mechanics and experiments

171
Distributed Bragg Grating Sensing-Strain Transfer Mechanics and Experiments Graham Ian Duck December 14, 2000 Thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy, Graduate Department of Aerospace Studies in the University of Toronto Copyright O Graham Ian Duck, 2001

Upload: others

Post on 07-Feb-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mechanics and Experiments

Distributed Bragg Grating Sensing-Strain Transfer

Mechanics and Experiments

Graham Ian Duck

December 14, 2000

Thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy, Graduate Department of Aerospace Studies in the University of Toronto

Copyright O Graham Ian Duck, 2001

Page 2: Mechanics and Experiments

National Library 1+1 ,,,a BibIiotheque nationale du Canada

Acquisitions and Acquisitions et Bibliographic Services services bibliographiques

395 Wellington Street 395, rue Wellington Ottawa ON K I A O N 4 Ottawa ON K1 A O N 4 Canada Canada

The author has granted a non- exclusive licence allowing the National LLbrary of Canada to reproduce, loan, distribute or sell copies of this thesis in microform, paper or electronic formats.

L'auteur a accorde une licence non exclusive pennettant a la Bibliotheque nationale du Canada de reproduire, prgter, distribuer ou vendre des copies de cette these sous la fonne de microfiche/^ de reprocluction sur papier ou sur format electronique .

The author retains ownership of the L'auteur conserve la propriete du copyright in this thesis. Neither the droit d'auteur qui protege cette these. thesis nor substantial extracts from it Ni la these ni des extraits substantiels may be printed or otherwise de celle-ci ne doivent Etre imprixnes reproduced without the author's ou autrement reproduits sans son permission. autorisation.

Page 3: Mechanics and Experiments

Abstract Distributed Bragg Grating Sensing-Strain Transfer Mechanics and Experi-

ments, Ph. D Thesis by Graham Ian Duck.

The University of Toronto Institute for Aerospace Studies, 200 1.

The optical fiber sensor is a tool by which a measurement of a thermo-mechanical field

inside a mechanical host can be made. It is possible to interrogate the optical fiber sensor

with light- The parameters of the in-fiber sensor determine the response of the sensor to

Lightwave interrogation. Having the sensor response, a reconstruction of the in-fiber structure

can be attempted. Successful reconstruction of the in-fiber structure allows an inference

of the behaviour of the mechanical substrate. The substrate behaviour is ultimately the

information of interest. Extraction of such a measurement may be seen as two sub-problems.

The first problem addresses the ability of the mechanical system to transfer surrounding

material information into the optical fiber sensor. It is not the field values inside the optical

fiber sensor that are particularly interesting, instead, it is the degree to which the in-fiber

field values accurately represent the field values of the surrounding material. It is important

to examine the mechanical system's ability to transfer strain from surrounding material to

optical fiber sensor. Certainly, if strain information is not transferred into the optical fiber,

even an ideal interrogation of the in-fiber sensor cannot report the full strain field of the

mechanical host.

The second issue concerns the sensor itself. The reflectivity characteristics of the loaded

sensor must first be recovered experimentally, then interpreted in meaningful way. The

experimental arrangement to recover the sensor response and interpretation of this data

constitute the optical measurement problem.

This thesis examines the two problems.

Page 4: Mechanics and Experiments

Acknowledgments

First, I would like to thank my supervisor Dr. Raymond Measures for including me in his

dynamic group and allowing me the freedom to grow both aicademicdy and professionally

in the field of fiber optic sensing, and for providing me a stepping stone into the important

field of telecommunications.

My senior colleagues Sharon Huang, Michel Leblanc and Myo Ohn offered advice and

time without limit. Without their input and support, this work would not Have been possible.

I would also like to extend sincere thanks to colleagues Tino Alavie, Rob Maaskant, Shai

Berger, Jason Bigue, Murray Lowery, Ning-Yao Fan, Trent Coroy: and Paul kldvihill.

I extend deepest gratitude to my new colleagues, Henry Postolek, David Moss, Martin

Matthews, Dan Grobnic, Jeffrey KalbBeisch, Dominique Brichard, Michael Lamont and

David Markin for their expertise and support.

I would like to thank professors Dr. Rod Tennyson, Dr. Jorn Hansen and Dr. David

Zingg for their friendly guidance and technical tips.

Wholehearted thanks go out to my dear fiiends Rob C. and Carol-lynn D. for their

consistent and humour-filled support - Thanks Eddie V. and Kurt C. for the many hours of musical accompaniment during the

preparation of this document.

Mom and dad, I am not sufficiently articulate to express how lucky I am that I have

parents like you.

Finally, I acknowledge the generous financial support of fSIS Canada for making investi-

gation of this technology possible.

Page 5: Mechanics and Experiments

Contents

1 Introduction 1

I . 1 Problem definition and document organization . . . . . . . . . . . . . . . . . 3

. . . . . . . . . . . . . . . . . . . . . . . . . . 12 A note on co-ordinate systems 6

2 Optical description of the fiber Bragg grating 8

2.1 TheBragggrating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Coupled mode equation 10

2.3 T-matrix formulation for grating simulation . . . . . . . . . . . . . . . . . . 17

2.4 Inverse Fourier transform method for low reflectivity gratings . . . . . . . . . 23

2.5 Hypothetical infinitely long and infinitely short grating discussion . . . . . . 31

3 Optical Response of fiber Bragg grating to strain and temperature 33

3.1 f hase sensitivity of fiber optic sensors . . . . . . . . . . . . . . . . . . . . . . 34

3.2 Experimental evaluation of sensitivities . . . . . . . . . . . . . . . . . . . . . 44

4 Interpretation of mechanical interactions: emphasis on strain transfer be-

tween host and fiber sensor 53

4.1 Strain transfer considerations for point sensors . . . . . . . . . . . . . . . . . 54

4.1.1 Shear lag formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.2 Axial strain transfer fiom host to fiber: Finite element models . . . . . . . . 61

4.2.1 Surface bonded sensor model . . . . . . . . . . . . . . . . . . . . . . . 61

4-22 Embedded sensor model . . . . . . . . . . . . . . . . . . . . . . . . . 72

4.2.3 Finite element analysis of linear strain gradient . . . . . . . . . . . . 76

Page 6: Mechanics and Experiments

4.3 Transfer of arbitrary strain distribution . . . . . . . . . . . . . . . . . . . . . 92

4.4 Equivalent embedded geometric ratio for s d w e bonded sensor . . . . . . . 98

4.5 Conclusions of mechanical strain transfer analysis . . . . . . . . . . . . . . . 103

5 Demonstration of method and measurements 105

. . . 5.1 Review of methods for interrogating gratings with non-uniform profiles 106

