

i) We have checked the germanium lone pair center (GLPC) distribution in fibers and
preforms using confocal microscopy luminescence technique. We demonstrated that the
GLPC radial distribution is different in fibers and preforms. This difference can change
the photosensitivity of the fiber regarding the preform sensitivity. This effect could be
due to a mechanical stress introduced by the drawing process.

ii) Different paramagnetic defect species, like GeE0 , Ge(1) and Ge(2), were induced by ra-
diation exposure (X-10 keV and γ) and they were revealed by electron paramagnetic
resonance measurements. Their concentration was studied as a function of the irradi-
ation dose. The comparison with the optical absorption spectra points out the main
role of Ge(1) on the optical transmission loss of fibers in the UV region (below 5 eV)
allowing their localization in the central part of canonical samples containing an higher
concentration of GeO2 .

Regarding the study on F-doped samples, in this Thesis we focused the attention on one
main aspect: the influence of the radiation exposure on the generation of paramagnetic species.
Multiple step preforms and fibers, doped with different F concentration, were studied to find
out the effects of UV- and γ-irradiation. Regardless the samples, EPR spectra evidence that
the E0 center is the main defect induced by the conversion of precursors, whereas the absence
of EPR signal of NBOHC rules out the occurrence of radiolysis of strained Si − O bonds.
These results prove the effectiveness of Si − F group in reducing the generation of defects in
silica, also after drawing, thus improving the optical transmission of F-doped core silica fibers.

Phosphorus doped samples in this study have been the subject of specific results, con-
sidering the lack of peculiar literature. P-related point defects in amorphous silica, based on
photoluminescence, absorption, and electron paramagnetic resonance measurements has been
108 Conclusions

investigated. By photoluminescence measurements excited by laser or synchrotron light we


detected an emission band peaked at 3.0eV with a lifetime in the range of ms. The excitation
spectrum of the 3.0 eV emission consists of two transitions peaked at 4.8eV and 6.4eV, the
former giving rise also to a measurable absorption band. We attribute this optical activity
to a P-related point defect embedded in SiO2 , based on the spatial correlation between the
emission intensity and the P doping level. A detailed spectroscopical investigation allows us to
propose a scheme of the electronic levels of this P-related defect, in which the 4.8eV and 6.4eV
excitation channels arise from transitions from the ground to two excited singlet states, while
the long-lived 3.0eV emission is associated to a spin-forbidden transition from an excited triplet
to the ground state. Finally, electron spin resonance measurements on X-irradiated samples
lead us to propose a tentative microscopic model of the defect as a diamagnetic 4-coordinated
P impurity substitutional to a Si atom.

Results on Ge-doped samples form the main part of our experimental results, entering a
wide scientific literature field. Otherwise, results on F-doped and P-doped fibers and preforms
are more innovative from the experimental and technological points of view, so interesting
questions remain open and could be investigated by further experiments.
List of related papers

1. G. Origlio, S. Girard, F. Messina, M. Cannas, A. Boukenter, R. Boscaino and Y. Ouer-


dane, 10 keV X-ray irradiation effects on phosphorus-doped fibers and preforms: electron
spin resonance and optical studies, submitted for publication in IEEE Transactions on
Nuclear Science

2. G.Origlio, F. Messina, M. Cannas, R. Boscaino, S. Girard, A. Boukenter and Y. Ouer-


dane, Optical properties of phosphorus-related point defects in silica fiber preforms, Phys.
Rev. B., in press

3. G. Origlio, M. Cannas, S. Girard, R. Boscaino, A. Boukenter, , and Y. Ouerdane, Influ-


ence of the drawing process on the defect generation in multistep index germanium-doped
optical fibers, Optics Lett. 34, 2282 (2009)

4. G. Origlio, S. Girard, M. Cannas, Y. Ouerdane, R. Boscaino, and A. Boukenter, Para-


magnetic germanium-related centers induced by energetic radiation in optical fibers and
preforms, J. Non-Cryst. Solids 355, 1054 (2009)

5. S. Girard, C. Marcandella, G. Origlio, Y. Ouerdane, A. Boukenter, and J-P. Meunier,


Radiation-induced defects in fluorine-doped silica-based optical fibers: Influence of a pre-
loading with H2 , J. Non-Cryst. Solids 355, 1089 (2009)

6. G. Origlio, S. Girard, M. Cannas, A. Boukenter, R. Boscaino, and Y. Ouerdane. Optical


and photonic material hardness for energetic environments. 9me Colloque sur les Sources
Cohérentes et Incohérentes UV, VUV et X (UVX), S. Jacquemot, A. Klisnick and T.
Ruchon Eds. 127, EDS Sciences.

7. S. Girard, Y. Ouerdane, G. Origlio, C. Marcandella, A. Boukenter, N. Richard, J. Baggio,


P. Paillet, M. Cannas, J. Bisutti, J-P. Meunier, and R. Boscaino, Radiation effects on
silica-based preforms and optical fibers - I: Experimental study with canonical samples,
IEEE Transactions on Nuclear Science 55, 3473 (2008)

8. S. Girard, N. Richard, Y. Ouerdane, G. Origlio, A. Boukenter, L. Martin-Samos, P.


Paillet, J-P. Meunier, J. Baggio, M. Cannas, and R. Boscaino, Radiation effects on silica-
110 List of related papers

based preforms and optical fibers - II: Coupling ab initio simulations and experiments,
IEEE Transactions on Nuclear Science 55, 6 (2008)

9. G. Origlio, A. Boukenter, S. Girard, N. Richard, M. Cannas, R. Boscaino, and Y. Ouer-


dane, Irradiation induced defects in fluorine doped silica, Nucl. Instrum. Methods B,
266, 2918 (2008)

10. G. Origlio, S. Girard, R. Boscaino, A. Boukenter, M. Cannas, and Y. Ouerdane, Optical


and photonic material hardness for energetic environments, 9me Colloque sur les Sources
Cohérentes et Incohérentes UV, VUV et X (UVX), (2008)

11. M. Cannas and G. Origlio, Ultraviolet optical properties of silica controlled by hydrogen
trapping at Ge-related defects Phys. Rev. B, 75, 233201 (2007)
List of communications to congresses

1. G. Origlio, S. Girard, F. Messina, M. Cannas, A. Boukenter, R. Boscaino and Y. Ouer-


dane. 10-keV X-ray irradiation effects on phosphorus doped fibers and preforms: elec-
tron spin resonance and optical studies. 10th European Conference on Radiation Effects
on Components and Systems (RADECS). Bruges, Belgium, September 2009.