5.1.1 Phase-based spectrum (PSB) strategy . . . . . . . . . . . . . . . . . . 106

5.1.2 Intensity based spectrum (ISB) strategy . . . . . . . . . . . . . . . . 108

5.1.3 Low coherence interferometry . . . . . . . . . . . . . . . . . . . . . . 110

5.2 Implementation of interferometric phase recovery experiment . . . . . . . . . 112

5.2.1 DFB tunable source . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5.2.2 Representative grating interrogation . . . . . . . . . . . . . . . . . . . 119

. . . . . . . 5.3 Interferometry-based phase recovery distributed sensing strategy 125

5.3.1 Uneven thermal apparent loading: experiment a1 demonstration of spa-

tial resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

5.3.2 Grating bonded in the vicinity of a circular hole . . . . . . . . . . . . 130

5.4 Delay-based phase recovery distributed sensing strategy . . . . . . . . . . . . 134

5.4.1 Measurement principle . . . . . . . . . . . . . . . . . . . . . . . . . 134

5.5 Conclusions of measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6 Conclusions and contribution summary 146

6.1 Summary of contributions and conclusions . . . . . . . . . . . . . . . . . . . 147

6.1.1 Contributions in understanding the optical response . . . . . . . . . . 147

. . . 6.1.2 Contributions towards understanding the strain transfer problem 147

. . . . . 6.1.3 Contributions in the area of distributed strain measurements 148

6.2 Suggested future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

Page 7: Mechanics and Experiments

List of Figures

. . . . . . . . . . . . . . . . . . . . . 1.1 Conceptual flow of optical fiber sensing 5

1.2 Cartesianandcyllndricalcoordinatesusedinthisdocument . . . . . . . . . . 7

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Fiber Bragg grating 9

2.2 Uniform grating reflectivity and phase response (bn = 2 x lo-', L = 5cm.

no = 1.5. h, = 500nm) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3 Uniform grating reflectivity and delay response (bn = 2x10-'. L = 5cm.

no = 1.5. A, = 500nm) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4 Cascaded uniform gratings model a complex grating . . . . . . . . . . . . . . 24

2.5 Incremental increase of reflectivity linearly related to local coupling strength 25

2.6 Range of reflectivities as predicted by the coupled mode solution and the FT

approximation-d gratings modeled with L = 5cm . . . . . . . . . . . . . . . . 28

2.7 Ratio of maximum reflectivity as predicted by the coupled mode solution and

the FT approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

. . . 2.8 Quality of coupling distribution recovery using Born (FT) approximation 30

3.1 Butter and Hocker Experiment . . . . . . . . . . . . . . . . . . . . . . . . . 35

3.2 Change in effective index and periodicity of a grating subject to strain field . 41

3.3 Pull test experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3.4 Pinch test experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3.5 Pinch test load apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.6 Apparatus to measure fiber stiffness . . . . . . . . . . . . . . . . . . . . . . . 50

4.1 Fiber segment in infinite host under uni-axial tension . . . . . . . . . . . . . . 55

Page 8: Mechanics and Experiments

. . . . . . . . . . . . . . . . . . . . . . 4.2 Sum of axial acting forces in the fiber 55

4.3 Swn of axial acting forces in the surrounding material . . . . . . . . . . . . . 56

4.4 Shear lag plots for N = 0.1. N = 0.2. a n d N = 0.3. . . . . . . . . . . . . . . 59

. . . . . . . . . . . . . . . . . . . . 4.5 Common optical fiber sensor installations 60

. . . . . . . . . . . . . . . . . . . . . . . . . 4.6 Model for surface mounted fiber 62

. . . . . . . . . . . . . . . 4.7 Node distribution across Reissner/Mindlin element 66

4.8 Quasi-t hree-dimensional adhesive element node posit ions . . . . . . . . . . . . 68

. . . . . . . . . . . . . . . . . . . . 4.9 Quasi-three-dimensional adhesive element 69

. . . . . . . . . . . . . . . . . . . . . . . 4.10 Axi-symmetric finite element model 73

. . . . . . . . . . . . . . . . . . . . . . . 4.11 Imposed strains and displacements 77

4.12 Equivalent transferred linear strain gradient . . . . . . . . . . . . . . . . . . 78

4.13 Location of lamina modeled as having no stifiess . . . . . . . . . . . . . . . 81

4.14 Shear stress distribution at fiber/surrounding material interface for embedded

. . . . . . . . . . . . . . . . . . . . L - O 82 installation . " = 1.05, 5 = 70, ;; - f Em

4.15 Quality of transfer for surface mounted installation (Vertical axis represents

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . & x 1000) 86

4.16 Gradient elongation. 5. for stiffness ratio of 2 = 14 (Surface mounted) . . 87

4.17 Gradient elongation. 5. for stiffness ratio of = 35 (Surface mounted) . . 88

4.18 Quality of transfer for embedded installation (Vertical axis represents 3 x 1000) 89

4.19 Gradient elongation. u; for stiffness ratio of 2 = 14 (Embedded) . . . . . 90 rf

4.20 Gradient elongation. e. for stiffness ratio of 2 = 35 (Embedded) . . . . . 91

4.21 Illustration of problem: axial strain transfer from host to sensor . . . . . . . 92

4.22 Transfer function. H ( k ) dependence on spatial frequency k . . . . . . . . . . 95

4.23 Transfer of far field sinusoids . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

4.24 Transfer o f far field Gaussians . . . . . . . . . . . . . . . . . . . . . . . . . . 99

4.25 Peak suppression and broadening of Gaussian far field strain profile . . . . . 100

4.26 Surface bonded installation modelled as embebbed installation with equivalent

radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.1 Phase accumulation due to reflection at z of length dz . . . . . . . . . . . . . 107

5.2 Closed loop schematic of[l] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

Page 9: Mechanics and Experiments

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Open loop schematic of[l]

. . . . . . . . . . . 5.4 DBR Source interferometric experiment with lock-in phase

. . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Phase sbifted grating as DFB

. . . . . . . . . . . . . . . . . . . . . . . . 5.6 DFB laser wavelength calibration

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7 Interferometric experiment

5 -8 Cosine of inferred phase against interferometric reflectivity . . . . . . . . . .

5.9 Reflectivity and coupling distribution experiment and simulation comparison

5.10 Grating of which one half is bonded to an alllminum substrate . . . . . . . . 5.11 Measured reflectivities from sensor partially bonded to an aluminum substrate

subject to thermal load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.12 Apparent strains reported from sensor bonded on alumhum substrate subject

tothermalload . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.13 Grating bonded to an aluminum plate in the vicinity of a circular hole . . .

5.14 Experimental and simulated reflectivities for different loadings . . . . . . . . .

5.15 Measured and analytical strain profiles for grating bonded near circular hole

for different axial loads (upper plot) . Deviation between theory and experi-

ment (lower plot) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.16 Repeated measurements of two different mountings of the same load (left) and

deviations horn elasticity solution (right) . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . 5.17 Phase shift modulation method

5.18 Delay-based phase recovery experiment . . . . . . . . . . . . . . . . . . . . .

5.19 Beam loaded in bending . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.20 Reflectivity and delay of unloaded and loaded samples . . . . . . . . . . . . .

5.21 Strain profiles recovered from delay based interrogation . . . . . . . . . . . .

5.22 Comparison of two gratings lcm and 5cm of like peak reflectivity subject to

Like localized strain perturbation . . . . . . . . . . . . . . . . . . . . . . . . .

Page 10: Mechanics and Experiments

Nomenclature

A: area (FEM context)

A,: cross sectional area of optical fiber coating

Af: cross sectional area of optical fiber

aij: individual elements of stress, moment, and transverse resultant matrix

a,: m-th Fourier co-efficient of 672 series

A,: Amplitude envelope of m-th optical mode

a (2): backward-propagating envelope with electric field phase dependence

A (2): spatially varying envelope of backward moving wave

a: angle of polarization interrogation (in context of Jone's calculus)

a f : fiber thermal expansion co-efficient

cr j: thermal eG-ion co-efficient tensor

g: magnetic flux density vector

[B]: shorthand for [L] [a] matrix product (FEM context)

bij: individual elements of stress, moment, and transverse resultant matrix

b (2): forward propagating envelope with electric field phase dependence

B (2) : spatially varying envelope of forward moving wave

P,: propagation constant of m-th optical mode

C: stress-optic co-efficient

[Dl : stiffness matrix

5: electric flux density vector

df: fiber diameter

Djj: stiffness tensor

dij : individual elements of stress, moment, and transverse result ant matrix

(6): generalized displacement vector (FEM context)

{6Ie: discrete generalized element nodal displacement vector (FEM context)

6,B: detuning between optical wave and Bragg resonance

&: Kronecker delta function

Page 11: Mechanics and Experiments

672: index modulation

t?: electric field vector em, ,. . - t o td strain tensor

eh-i: mechanical strain tensor

{e } : generalized strain vector (FEM context)

Ec: elastic modulus of optical fiber coating

Ef: elastic modulus of optical fiber

Em: elastic modulus of surrounding material

~ f : electric permittivity of optical fiber

€9: electric pennit tivity due to grating corrugation

~ o : vacuum electric permittivity

Ei j : electric permittivity tensor

€ f i j : electric permittivity tensor of optical fiber

c g i j : electic permittivity tensor due to grating corrugation

q: transform plane co-ordinate (FEM context)

{Fig: global nodal force vector (FEM context)

f : superscript indicating an in-fiber quantity

{F)': element nodal force vector (FEM context)

f-: RF modulation frequency

[a]: interpolation function matrix (FEM context)

(@] ': element interpolation function matrix (FEM context)

& i-th interpolation function (FEM context)

g: accelearation due to gravity

G,: surrounding material shear modulus

y: phase offset (interferometric and polarimetric experiment contexts)

g: magnetic field vector

H (k): transfer function describing strain transfer of spatial frequency k

i: drive current

I: optical intesity

lij : identity tensor

Page 12: Mechanics and Experiments

[J]: Jacobian matrix (FEM context)

k: spatial frequency (multiple contexts)

K: complex coupling strength

[K] e: element st f i e s s matrix (FEM context )

[KIg: global stiffness matrix (FEM context)

K : coupling strength

L: length (used in multiple contexts)

[L] : operator matrix (FEM context)

A: optical wavelength

A,: nominal optical wavelength

A,: eigenvalue of shear lag equation-physically interpreted as spatial wavelength

A,: spatial period of m-th Fourier component used in reconstruction of coupling distri-

bution

A: period of fiber Bragg grating index modulation

rn: mass

m: superscript indicating a surrounding material quantity

[MI: moment resultant matrix (FEM context)

n: index of refraction

n: effective index of refraction

N: shear lag parameter

[N]: stress resultant matrix (FEM context)

qij: impermeability tensor

u: Poisson's ratio

urn: surrounding material Poisson's ratio

vf: fiber Poisson's ratio

p: subscript indicating transverse polarization

P: mechanical line load in force per meter

p: polarization density vector

[PI : lower virtual node specification matrix (FEM context)

pijkl: full Pockei's strain-optic tensor

Page 13: Mechanics and Experiments

pl l : reduced index Pockel's co-efficient

p12: reduced index Pockel's co-efficient

$: phase (used in several contexts)

Q,: m-th modal solution of wave equation

[Q]: transverse shear resultant matrix (FEM context)

r : radial position (cylindrical co-ordinates)

r': position vector

r (bp) : electric field proportional reflectivity

R (w) : power proportional reflectivity

r f : fiber radius -

[R]: upper virtual node specification matrix (FEM context)

s: eigenvalue of coupled mode equation

S.: phase sensitivity to axial strain (gauge factor)

ST: phase sensitivity to temperature

ui: stress tensor

in-fiber axial normal stress

in-fiber normal stress averaged over cross section

in-fiber axial radial stress

in-fiber radial stress averaged over cross section

in-fiber azimuthal normal stress

in-fiber azimuthal normal stress average over cross section

surrounding material axial normal stress

surrounding material normal stress averaged over cross section

(0): generalized stress vector (FEM context)

t: time

T: temperature

zj: T-matrix entries - xj: complex conjugate of T-matrix entries

8: angle of polarizer (in context of Jone's calculus)

8: azimuthal position (cylindrical co-ordinates)

Page 14: Mechanics and Experiments

8: m-th transverse mode shape of optical wave

Oi: i-th rotation-like mid-plane variable of Reissner/Mindlin formulation

{ T ) : generalized surface traction vector (FEM context)

r: time delay

rh: in-fiber shear stress

TZ: surro~~lding material shear stress

q-: i-th cartesian displacement

ii$ i-th midplane cartesian displacement

u: radial displacement (cylindrical co-ordinates)

po: vacuum magnetic permeability

v: azimuthal displacement (cylindrical cc>-ordinates)

V: fringe visibility

V: volume (FEM context)

w : axial displacement (cylindrical cuordinat es )

W (A): wavelength dependent window function

w : optical frequency (in radians/second)

x, : m-t h cartesian dimension

Cij: normalized sensitivity of refkactive index to temperature

X i j : electric susceptibility tensor

<: transform plane co-ordinate (FEM context)

z: axial position (cylindrical co-ordinates)

Page 15: Mechanics and Experiments

Chapter 1

Introduction

Page 16: Mechanics and Experiments

Fiber optic structural sensing is now widely accepted as a high resolution, versatile strat-

egy for extracting therm+mechanical information kom mechanical structures[2] [3]. Butter

and Hocker[4] first derived and demonstrated a phase strain relationship between the load

state of a Michelson interferometer and an optical phase response seen as a n interference

pattern on a screen. Since this initial demonstration the field has evolved considerably. This

work intends to contr ib~te to the evolution of fiber optic sensing, in the particular area of

distributed Bragg grating sensing.