2. G. Origlio, S. Girard, A. Boukenter, C. Marcandella, M. Cannas and Y. Ouerdane


Influence du rayonnement X sur des fibres et préformes canoniques dopées au phosphore.
Journées Nationales d’Optique Guidée et Horizons de l’Optique (JNOG), Lille, France,
July 2009.

3. G. Origlio, S. Girard, M. Cannas, A. Boukenter, R. Boscaino, and Y. Ouerdane. Dur-


cissement de matériaux pour l’optique et la photonique destinés à l’utilisation dans un
environnement énergétique. 9me Colloque sur les Sources Cohérentes et Incohérentes
UV, VUV et X (UVX), Dourdan, France, October 2008.

4. G. Origlio, S. Girard, M. Cannas, Y. Ouerdane, R. Boscaino, and A. Boukenter. Para-


magnetic germanium-related centers induced by energetic radiation in silica based de-
vices. The 7th symposium SiO2 , advanced dielectrics and related devices, Saint-Etienne,
France, July 2008.

5. S. Girard, C. Marcandella, G. Origlio, Y. Ouerdane, A. Boukenter, and J-P. Meunier.


Radiation-induced defects in fluorine-doped silica-based optical fibers: Influence of the
H2 -loading. The 7th symposium SiO2 , advanced dielectrics and related devices, Saint-
Etienne, France, July 2008.

6. . Girard, N. Richard, Y. Ouerdane , G. Origlio, A. Boukenter, L. Martin-Samos, P.


Paillet, J.-P. Meunier, M. Cannas, R. Boscaino. Radiation-induced effects in silica-
based glasses: experimental and theoretical results, The 7th symposium SiO2 , advanced
dielectrics and related devices, Saint-Etienne, France, July 2008.

7. S. Girard, Y. Ouerdane, G. Origlio, C. Marcandella, A. Boukenter, N. Richard, J. Baggio,


P. Paillet, M. Cannas, J. Bisutti, J-P. Meunier, and R. Boscaino. Radiation effects on
silica-based preforms and optical fibers - I: Experimental study with canonical samples.
112 List of communications to congresses

The Nuclear and Space Radiation Effects Conference (NSREC), Tucson, Arizona, July
2008.

8. S. Girard, N. Richard, Y. Ouerdane, G. Origlio, A. Boukenter, L. Martin-Samos, P.


Paillet, J-P. Meunier, J. Baggio, M. Cannas, R. Boscaino. Radiation Effects on Silica-
Based Preforms and Optical Fibers - II: Coupling Ab initio Simulations and Experiments,
The Nuclear and Space Radiation Effects Conference (NSREC), Tucson, Arizona, July
2008.

9. G. Origlio, A. Boukenter, M. Cannas, S. Girard, R. Boscaino, and Y. Ouerdane. Photo-


luminescence properties of point defects in Ge-doped fibers and preforms. The 15th Inter-
national Conference on Luminescence and Optical Spectroscopy of Condensed Matter
(ICL), Lyon, France, July 2008.

10. G. Origlio, A. Boukenter, S. Girard, N. Richard, M. Cannas, R. Boscaino, and Y. Ouer-


dane. Irradiation induced defects in fluorine doped silica. Radiation Effects in Insulators
(REI), Caen, France, August 2007.
Bibliography

1. G. P. Agrawal, Fiber optic communication systems (John Wiley & Sons, New York,
1997), 2nd edn.
2. K. Nagayama, M. Kakui, M. M. I. Saitoh, and Y. Chigusa, “Ultra-low-loss
(0.1484 dB/km) pure silica core fibre and extension of transmission distance,” Electron.
Lett. 38, 1168 (2002).
3. S. Ungar, Fibre optics: theory and applications (John Wiley & Sons, New York, 1990).
4. C. Yeh, Handbook of fiber optics: theory and applications (Academic Press, San Diego,
1990).
5. D. K. Mynbaev and L. L. Scheiner, Fiber-Optic Communications Technology (Prentice
Hall, 2000).
6. G. Keiser, Optical Fiber Communications (McGraw-Hill, Nex York, 1991), 2nd edn.
7. P. C. Becker, N. A. Olsson, and J. R. Simpson, Erbium-Doped Fiber Amplifiers, Fun-
dameltal and Technology (Academic Press, London, 1999).
8. H. Murata, Handbook of Optical Fibers and Cables (CRC Press, 1996), 2nd edn.
9. K. O. Hill, Y. Fujii, D. C. Johnson, and B. S. Kawasaki, “Photosensitivity in otpical
fibers waveguides: Application to reflection fiber fabrication,” Appl. Phys. Lett. 32, 647
(1978).
10. G. Meltz, W. W. Morey, and W. H. Glenn, “Formation of Bragg gratings in optical fibers
by a transverse holographic method,” Opt. Lett. 14, 823 (1989).
11. D. P. Hand and P. S. J. Russell, “Photoinduced refractive-index xhanges in germanosili-
cate fibers,” Opt. Lett. 15, 102 (1990).
12. N. Ashcroft and N. D. Mermin, Solid State Physics (Holt-Saunders International Edi-
tions, JAPAN, 1981).
13. R. M. Atkins, V. Mizrahi, and T. Erdogan, “Induced vacuum UV spectral changes in
optical fibre preforme cores: support for a colour centre model of photosensitivity,” Electr.
Letters 29, 385 (1993).
114 Bibliography

14. M. Essid, J. L. Brebner, J. Albert, and K. Awazui, “Ion implantation induced photosen-
sitivity in Ge-doped silica: Effect of induced defects on refractive index changes,” Nucl.
Instr. and Meth. In Phys. Res. B 141, 616 (1998).
15. H. S. Nalwa, Silicon-based Materials and Devices (Academic Press, USA, 2001).
16. M. Parent, J. Bures, S. Bures, and J. Lapierre, “Proprietes de polarisation des reflecteurs
de Bragg induits par photosensibilite dans les fibres optiques monomodes,” Appl. Optics
24, 354 (1985).
17. U. Osterberg and W. Margulis, “Dye laser pumped by Nd:YAG laser pulses frequency
doubled in a glass optical fiber,” Opt. Lett. 11, 516 (1986).
18. U. Osterberg and W. Margulis, “Experimental studies on efficient frequency doubling in
glass optical fibers,” Opt. Lett. 12, 57 (1987).
19. V. B. Neustruev, “Colour centres in germanosilicate glass and optical fibres,” J. Phys.
Condens. Matter 6, 6901 (1994).
20. D. formation in SiO2 :GeO2 glasses studied by irradiation with excimer laser light, “H.
Hosono and H. Kawazoe and J. Nishii,” Phys. Rev. B 53, R11 921 (1996).
21. M. Cannas and G. Origlio, “Ultraviolet optical preperties of silica controlled by hydrogen
trapping at Ge-related defects,” Phys. Rev. B 75, 233 201 (2007).
22. S. Girard, J. Keurinck, Y. Ouerdane, J.-P. Meunier, and A. Boukenter, “γ Rays and
Pulsed X-Ray Radiation Responses of Germanosilicate Single-Mode Optical Fibers: In-
fluence of Cladding Codopants,” J. Lightwave Technol. 22, 1915 (2004).
23. J. W. Fleming and D. L. Wood, “Refractive index dispersion and related properties in
fluorine doped silica,” Appl. Optics 22, 3102 (1983).
24. K. Kajihara, Y. Ikuta, M. Oto, M. Hirano, L. Skuja, and H. Hosono, “UV-VUV laser
induced phenomena in SiO2 glass,” Nucl. Instr. and Meth. In Phys. Res. B 218 (2004).
25. H. Hosono, M. Mizuguchi, and L. Skuja, “Fluorine-doped SiO2 glasses for F2 excimer
laser optics: fluorine content and color-center formation,” Opt. Lett. 24, 1549 (1999).
26. A. A. Langford, A. H. Mahan, M. L. Fleet, and J. Bender, “Effect of fluorine on the
structural and electronic properties of a-Si:H:F,” Phys. Rev. B 41, 8359 (1990).
27. R. E. Youngman and S. Sen, “The nature of fluorine in amorphous silica,” J. Non-Cryst.
Solids 337, 182 (2004).
28. H. Hosono and Y. Ikuta, “Interaction of F2 excimer laser with SiO2 glasses: Towards
the third generation of synthetic SiO2 glasses,” Nucl. Instr. and Meth. In Phys. Res. B
166-167, 691 (2000).
29. M. Oto, S. Kikugawa, T. Miura, M. Hirano, and H. Hosono, “Fluorine doped silica glass
fiber for deep ultraviolet light,” J. Non-Cryst. Solids 349, 133 (2004).
Bibliography 115

30. L. Skuja, K. Kajihara, Y. Ikuta, M. Hirano, and H. Hosono, “Urbach absorption edge of
silica: reduction of glassy disorder by fluorine doping,” J. Non-Cryst. Solids 345 & 346,
328 (2004).
31. K. Saito and A. J. Ikushima, “Effects of fluorine on structure, structural relaxation, and
absorption edge in silica glass,” J. Appl. Phys. 91, 4886 (2002).
32. G. Cody, T. Tiedje, B. Abeles, B. Brooks, and Y. Goldstein, “Disorder and the Optical-
Absorption Edge of Hydrogenated Amorphous Silicon,” Phys. Rev. Lett. 47, 1480 (1981).
33. R. E. Youngman and S. Sen, J. Non-Cryst. Solids 337, 182 (2004).
34. D. Ehrt, P. Ebeling, and Natura, “UV Transmission and radiation-induced defects in
phosphate and fluoride-phosphate glasses,” J. Non-Cryst. Solids 263& 264, 240 (2000).
35. E. M. Dianov, M. V. Grekov, I. A. Bufetov, S. A. Vasiliev, O. I. Medvedkov, V. G.
Plotnichenko, V. V. Koltashev, A. V. Belov, M. M. Bubnov, S. L. Semjonov, and A. M.
Prokhorov, “CW high power 1.24 µm and 1.48 µm Raman lasers based on low loss
phosphosilicate fibre,” Electron. Lett. 33, 1542 (1997).
36. M. C. Paul, D. Bohra, A. Dhar, R. Sen, P. K. Bhatnagar, and K. Dasgupta, “Radiation
response behavior of high phosphorous doped step-index multimode optical fibers under
low dose gamma irradiation,” J. Non-Cryst. Solids 355, 1496 (2009).
37. B. Malo, J. Albert, F. Bilodeau, T. Kitagawa, D. C. Johnson, K. O. Hill, K. Hattori,
Y. Hibino, and S. Gujrathi, “Photosensitivity in phosphorus-doped silica glass and optical
waveguides,” Appl. Phys. Lett. 65, 394 (1994).
38. H. Hosono, K. Kajihara, M. Hirano, and M. Oto, “Photochemistry in phosphorus-doped
silica glass by ArF excimer laser irradiation: Crucial effect of H2 loading,” J. of Appl.
Phys. 91, 4121 (2002).
39. J. Bisutti, S. Girard, and J. Baggio, “Radiation effects of 14 MeV neutrons on ger-
manosilicate and phosphorus-doped multimode optical fibers,” J. Non-Cryst. Solids 353,
461 (2007).
40. D. L. Griscom, E. J. Friebele, and K. J. Long, “Fundamental defect centers in glass:
Electron spin resonance and optical absorption studies od irradiated phosphorus-doped
silica glass and optical fibers,” J. of Appl. Phys. 54, 3743 (1983).
41. E. Regnier, I. Flammer, S.Girard, F. Gooijer, F. Achten, and G. Kuytand, “Low-Dose
Radiation-Induced Attenuation at InfraRed Wavelengths for P-Doped, Ge-Doped and
Pure Silica-Core Optical Fibres,” in IEEE Transactions on Nuclear Science, 54, ed., p.
1115 (2007).
42. H. S. Nalwa, Silicon-Based Materials and Devices (Academic, San Diego, 2001).
43. S. F. Paul, J. L. Goldstein, R. D. Durst, and R. J. Fonck, “Effect of highenergy neutron
flux on fiber optics in an active diagnostic on TFTR,” Rev. Sci. Instrum. 66, 1252 (1995).
116 Bibliography