The availability of components for the field of optical fiber sensing is largely due to

the telecommunications community. The confluence of technological advances in optimizing

the attenuation and dispersion properties of optical fibers and the advent of a laser source

to provide coherent light at the appropriate wavelength range supported, and continues to

suppoa, a worldwide research interest in teletommunications [5]. The materials and devices

employed in thesis for purposes of structural sensing: optical fiber, the in-fiber Bragg grating,

and the various light sources owe their development to telecommunications research.

Fiber sensors deployed as structural sensors offer several advantages over conventional

resistive foil strain gauges[6]. The small diameter of the sensor lends itself well to structural

integration-in particulax in the field of composite materials. Within fiber optic sensing,

the in-fiber Bragg grating has emerged as a leading alternative for structural sensing[7] [8].

The dominant advantage of the fiber Bragg sensor is that its response to thermo-mechanical

perturbations is seen in a n absolute wavelength change; this contrasts with interferometric

sensors whereby relative int erferometric phase excursions do not distinguish between, for

example, tensile or compressive strain perturbations of the same magnitude. The wavelength

encoded shift has been demodulated[9] by a filter having a near-linear wavelength response.

This technique provides far successful point measurements.

A second, very significant, advantage of the Bragg grating sensor is demonstrated later:

the full spectral response (including phase) of the Bragg grating encodes the full distributed

strain field inside the sensar[lO]. Where the point sensing strategy reports a measurement in-

dicating the average thermo-mechanical perturbation over the sensor's length, the distributed

approach attempts to recover the t herm~mechanical perturbation as spatially varying func-

tion over the spatial extent of the sensor. The potential of this implementation of distributed

Page 17: Mechanics and Experiments

Bragg grating sensing has been investigated in Msious modalities through the 1990's a t the

Fiber Optic Smart Structures laboratory at the University of Toronto Institute for Aerospace

Studies and at other institutions.

This work continues the exploration of the distributed Bragg grating sensing and seeks

to make a contribution t o this field.

1.1 Problem definition and document organization

The optical fiber sensor is a tool by which a measurement of a thermo-mechanical field

inside a mechanical- substrate or host can be made. It is possible to interrogate the optical

fiber sensor with Light. The parameters of the in-fiber sensor determine the response of the

sensor to lightwave interrogation [ll] [12]. Having the sensor response, a reconstruction of

the in-fiber structure can be attempted. Successful reconstruction of the in-fiber structure

allows an inference of t he behaviour of the mechanical substrate. The substrate behaviour is

ultimately the information of interest. The progression of these relationships constitutes the

fundamental problem addressed by this thesis. In this document: this problem is organized

as two broad subproblems:

0 the mechanical strain transfer problem and?

the optical measurement problem.

The &st problem addresses the ability of the mechanical system to transfer surrounding

material information into the optical fiber sensor. It is not the field values inside the optical

fiber sensor that are particularly interesting, instead, it is the degree to which the in-fiber

field values accurately represent the field values of the surrounding material. It is important

to examine the mechanical system's ability to transfer strain from surrounding material to

optical fiber sensor. Certainly, if strain information is not transferred into the optical fiber,

even an ideal interrogation of the in-fiber sensor cannot report the full strain field of the

mechanical host. This question is addressed in literature for the case of a single ended point

sensor by LeBlanc 1131. The reference brings attention to the fact that strain mismatch is

expected between sensor and substrate.

Page 18: Mechanics and Experiments

The second issue concerns the sensor itself. The reflectivity characteristics of the loaded

sensor must first be recovered experimentally, then interpreted in meaningful way The

experimental arrangement to recover the sensor response and interpretation of this data

constitute the optical measurement problem. The background of this problem is provided in

large part by the candidate's predecessors at the University of Toronto. Huang et. al. [lo] describe the dependence of spectral characteristic on the internal strain state of the fiber

B r a g grating. The origins of the particular modality of distributed Bragg studied in this

document are found in two initially distinct approaches, both pursued a t the University of

Toronto. Leblanc [14] proposes an intensity based spectrum analysis, while, concurrently,

Huang [15] identifies the potential of a phase based spectrum approach to distributed sensing.

Further, Ohn [16] consolidates the intensity and phase based approaches. The work of O h

is the direct precursor to the work presented in this thesis. As well, Volanthen [I?] [18] [I], at

the University of Southampton, develops an alternative experimental approach for recovering

in-grating parameters with optical interrogation.

With a view to examining these problems, the following document structure, sketched

schematically in fiewe 1.1, is adopted.

Chapter 2 relates the background necessary to understand the Bragg grating sensor's

response in terms of its in-fiber structure. Starting with Maxwell's equations, the governing

coupled mode equations are derived in detail. An exact reflectivity prediction for a uniform

Bragg grating is discussed. Arrows of fiewe 1.1 describe either 'forward' or 'inverse' scatter-

ing directions. The prediction of spectral response given knowledge of the physical structure

of the grating is an example of the forward scattering problem. The forward scattering

solution is outlined for the case of the arbitrary (norr-uniform) grating with the numeri-

cal T-Matrix formulation. The inverse scattering prob!em refers to recovery of the in-fiber

parameters given optical properties, reflectivity and phase of the grating. The inverse strat-

egy employed in this work is derived and the consequences of simplifving assumptions are -

discussed.