44. K. Obara, S. Kakudate, K. Oka, E. Tada, Y. Morita, and M. Sei, “Development of optical
components for in-vessel viewing systems used for fusion experimental reactor,” in SPIE
2425, 115 (1994).
45. M. O. Zabezhailov, A. L. Tomashuk, I. V. Nikolin, and K. M. Golant, “The Role of
Fluorine-Doped Cladding in Radiation-Induced Absorption of Silica Optical Fibers,” in
IEEE Transactions on Nuclear Science 49, 1410 (2002).
46. S. Girard, Analyse de la réponse des fibres optiques soumise à divers environnements
radiatifs, Ph.D. thesis, University Jean Monnet, Saint-Etienne (2003).
47. H. Hosono and K. Kawamura, “Interaction of the SiO2 glasses with high energy ion beams
and vacuum UV excimer laser pulses,” in Defects in SiO2 and Related Dielectrics: Science
and Technology, G. Pacchioni, L. Skuja, and D. L. Griscom, eds., p. 213 (2000).
48. F. M. Gelardi and S. Agnello, “Gamma rays induced conversion of native defects in nat-
ural silica,” in Defects in SiO2 and Related Dielectrics: Science and Technology, G. Pac-
chioni, L. Skuja, and D. L. Griscom, eds., p. 285 (2000).
49. K. C. Kao and G. A. Hockham, “Dielectric-fibre surface waveguide for optical frequen-
cies,” in Proc. of IEEE 113 (1966).
50. L. Skuja, “Optical properties of defects in silica,” in Defects in SiO2 and Related Di-
electrics: Science and Technology, G. Pacchioni, L. Skuja, and D. L. Griscom, eds., p. 73
(2000).
51. L. Skuja, “Optically active oxygen-deficiency-related centers in amorphous silicon diox-
ide,” J. Non-Cryst. Solids 239, 16 (1998).
52. D. L. Griscom, “γ-Ray-induced visible/infrared optical absorption bands in pure and
F-doped silica-core fibers: are they due to self-trapped holes?” J. of Non-Cryst. Solids
349, 139 (2004).
53. A. L. Tomashuk, K. M. Golant, E. M. Dianov, O. I. Medvedkov, O. A. Plaksin, V. A.
Stepanov, P. A. Stepanov, P. V. Demenkov, V. M. Chernov, and S. N. Klyamkin, “Ra-
diation Induced Absorption and Luminescence in Specially Hardened Large-Core Silica
Optical Fibers,” in IEEE Transactions on Nuclear Science 47, 693 (2000).
54. J. Nishii, N. Fukumi, H. Yamanaka, K. Kawamura, H. Hosono, and H. Kawazoe, “Photo-
chemical reactions in GeO2 -SiO2 glasses induced by ultraviolet irradiation: Comparison
between Hg lamp and excimer laser,” Phys. Rev. B 52, 1661 (1995).
55. H. Hosono, Y. Abe, D. Kinser, R. Weeks, and K. Kawazoe, “Nature and origin of the
5-eV band in SiO2 :GeO2 glasses,” Phys. Rev. B 46, 1445 (1992).
56. T. Uchino, “Structure and Properties of Amorphous Silica and its Related Materilas:
Recent Developments and Future Directions,” J. of the Ceram. Soc. of Japan 113, 17
(2005).
Bibliography 117

57. D. L. Griscom and E. J. Friebele, “Fundamental radiation-induced defects centers in


syntetic fused silicas: Atomic chlorine, delocalized E0 centers, and a triplet state,” Phys.
Rev. B 34, 7524 (1986).
58. D. L. Griscom, “The Nature of point defects in amorphous silicon dioxyde,” in Defects in
SiO2 and Related Dielectrics: Science and Technology, G. Pacchioni, L. Skuja, and D. L.
Griscom, eds., p. 117 (2000).
59. D. L. Griscom, “Optical properties and structure of defects in silica glass,” Journ. Ceram.
Soc. of Japan 99, 923 (1991).
60. L. Skuja, H. Hosono, and M. Hirano, “Laser-induced color centers in silica,” in Proc. of
SPIE 4347, 155 (2001).
61. L. Hobbs and X. Yuan, “Topology and topological disorder in silica,” in Defects in SiO2
and Related Dielectrics: Science and Technology, G. Pacchioni, L. Skuja, and D. L.
Griscom, eds., p. 37 (2000).
62. A. C. Wright, “Defect-free vitreous network: The idealised structure of SiO2 and related
glasses,” in Defects in SiO2 and Related Dielectrics: Science and Technology, G. Pac-
chioni, L. Skuja, and D. L. Griscom, eds., p. 1 (2000).
63. G. Pacchioni, “Ab initio theory of point defects in SiO2 ,” in Defects in SiO2 and Related
Dielectrics: Science and Technology, G. Pacchioni, L. Skuja, and D. L. Griscom, eds., p.
161 (2000).
64. W. H. Zachariasen, “The atomic arrangement in glass,” J. Am. Chem. Soc. 54, 3841
(1932).
65. S. R. Elliott, Physics of Amorphous Materials (Longman, London, 1990).
66. H. Bach and N. Neuroth, The properties of optical glasses (Springer-Verlag, 1995).
67. R. A. B. Devine, J. P. Duraud, and E. Dooryhée, Structure and Imperfections in Amor-
phous and Crystalline SiO2 (Wiley, Chichester, New York, 2000).
68. R. A. Weeks, “Paramagnetic resonance of lattice defects in irradiated quartz,” J. Appl.
Phys. 27, 1376 (1956).
69. D. L. Griscom, “Characterization of three E0 -center variants in X- and γ-irradiated high
purity a-SiO2 ,” Nucl. Inst. and Methods B 1, 481 (1984).
70. R. A. Weeks, “The many varieties of E0 centers a review,” J. Non-Cryst. Solids 179, 1
(1994).
71. G. Buscarino, S. Agnello, and F. M. Gelardi, “Characterization of E0δ and triplet point
defects in oxygen-deficient amorphous silicon dioxide,” Phys. Rev. B 73, 045 208 (2006).
72. G. Buscarino, S. Agnello, F. M. Gelardi, and A. Parlato, “Electron paramagnetic res-
onance investigation on the hyperfine structure of the E0δ center in amorphous silicon
dioxide,” J. Non-Cryst. Solids 353, 518 (2007).
118 Bibliography

73. G. Buscarino, R. Boscaino, S. Agnello, and F. M. Gelardi, “Optical absorption and


electron paramagnetic resonance of the E0α center in amorphous silicon dioxide,” Phys.
Rev. B 77, 155 214 (2008).

74. V. A. Radzig, “Defects on activated silica surface,” in Defects in SiO2 and Related Di-
electrics: Science and Technology, G. Pacchioni, L. Skuja, and D. L. Griscom, eds., p.
339 (2000).

75. D. L. Griscom, “E0 center in glassy SiO2 : 17


O, 1 H and "very weak" 29
Si superhyperfine
structure,” Phys. Rev. B 22, 4192 (1980).

76. M. G. Jani, R. B. Bossoli, and L. E. Halliburton, “Further characterization of E01 center


in crystalline SiO2 ,” Phys. Rev. B 27, 2285 (1983).

77. H. Nishikawa, E. Watanabe, D. Ito, and Y. Ohki, “Kinetics of enhanced photogeneration


of E’ centers in oxygen deficient silica,” J .Non-Cryst.Solids 179, 179 (1994).

78. R. Boscaino, M. Cannas, F. M. Gelardi, and M. Leone, “ESR and PL centers induced by
gamma rays in silica,” Nucl. Instr. Meth Phys. Res. B116, 373 (1996).

79. L. Skuja, J. Non-Cryst. Solids 239, 16 (1998).

80. H. Hanafusa, Y. Hibino, and F. Yamamoto, “Formation mechanism of drawing-induced


E0 centers in silica optical fibers,” J. Appl. Phys. 58, 1356 (1985).

81. Y. Hibino and H. Hanafusa, “ESR Study on E0 -Centers Induced by Optical Fiber Drawing
Process,” Japan. J. Appl. Phys. 22, L766 (1983).

82. E. P. O’Reilly and J. Robertson, “Theory of defects in vitreous silicon dioxide,” Phys.
Rev. B 27, 3780 (1983).

83. H.Imai, K.Arai, H.Imagawa, H.Hosono, and Y.Abe, “Two types of oxygen-deficient cen-
ters in synthetic silica glass,” Phys.Rev. B 38, 12 772 (1988).

84. L.N.Skuja, A.N.Streletsky, and A.B.Pakovich, “A new intrinsic defects in amorphous


SiO2 : twofold coordinated silicon,” Solid State Comm. 50, 1069 (1984).

85. L. Skuja, “The origin of the intrinsic 1.9 eV luminescence band in glassy SiO2 ,” J. Non
Cryst. Solids 179, 51 (1994).

86. H. Hanafusa, Y. Hibino, and F. Yamamoto, “Formation of drawing-induced defects in


optical fibers,” J. Non Cryst. Solids 95, 655 (1987).

87. Y. Hibino and H. Hanafusa, “Defect structure and formation mechanism of drawing-
induced absorption at 630nm in silica optical fibers,” J. of Appl. Phys. 60, 1797 (1986).

88. M. Cannas, L. Vaccaro, and B. Boizot, “Spectroscopic parameters related to non-bridging


oxygen hole centers in amorphous SiO2 ,” J. Non-Cryst. Solids 352, 203 (2006).
Bibliography 119

89. G. D. L. and M. Mizuguchi, “Determination of the visible range optical absorption spec-
trum of peroxy radicals in gamma-irradiated fused silica,” J. Non-Cryst. Solids 239, 66
(1998).