The expected response of in-fiber parameters to thenno-mechanical perturbations is the

subject of chapter 3. The phase strain relationships for the distributed optical fiber sensor

are detailed. Additionally, contributions are made in the experimental evaluation of the

Page 19: Mechanics and Experiments

Optical Data Representing Thermo-mechanical Field

Fiber Bragg grating optical d A response

Chapter 3: e of in-fiber parameters

to in-fiber thermo-mechanical

Chapter 4: ship between

host material strain and in fiber strain

Thermo-mechanical Field in Su bstrate/Host Material

Figure 1.1: Conceptual Row of optical fiber sensing

Page 20: Mechanics and Experiments

sensitivity parameters required for the measurements that follow in the document. Experi-

mental reports of the strain-optic coefficients, fiber stiffness, thermal and strain sensitivities

are reported.

Chapter 4 investigates the mechanical load transfer relationship between the in-fiber

field and the host substrate field. This content is particularly relevant to the new field of

distributed sensing where fields are expected to vary in space along the sensor length. In

collaboration with co-authour G d a u r n e Renaud contributions are made in the area of pre-

dicting the in-fiber strain field given the substrate/host field. As many optical fibers are

surface mounted to structures, a finite element modeling approach is described to estimate

the strain transfer behaviour kom host/substrate to surface mounted optical fiber. Further,

with colleague and c-authour Michel LeBlanc, a novel theoretical relationship between in-

fiber field and the substrate/host field is presented for the case of an axi-symmetric embedded

fiber sensor. A strategy to represent the surface mounted fiber as a geometrically equiva-

lent embedded fiber is proposed so that the aforementioned theoretical predictions can be

extended to the surface mounted case.

Chapter 5 puts into application the content of chapters 2 through 4. Experimental results

of a demonstrated experimental strategy[lb] are reported in further detail, with particular

emphasis on specifying the resolution of the sensor against theoretical elasticity solutions.

Additional data treatment techniques are discussed. With colleague and co-authour &My0

0 hn, an alternative experimental strategy is demonstrated and characterized.

Conclusions of this work are made in the final chapter. Included is a summary of contri-

butions.

1.2 A note on co-ordinate systems

In this document, Cartesian, (XI, 12, x3), and cylindrical co-ordinates (2, r, 6) are used where

appropriate. In the Cartesian case, xl lies parallel to the optical axis. In the cylindrical

system, z lies parallel the optical axis. Co-ordinate systems are shown in figure 1.2.

Page 21: Mechanics and Experiments

Figure 1.2: Cartesian and cylindrical coordinates used in this document.

Page 22: Mechanics and Experiments

Chapter 2

Optical description of the fiber Bragg

grating -

Page 23: Mechanics and Experiments

Figure 2.1: Fiber Bragg grating

2.1 The Bragg grating

The in-fiber Bragg grating is an all fiber device that has the property of reflecting a narrow

band spectrum related to its in-fiber parameters. Figure 2.1 illustrates a fiber B r a g grating

interacting with forward and backward propagating waves. In the context of structural

sensing the grating is typically interrogated in reflection. The all fiber device employs a

spatially periodic index variation to couple incident light from mode to mode. The period of

the index variation, A, is typically sub-micron. The index variation, hn, is small compared

to the nominal effective index, the order of

To induce a grating in an optical fiber, one takes advantage of the phenomenom of

photosensitivity, first observed by Hill[19]. Photosensitivity refers to a fiber's potential to

have its index of refraction modified by light exposure. The writing technique of Hill is

Page 24: Mechanics and Experiments

not widely used due to its self-written nature. Self written gratings arise from standing

wave pattern inside the fiber core: generated by reflection of in-fiber propagating Light-

This method accommodates little control of either the spatial period of the grating, as it

was determined by the launch light, and offered little spatial control of the distribution

of the index modulation strength-apodization-an issue critical to tailoring the reflection

spectrum-

The breakthrough contribution of Meltz et. aZ.[20] resolves some of these difEculties by

proposing a new transverse writing method. The authors indicate how gratings can be ex-

ternally impressed into an optical fiber by introducing the fiber into a standing wave pattern

generated by two hee space interfering beams. Called the transverse holographic method,

this st rate,^ permits wavelength control, and some apodization control. A further improve-

ment is reported by Anderson et. aZ.[21]. This group describes a difiactive optical element

(now commonly called a phase mask) through which a transmitted beam emerges to produce

a standing intensity field whose parameters are determined by the physical characteristics of

only the mask and the writing beam. The gratings used in this work are written with the

phase mask technique.

2.2 Coupled mode equation

The fundamental description of the optical response of a fiber Bragg grating is provided

by coupled mode theory- Coupled mode theory allows the prediction of an optical wave's

behaviour in the presence of a periodic or near periodic variation in the index of refraction

of the waveguide.

Before embarking on the mathematical derivation of the coupled mode relationship, a

qualitative description is useful. The physical key to :understanding the Bragg reflection

is the interaction between the lightwave and its medium. The optical wave oscillates in

space with a period defined by its wavelength and t&e effective index of the waveguide.

This wave interacts with the guiding material by way cf the medium's polarization density.

For sufficiently non-dispersive media, the polarization density is driven to oscillate in phase

by the optical wave. Should the medium be dispersive, the magnitude of the polarization

Page 25: Mechanics and Experiments

density's oscillation response will depend on the frequency of the optical wave. This in turn

induces a frequency dependent phase velocity for that particular optical fkeque~cy. If the

polarization densie itself is spatially periodic, the optical wave has opportunity to resonate

with the medium. An optical mode is able to transfer energy to the medium, while the

medium, through accelerating charges implicit in the resonating of the polarization density,

is able to couple energy back into optical modes. Given a precise knowledge of the make-

up of the periodic structure, coupled mode theory allows the prediction of power transfer,

through the medium, between optical modes.

The natural starting point is Maxwell's equations,

The fields in equations 2.1 through 2.4 are the electric field, @', t ) ; the magnetic field,

<(F, t ) ; the electric flux density, 5(F, t ) ; and the magnetic flux density, @F, t).

Two constitutive relationships describe the role of the medium in interacting with an

electromagnetic field,

where p(F, t) is the polarization density of the medium. The free space electric permittivity

appears as E, and p0 denotes the magnetic permeability of the vacuum.

The wave equation, is derived as equation 2.6 is substituted into equation 2.1, then

d x 3? (F, t ) is replaced in the resuiting expression with the relationship 2.2. a - 5 (7, t) is

expanded as in equation 2.5 and the curl operation, v x , is performed on the result. The

vector identity x 9 x Z(F, t ) = d (d - - f2g(F, t ) is employed to finally achieve

Page 26: Mechanics and Experiments

A more detailed development of the wave equation many can be found in most optics text-

books, Saleh and Teich[22] for example.