90. D. L. Griscom, “Radiation hardening of pure-silica-core optical fibers: Reduction of in-


duced absorption bands associated with self-trapped holes,” Appl. Phys. Lett. 71, 175
(1997).

91. D. L. Griscom, “Self-trapped holes in amorphous silicon dioxyde,” Phys. Rev. B 40, 4224
(1989).

92. L. Skuja, “Isoelectronic series of twofold coordinated Si, Ge, and Sn atorns in glassy SiO2:
a luminescence study,” J. Non-Cryst. Solids p. 77 (1992).

93. C. M. Carbonaro, V. Fiorentini, and F. Bernardini, “Stability of Ge-related point defects


and complexes in Ge-doped SiO2 ,” Phys. Rev. B 66, 233 201 (2002).

94. A. Sakoh, M. Takahashi, T. Yoko, J. Nishii, H. Nishiyama, and I. Miyamoto, “Photo-


chemical process of divalent germanium responsible for photorefractive index change in
GeO2 -SiO2 glasses,” Opt. Express 11, 2679 (2003).

95. M. Takahashi, A. Sakoh, K. Ichii, Y. Tokuda, T. Yoko, and J. Nishii, “Photosensitive


GeO2 -SiO2 films for ultraviolet laser writing a channel waveguides and Bragg gratings
with Cr-loaded waveguides structure,” Appl. Opt. 42, 4594 (2003).

96. M. Fujimaki, T. Watanabe, T. Katoh, T. Kasahara, N. Miyazaki, Y. Ohki, and


H. Nishikawa, “Structures and generation mechanisms of paramagnetic centers and ab-
sorption bands responsible for Ge-doped SiO2 optical-fiber gratings,” Phys. Rev. B 57,
3920 (1998).

97. J. Nishii, H. Yamanaka, H. Hosono, and H. Kawazoe, “Characteristics of 5-eV absorption


band in sputter deposited Ge02 -SiOa2 thin glass films,” Appl. Phys. Lett. 64, 282 (1994).

98. M. Cannas, Point Defects in Amorphous SiO2 : Optical Activity in the Visible, UV and
Vacuum-UV Spectral Regions, Phd thesis, Dipartimento di Scienze Fisiche ed Astro-
nomiche, Università di Palermo, Italy (1998).

99. S. Kannan, J. L. M. E. Fineman, and G. H. Sigel-Jr., Appl. Phys. Lett. 63, 3440 (1993).

100. S. Girard, J.-P. Meunier, Y. Ouerdane, A. Boukenter, B. Vincent, and A. Boudrioua,


Appl. Phys. Lett. 84, 4215 (2004).

101. E. J. Friebele, D. L. Griscom, and G. H. Siegel, “Defect centers in a germanium-doped


silica-core optical fiber,” J. Appl. Phys. 8, 3424 (1974).

102. T. E. Tsai, D. L. Griscom, E. J. Friebele, and J. W. Fleming, “Radiation induced defect


centers in high-purity GeO2 glass,” J. Appl. Phys. 62, 2264 (1987).
120 Bibliography

103. N. Chiodini, F. Meinardi, F. Morazzoni, A. Paleari, and R. Scotti, “Optical transitions of


paramagnetic Ge sites created by x-ray irradiation of oxygen-defect-free Ge-doped SiO2
by the sol-gel method,” Phys. Rev. B 60, 2429 (1999).
104. G. Pacchioni and C. Mazzeo, “Paramagnetic centers in Ge-doped silica: A first-principles
study,” Phys. Rev. B. 62, 5452 (2000).
105. J. Nishi, K. Kintaka, H. Hosono, H. Kawazoe, M. Kato, and K. Muta, “Pair generation of
Ge electron centers and self-trapped hole centers in GeO2 -SiO2 glasses by KrF excimer-
laser irradiation,” Phys. Rev. B 60, 7166 (1999).
106. M. Fujimaki, T. Kasahara, S. Shimoto, N. Miyazaki, S. Tokuhiro, K. S. Seol, and
Y. Ohki, “Structural changes induced by KrF excimer laser photons in H2 -loaded Ge-
doped SiO2 glass,” Phys. Rev. B 60, 4682 (1999).
107. E. Friebele and D. Griscom, “Defects in Glasses,” in Mater Res. Soc. Symp., M. W.
F.L. Galeneer, D.L. Griscom, ed., 61, 319 (1986).
108. C. P. Slichter, Principles of Magnetic Resonance (Springer-Verlag, Hong Kong, 1991).
109. J. W. Chan, T. Huser, J. S. Hayden, S. H. Risbud, and D. M. Krol, “Fluorescence Spec-
troscopy of Color Centers Generated in Phosphate Glasses after Exposure to Femtosecond
Laser Pulses,” J. Am. Ceram. Soc. 85, 1037 (2002).
110. R. A. Weeks and P. J. Bray, “Electron Spin Resonance Spectra of Gamma-Ray-Irradiated
Phosphate Glasses and Compounds: Oxygen Vacancies,” J. Chem. Phys. 48, 1 (1968).
111. P. Ebeling, D. Ehrt, and M. Friedrich, “X-ray induced effects in phosphate glasses,” Opt.
Mater. 20, 101 (2002).
112. G. Pacchioni, D. Erbetta, D. Ricci, and M. Fanciulli, “Electronic Structure of Defect
Centers P1, P2, and P4 in P-Doped SiO2,” J. Phys. Chem. B 105, 6097 (2001).
113. U. Hoppe, G. Walter, A. Bartz, D. Stachel, and A. Hannon, “The P-O bond lengths in
vitreous P2 O5 probed by neutron diffraction with high real-space resolution,” J. Phys.:
Condens. Matter 10, 261 (1998).
114. R. K. Brow, “Review: the structure of simple phosphate glasses,” J. Non-Cryst. Solids
263-264, 1 (2000).
115. V. G. Plotnichenko, V. O. Sokolov, V. V. Koltashev, and E. M. Dianov, “On the structure
of phosphosilicate glasses,” J. Non-Cryst. Solids 306, 209 (2002).
116. G. Walter, G. Goerigk, and C. Rüssel, “The structure of phosphate glass evidenced by
small angle X-ray scattering,” J. Non-Cryst. Solids 352, 4051 (2006).
117. S. Girard, J. Baggio, J.-L. Leray, J.-P. Meunier, and A. B. Y. Ouerdane, “Vulnerability
analysis of optical fibers for Laser Megajoule facility: Preliminary studies,” in IEEE
Transactions on Nuclear Science 52, 1497 (2005).
Bibliography 121

118. E. Friebele, C. Askins, C. Shaw, M. E. Gingerich, C. C. Harrington, D. L. Griscom,


T.-E. Tsai, U.-C. Paek, and W.H.Schmidt, “Correlation of single-mode fiber radiation
response and fabrication parameters,” Appl. Opt. 30, 1944 (1991).
119. S. Girard, Y. Ouerdane, A. Boukenter, and J.-P. Meunier, “Transient radiation responses
of silica-based optical fibers: influence of Modified Chemical Vapor Deposition process
parameters,” J. Appl. Phys. 99, 023 104 (2006).
120. B. Tortech, S. Girard, E. Régnier, Y. Ouerdane, A. Boukenter, J.-P. Meunier, M. V.
Uffelen, A. Gusarov, F. Berghmans, and H. Thienpont, “Core versus cladding effects of
proton irradiation on erbium-doped optical fiber: micro-luminescence study,” in IEEE
Transactions on Nuclear Science 55, 2223 (2008).
121. iXFiber Website: http://www.ixfiber.com.
122. G. Origlio, S. Girard, M. Cannas, Y. Ouerdane, R. Boscaino, and A. Boukenter, “Para-
magnetic germanium-related centers induced by energetic radiation in optical fibers and
preforms,” J. Non-Cryst. Solids 355, 1054 (2009).
123. S. Girard, N. Richard, Y. Ouerdane, G. Origlio, A. Boukenter, L. Martin-Samos, P. Pail-
let, J.-P. Meunier, J. Baggio, M. Cannas, and R. Boscaino, “Radiation Effects on Silica-
Based Preforms and Optical Fibers - II: Coupling Ab initio Simulations and Experi-
ments,” in IEEE Transactions on Nuclear Science 55, 3508 (2008).
124. K. Médjahdi, A. Boukenter, Y. Ouerdane, F. Messina, and M. Cannas, “Ultraviolet-
induced paramagnetic centers and absorption changes in singlemode Ge-doped optical
fibers,” Opt. Express 14, 5585 (2006).
125. P. Paillet, J. R. Schwank, M. R. Shaneyfelt, V. Ferlet-Cavrois, R. L. Jones, O. Flament,
and E. W. Blackmore, “Comparison of charge yield in MOS devices for different radiation
sources,” in IEEE Transactions on Nuclear Science 49, 2656 (2002).
126. S. Girard, Y. Ouerdane, G. Origlio, C. Marcandella, A. Boukenter, N. Richard, J. Baggio,
P. Paillet, M. Cannas, J. Bisutti, J.-P. Meunier, and R. Boscaino, “Radiation Effects
on Silica-Based Preforms and Optical Fibers - I: Experimental Study with Canonical
Samples,” in IEEE Transactions on Nuclear Science 55, 3473 (2008).
127. D. Harris and M. Bertolucci, Symmetry and Spectroscopy: An Introduction to Vibrational
and Electronic Spectroscopy (Oxford University Press, 1978).
128. A. Meijerink, Luminescence of Solids (Plenum Press, New York, 1998).
129. J. R. Lakowicz, Principles of Fluorescence Spectroscopy (Springer, New York, 2006).
130. Opotek Website: www.opotek.com.
131. Desy Website: www-hasylab.desy.de.
132. J. R. Ferraro, K. Namamoto, and C. W. Brown, Introductory Raman Spectroscopy (Aca-
demic Press, UK, 2003).
122 Bibliography

133. A. Thorne, U. Litzen, and S. Johansson, Spectrophysics: principles and applications


(Springer-Verlag, Germany, 1999).

134. S. Girard, D. Griscom, J. Baggio, B. Brichard, and F. Berghmans, “Transient optical


absorption in pulsed-X-ray-irradiated pure-silica-core optical fibers: Influence of self-
trapped holes,” J. Non-Cryst. Solids 352, 2637 (2006).

135. G. Origlio, A. Boukenter, S. Girard, N. Richard, M. Cannas, R. Boscaino, and Y. Ouer-


dane, “Irradiation induced defects in fluorine doped silica,” Nucl. Instr. and Meth. In
Phys. Res. B 266, 2918 (2008).

136. A. Abragam and B. Bleaney, Electron paramagnetic resonance of transition ions (Claren-
don, Oxford, 1970).

137. C. P. P. Jr., Electron spin resonance (Wiley, New York, 1967).

138. G. E. Pake and T. L. Estle, The physical principles of electron paramagnetic resonance
(W. A. Benjamin Inc., Massachussets, 1973).

139. S. Agnello, Gamma ray induced processes of point defect conversion in silica, Phd thesis,
Dipartimento di Scienze Fisiche ed Astronomiche, Università di Palermo (Italy) (2000).

140. S. Agnello, R. Boscaino, M. Cannas, and F. M. Gelardi, “Instantaneous diffusion effect


on spin-echo decay: Experimental investigation by spectral selective excitation,” Phys.
Rev. B 64, 174 423 (2001).