The medium polarization, induced by the electric field is most generallyt anisotropic,

non-linear, inhomogeneous and dispersive,

in which X i j represents the electric susceptibility of the medium. For purposes of arriving

at the coupled mode equations, it is sufficient to model the grating as a linear, spatially

inhomogeneous, isotropic structure, where equation 2.8 reduces to

In terms of the medium permittivity, equation 2.9 is recast as

where E (3 indicates the full spatially v w g permittivity of the medium. Equation 2.10

can be rewritten to explicitly reflect the permittivity contributions from the unperturbed

fiber, q (3 : and from the periodicity of the Bragg grating, E , (3,

In terms of the effective index of the waveguide and the index modulation of the grating,

expanding the square as a two term binomial, equation 2.11 appears as

The A quantity indicates the effective index of the waveguide. For a guided wave, a spatial

distribution of index of refraction exists for purposes of binding light propagation in the

waveguide. After solving the appropriate waveguide equation, one may find the propagation

behaviour of the bound modes. From the speed of the modes, the effective index, E is found.

Typically. lies between the maximum index of the core and the minimum index of the

cladding.

Page 27: Mechanics and Experiments

In this document, the fiber will be considered weakly guiding[23]]. This assumption allows

the reduction of equation 2.7 from a full vector expression to a scalar equation, for which

an approximation to the transverse component of the electric fieltd will be solved. Thus the

wave equation to be solved is

Note that the left hand side of equation 2.13 is the governing relationship born which the

permissible mode-field solutions in the unperturbed waveguide cam be found. The right hand

side of the equation accounts for the periodic structure in the wa~eguide.

Following Yariv's normal mode analysis[24], the electric field i s expanded in terms of the

m possible waveguide modes at frequency w ,

In equation 2.14, O;" (r, 0) represents the transverse shape functzion of the m-th mode of

propagation, with forward propagation constant, Dm. The transvezrse shape functions can be

found by solving the wave equation for an unperturbed waveguide The expression

represents a solution to the equation

Further details of such a solution: and the precise form of O;L (r, 0) can be found in Okoshi[25].

As well, the modes of the unperturbed waveguide solution are orthogonal and normalized in

the sense that

To account for coupling between modes, the envelope of the electric field is included as

A, (2). The latter term describes, as a function of z; the growth or decay of a mode along

the waveguide as energy is transfered from mode to mode.

The mode field expansion of equation 2.14 is substituted into t h e wave equation 2.13. Af-

ter some manipulation, including eliminating the time dependence, the following expression

Page 28: Mechanics and Experiments

is recovered

1 = -2n (3 (I) w2 C ZAm ( r ) 8;L (r, 0) e-i8.1z.

m

Because Q, (c t) is a solution to equation 2.16, BF (r, 8) homogeneously solves the Helmholtz

equation, thus the content of the first bracketed group of terms is zero. Equation 2.18 reduces

At this point, equation 2.19 is multiplied by 8; (r, 8) and integration is performed over the

cross section: resulting in

xe-"' /Lo lIo o: (T, 8 ) OF (r, 0) rdrdi3

1 = - E / I ~ E ~ S ~ (I) w2 ;Am ( z )

In view of equation 2.17, on the left hand side of equation 2.20 only two possible values of m

will result in non-trivial values after integration, m = +s. The reduced equation will appear

The indices f s indicate modes of the same order, with identical transverse shape functions,

but with propagation constants of opposite sign. The opposite sign of the propagation

constants indicate forward and backward propagating waves.

Page 29: Mechanics and Experiments

Further reduction of equation 2.21 merits some comment. In the original work of Yariv[24]

two assumptions are made. The first is that the envelope of the electric field is slowly varying.

This assumption is argued to provide the basis for the following simplifying assumption

This assumption is used to justify the elimination of the second order derivative terms relative

to the &st order derivative terms in equation 2.21. The second assumption is that only terms

on the right hand side of equation 2.21 that are in spatial resonance may drive any terms

appearing on the left hand side of equation 2.21. The latter assumption provides for the

separation of equation 2.21 into the two coupled mode equations relating the forward to

backward propagating waves. Sipe[26] has shown that both these assumptions are incorrect

for the general periodic structure. Using a total field analysis, Sipe explicitly indicates that

contradicting the second assumption of Yariv. Kashyap [27] argues that the assumptions of

Yariv are admissible in the case of a weakly coupled interaction and preserves the steps of

the original derivation.

It occurs that the result of Yariv is not at all an approximation, but exact, due to a coin-

cidental cancelation errors. To arrive at the final coupled mode equations, the assumptions

of normal mode analysis will be used.

The index of refraction perturbation will be expanded as a periodic spatial function,

written as

q=-a0

where A indicates the period of the cormgation. The index modulation is a real quantity,

thus the sum of equation 2.24 is re-expressed as w

671 ( z ) = 2 a, cos (q?) . q=O

Equation 2.24 is inserted into equation 2 -2 1 and the weakly coupled assumption is invoked

yielding

Page 30: Mechanics and Experiments

/d /,a Qp (r, 0 ) 8; (r, 0 ) l-drde-

In equation 2.26, there can be interaction between terms on the right side and terms

on the left side only if there is spatial phase matching between the terms. Consider the

M e - i f i 8 z a2 expression. the only terms capable of resonating with this term are those with

a phase matched spatial period. SpeciGcally, those terms that satisfy

In equation 2.27, q- indicates the order of the Fourier component of the corrugation that

couples the modes. Equation 2.27 is referred to as the phase matched condition or the Bragg

condition. Thus for m = s in the right hand side sum of equation 2.26, it is found

d A-, zEe,w -- - -As ( 2 ) &%-%) dz 2 oq JW T-=o 1% e=o [OF (r, 6)12 rdrde.

The coupling constant, K is introduced to simphfy the notation. K represents

Using equations 2-17 and 2.25, K is rewritten as

In equation 2.30 An represents the Fourier co-efficient of the resonant spatial period of the

index variation due to the grating. Thus, the coupling constant is seen to be a measure of the

periodic structure's ability to couple power between optical modes. The detuning, 60, which

is a measure of the mismatch between the period of the grating and twice the propagation

constant of the optical wave, is now introduced

A periodic, but non-sinusoidal corrugation would yield resonances at q values other than

unity. The second order resonance occurs at a resonant wavelength half that of the first

order. In practical interrogations; this second order resonant wavelength will be outside the

Page 31: Mechanics and Experiments

range of sources or detectors-approximately 775727~ for typical gratings used in this thesis.

It is appropriate to study only the first resonance where q = 1.