141. S. Agnello, R. Boscaino, M. Cannas, A. Cannizzo, F. M. Gelardi, S. Grandi, and


M. Leone, “Temperature and excitation energy dependence of decay processes of lu-
minescence in Ge-doped silica,” Phys. Rev. B 68, 165 201 (2003).

142. J. H. Lee and S. K. Oh, “Residual Stress Change of Optical Fiber for Ge Dopant Varia-
tion,” in SPIE 5279, 483 (2004).

143. P. Y. Fonjallaz, H. G. Limberger, R. P. Salathé, F. Cochet, and B. Leuenberger, “Ten-


sion increase correlated to refractive-index change in fibers containing UV-written Bragg
gratings,” Opt. Lett. 20, 1346 (1995).

144. N. Belhadj, Y. Park, S. LaRochelle, K. Dossou, , and J. Azaña, “UV-induced modification


of stress distribution in optical fibers and its contribution to Bragg grating birefringence,”
Opt. Express 16, 8727 (2008).

145. P. L. Chu and T. Whitbread, “Measurement of stresses in optical fiber and preform,”
Appl. Opt. 21, 4241 (1982).

146. G. Origlio, S. Girard, R. Boscaino, A. Boukenter, M. Cannas, and Y. Ouerdane, “Optical


and photonic material hardness for energetic environments,” in 9ème Colloque sur les
Sources Cohérentes et Incohérentes UV, VUV et X (UVX) (2008).
Bibliography 123

147. S. Agnello, R. Boscaino, M. Cannas, F. M. Gelardi, F. L. Mattina, S. Grandi, and


A. Magistri, “Ge related centers induced by gamma irradiation in sol-gel Ge-doped silica,”
J. Non Cryst. Solids 322, 134 (2003).
148. H. Hanafusa, Y. Hibino, and F. Yamamoto, “rawing condition dependence of radiation-
induced loss in optical fibres,” Elect. Lett. 22, 106 (1986).
149. A. Q. Tool, “Relation between inelastic and thermal expansion of glass in its annealing
range,” J. Am. Ceram. Soc. 29, 240 (1946).
150. C. M. Smith and L. A. Moore, “Properties and production of F-doped silica glass,” J.
Fluorine Chem. 122, 81 (2003).
151. P. Dumas, J. Corset, W. Carvalho, Y. Levy, and Y. Neuman, “Fluorine doped vitreous
silica analysis of fiber optic preforms by vibrationnal spectroscopy,” J. Non-Cryst. Sol.
p. 239 (1982).
152. F. L. Galeener and G. Lucovsky, “Longitudinal optical vibrations in glasses: GeO2 and
SiO2 ,” Phys. Rev. Lett. 1474, 1474 (37).
153. F. L. Galeener, “Band limits and the vibrational spectra of tetrahedral glasses,” Phys.
Rev. B 19, 4292 (1979).
154. F. L. Galeener, The Structure of Non-Crystalline Materials (Taylor & Francis, London,
1982).
155. F. L. Galeener, “Planar rings in glasses,” Solid State Commun. 44, 1037 (1982).
156. R. A. B. Devine, R. Dupree, I. Farnan, and J. J. Capponi, “Pressure-induced bond-angle
variations in amorphous SiO2 ,” Phys. Rev. B 35, 2560 (1987).
157. L. Vaccaro, M. Cannas, B. Boizot, and A. Parlato, “Radiation induced generation of
non-bridging oxygen hole center in silica: Intrinsic and extrinsic processes,” J. Non-Cryst.
Solids 353, 586 (2007).
158. G. Buscarino, Experimental investigation on the microscopic structure of intrinsic para-
magnetic point defects in amorphous silicon dioxide, Phd thesis, Dipartimento di Scienze
Fisiche ed Astronomiche, Università di Palermo (Italy) (2006).
159. R. Schnadt and A. Räuber, “Motional effects in the trapped-hole center in smoky quartz,”
Solids State Commun. 9, 159 (1971).
160. J. H. E. Griffiths and J. Owen, “Paramagnetic Resonance in Neutron-Irradiated Diamond
and Smoky Quartz,” Nature 173, 439 (1954).
161. L. E. Halliburton, N. Koumvakalis, M. E. Markes, and J. J. Martin, “Radiation effects
in crystalline SiO2 : The role of aluminum,” J. Appl. Phys. 52, 3565 (1980).
162. P. S. Rao, R. J. M. Eachern, and J. A. Weil, “On a proposed radiation-induced polaronic
hole in silicon dioxide,” J. Comput. Chem. 12, 254 (1991).
124 Bibliography

163. P. Karlitschek, G. Hillrichs, and K. Klein, “Photodegradation and nonlinear effects in


optical fibers induced by pulsed uv-laser radiation,” Optics Commun. 116, 219 (1995).
164. K. Okamoto, K. Toh, S. Nagata, B. Tsuchiya, T. Suzuki, N. Shamoto, and T. Shikama,
“Temperature dependence of radiation induced optical transmission loss in fused silica
core optical fibers,” J. Nucl. Mat. 329-333, 1503 (2004).
165. S.-I. Kitazawaa, S. Yamamoto, M. Asano, and S. Ishiyama, “Radiation−induced lumi-
nescence and photoluminescence from sol−gel silica glasses and phosphosilicate glasses
by 1 MeV H+ irradiations,” Physica B 349, 159 (2004).
166. M. D’Amico, F. Messina, M. Cannas, M. Leone, and R. Boscaino, “Homogeneous and
inhomogeneous contributions to the luminescence linewidth of point defects in amorphous
solids: Quantitative assessment based on time-resolved emission spectroscopy,” Phys.
Rev. B 78, 014 203 (2008).
E al fin della licenza, io tocco . . .

À la fin de l’envoi, je touche . . .

You might also like