The notation of equations 2-30 and 2.31 are invoked and the first coupled mode relation-

ship is found

With rn = -s in the right hand s u m of equation 2.26 the second coupled mode equation is

analogously derived. Finally, the notation is clarified by replacing the &s subscripts with

the following notation: A (I) represents the backwards propagating wave and B ( z ) denotes

the forward propagating wave. This notation will be used in the entirety of this document.

The h a l coupled mode equations are thus presented as

dl4 (4 = ~ K B (2) exp [-2i6flzl and, dz

where the E denotes the complex conjugate.

2.3 T-matrix formulation for grating simulation

Equations 2.33 can be solved analytically. The analytical solution of the uniform grating

will be seen to be an important building block in the prediction of the optical response of

arbitrarily complex intra-grating structure. Equations 2.33 can be reduced to two uncoupled

second order differential equations,

d2 A (z) dA (2.1 + 2i6fl- - KKA ( z ) = 0 dz2 dz

d2B (2) d B (2) dz2

- 2ibP- - KEB ( 2 ) = 0 , dz

admitting the general solutions

Page 32: Mechanics and Experiments

where s = is introduced. At this point, no restrictions will be imposed on the

boundary conditions of the grating and the derivation will proceed with the most general

boundary conditions: A (0) = Ao, B (0) = B,, A (L) = AL: and B (L) = BL. These

boundary conditions can be substituted into equations 2.35 and the envelope functions are

re-expressed

and r = L,

in terms of the magnitudes of the input and output envelopes at both z = 0

To arrive efficiently at a solution the derivative of A (2) f?om 2.36 is computed and substituted

into the first of equations 2.33 and the expression is evaluated at z = L resulting in

s cosh (sL) - ibp sinh (sL) ei6p A0 = AL [

s I + B ~ [

4 6 sinh (s L) -i6pL e

S I -

The analogous operation is performed for B, to recover the following equality,

iri sinh (sL) i6pL

S I + BL[

s cosh (sL) + is0 sinh (sL) S I

It is convenient to add the spatial dependence on the propagation constant to the envelope;

this includes the full wave dependence of the forward and backward propagating waves.

Specifically, a ( z ) = A ( r ) e-iBL and b (2) = B (z j eiBL. Equations 2 -37 and 2.38 express a

useful relationship that may be better seen by casting the equations in matrix form. They

Page 33: Mechanics and Experiments

where

s cosh (sL) - 260 sinh (s L) z1 = s

exp [-iAL]

- & and denote complex conjugates of zl and z2. It is seen that

the waveforms at z = 0 to the waveforms at z = L. Equation 2.39 is

equation 2.39 relates

referred to as the T-

Ma tnx relationship and provides the basis for the description of complex grating structures.

The T-matrix formalism is based on the work of Yamada and Sakuda[28], originally used to

describe the reflectivity of distributed feedback laser and applied and verified in an intra-

grating sensing context by Huang[lO] .

Typic-, the proportiofid reflectivity is sought from the grating. This is simply the

ratio

In the simple case of a spatially uniform grating, subject to single end interrogation, i.e. the

backward traveling is set as A (L) = 0, equation 2.41 reduces to

Equation 2.42 describes the ratio of electric fields incident and reflected from the grating.

The power proportional reflectivity is the square of the electric field proportional reflectivity,

in which T denotes the complex conjugate of r . The grating's full power reflectivity is h d y

written as ri2 sinh2 (sL)

R ( 6 4 = s2 cosh2 (sL) + 6P sinh2 (sL) '

The uniform grating response is plotted in figures 2.2 and 2.3 In the plotted example,

the following parameters were used: 6n = 2 x lo-'; L = 5cm, A. = 500n.m and = 1.5. It

is seen that the maximum reflectivity as a function of detuning occurs at the phase matched

condition, 6P = 0

R (60 = 0) = tanh2 (sL) . (2.45)

Page 34: Mechanics and Experiments

Photonic Bandgap

.-.-. -.-I_.__ ----. ---.

Wavelength Offset M [nml

Figure 2.2: Uniform grating reflectivity and phase response (6n = 2 x lo-', L = 5cm, rz, =

1.5, A, = 500nm)

Page 35: Mechanics and Experiments

Photonic Bandgap

Wavelength Offset AA [nml

Figure 2.3: Uniform grating reflectivity and delay- response (672 = ~ X I O - ~ , L = 5cm, no =

1.5, A, = 500nm)

Page 36: Mechanics and Experiments

In terms of wavelength, the center of the reflection spectrum given by 6D = 0 is

The time delay characteristic indicates the differential delays between different probe wave-

lengths. At larger detunings the value of delay approaches 250ps. Consider a simple time

of fight argument. A reflection from the physical mid-point of the grating requires 250ps

to make the return trip from the front of the grating. Effectively, detuned wavelengths are

reflected from the grating center. The lowest value of delay is seen at the phase matched

condition. The wavelength that best resonates with the structure is effectively reflected from

a point physically closer to the front of the grating. For this example, the lowest value of de-

lay is approximately 120ps. This indicates that the penetration depth for the phase matched

wavelength is shortened to approximately a quarter of the grating length. A stronger grat-

ing will display a further decreased penetration depth for the phase matched wavelength.

A third observation can be made regarding the periodicity of the oscillations in the delay

characteristic a t det uned wavelengths. The maxima of the delay characteristic coincide with

the minima in amplitude response. Consider a Fabry-Perot cavity having the same length

as the grating, 0.05m. The free spectral range of such a cavity- is 15pm. This spacing is seen

between successive maxima in delay on both the lower and higher wavelength sides of the

delay spectrum. At detuned wavelengths, the grating shows a Fabry-Perot cavity behaviour

where the maxima in delay indicate wavelengths that are trapped in the cavity.

Recalling s = J-=, it is seen that as the detuning becomes larger than the mag-

nitude of the coupling coefficient: s becomes a complex-quantity. The region in which the s

parameter is real is referred to as the photonic bandgap. This region includes a significant

part of the reflectivity spectrum and is used as an indication of the grating's bandwidth.

In figures 2.2 and 2.3, vertical bars indicate the width of the photonic bandgap. Setting

the coupling constant equal to the detuning, (expressing detuning in terms of wavelength % E M excursion, 60 = T), the half width of the photonic bandgap is

In the present example, the half-bandwidth is 0.ln.m. The first zeros occur in outside the pho-

tonic bandgap where the solution becomes oscillatory. Setting R (6P) = 0 in equation 2.44,

Page 37: Mechanics and Experiments

the zeroes are found as sL = *irn.ir where m is any positive integer. The locations of the

zeroes are approximated by

In the current example, they occur at k0.019nm fkom the nominal wavelength.

The availability of an analytical solution to the coupled mode equations for a uniform

grating allows a numerical solution to more complex grating structures. Consider figure 2.4.

Equation 2 -39 describes the relationship between the forward and backward propagating

waves at either side of the grating. A complex grating is modeled as a series of cascaded

uniform gratings, each having an arbitrary pitch, index modulation, and effective index. The

relationship between sequential gratings is natural. The (n - 1)-th subgrating segment is

related to the n-th segment by

where [TIn denotes the n-th uniform T-matrix of equation 2.39. The same relationship

holds between the n-th segment and the (n + 1)-th. Thus, multiplied over the entirety of

subgrating segments, the relationship between the left side of the first segment and the right

hand side of the last segment appears as

2.4 Inverse Fourier transform method for low reflec-

tivity gratings

There are several theoretical ways to attempt to invert the coupled mode equations. The

following derivation was proposed by Kogelnik(291 for thin-film optical waveguides in 1975. It

remains valid for periodic and near periodic structures in cylindrical waveguides. To arrive

at an inverse scattering approximation, first, the electric field proportional reflectivity of

Page 38: Mechanics and Experiments

Figure 2.4: Cascaded uniform gratings model a complex grating

equation 2.41 is written as a function of 2:

then the derivative with z is taken,

Expressions 2 -33 are substituted into equation 2.52 to yield

dr (2) = [i~e*''~~ + nr2 (t) e-2i6Bz] d,z (2.53)

Expression 2.53 indicates the rate of change of reflectivity over an infinitesimal length of

uniform grating, dz. Consider a concatenation of infinitesimal elements, where a variation

in the magnitude of K and A are possible along the grating. Equation 2.53 is approximated

by

dr (2) = i ~ e ~ ~ ~ ~ d z . (2.54)

In equation 2.54 the term proportional to r2 ( z ) has been neglected. This low reflectivity

assumption limits the accuracy in exchange for computational efficiency. The simplification

Page 39: Mechanics and Experiments

Figure 2 -5: Incremental increase of reflectivity linearly related to local coupling strength

is also equivalent to the application of the Born approximation. The Born approximation

is a conventional 'first order' approach to solving the non-linear wave equation. Specifically,

when an optical field is incident on a non-linear medium, the incident optical field excites the

non-linear medium and the medium responds by resonating, thus emitting a second order

optical field. The second order optical field, in turn, will also interact with the medium and

induce a further higher order field. The Born approximation neglects all but the &st order

field originating from the interaction from the incident field on the medium. Approximate

equation 2 -54 expresses the Born simplification: the incremental induced optical field, dr (2)

is seen to be dependent on the local medium alone-independent of the presence of the field

or the medium at other points.

Equation 2.54 indicates the Born approximated inversion of the coupled mode equation.

Specifically. as drawn in figure 2.5, the local increase of reflectivity is related only to the

local value of the coupling strength. The general case of the non-uniform grating will have

a spatial distribution of coupling strength. Thus, the coupling strength is expressed as a

spatially varying function,

K + K (z) . (2.55)

As well! the phase term needs to be refined to reflect the variation with axial position. It is

re-expressed as [30]

Page 40: Mechanics and Experiments

where the detuning is now written as a distributed function: reflecting any axial variation

of the propagation constant and pitch- The axially varying detuning is written as the sum

of a spatially invariant nominal detlrning and a term representing the deviation from the

nominal state,

bp (2) = m a + (W (2)) . (2.57)

qb is introduced to clarify the notation

Thus, the refmed phase term appears as

26,~ -+ 26P,2 + $J (z) .

K ( z ) is introduced to simplify the writing of equation 2.54. K (2) appears as

K ( z ) = i~ (2) e-iJI(z). (2.60)

Thus: substituting equations 2.55 through equation 2.60 into equation 2.54, and integrat-

ing the result from r = -00 to z = oo and considering only a grating where no backwards

traveling optical power is incident on the RHS of the grating, an expression for the reflectivity

at the LHS side of the grating is found as

Equation 2.61 is a Fourier relationship, and can be inverted to yield[29]

Expressing detuning explicitly in terms of wavelength, it is seen

Oa K (2) = --

A: L, r (A, 0) exp [laid - d ~ . A:

Equations 2 -61 and 2.62 approximate the relationship between the in-fiber structure

and the complex reflectivity. The relationship improves with lower reflectivities. The low

Page 41: Mechanics and Experiments

reflectivity approximation for a uniform

with I

gratimg can be obtained by solving equation 2.61

KO Eor 0 < z < L

0 elsewhere -

It is found that r (AP) = KO sin (b/3L) /bp. Thrus, the power-proportional reflectivity is

Equation 2.65 has an obvious physically in&d attribute-potentially: it predicts a peak

reflectivity larger than unity, for a [ K , I L prodmct larger than unity.

The reflectivity -expressed in equation 2.65 Es compared with the exact solution of equa-

tion 21.44 for a 5cm uniform- grating. The agreement between the exact solutions and the

approximate solutions: for gratings of differend strengths are plotted in figure 2.6. From

strongest to weakest gratings, with peak refleceivities of 99%, 50%: 25%, and 10% are con-

sidered (corresponding values of the product l K l L are 3.0, 0.88, 0.54, and 0.32 respectively).

For the strongest grating the approximation is : clearly invalid. The deviation between the

exact and approximate solution are extreme. The approximate solution erroneously pre-

dicts a peak reflectivity of 9. As reflectivity decreases the approximation improves. At 10%

reflectivity agreement is strong, with the Fouriier transform approximation overstating the

reflectivity somewhat at its peak: but otherwise capturing the oscillatory portion of the so-

lution outside the photonic bandgap. Figure 2 .7 indicates the decrease in integrity of the

approximation as the product IKIL increases. I:t is seen that with a IK[L as low as 0.4, the

approximation is more than 10% in error,

The strength of the inverse relationship, equation 2.62 is tested as well. Again, only

uniform gratings are considered. A uniform gratEng's reflectivity is found using equation 2.42,

this reflectivity, r (6p) is then used in equation 2.62 and the deviation between the imposed

coupling distributions and the prediction are pPotted in figure 2.8. The same four gratings

are analyzed as in the previous paragraph. The .agreement is very poor in the case of a I KI L

product of 3. Note that the magnitude of K is well represented at the front of the grating but

deviates dramatically towards the far end. T h i s is a consequence of power that is reflected

from the fiont of the grating is better approximated by the Born approximation. Deeper

Page 42: Mechanics and Experiments

Figure 2.6: Range of reflectivities as predicted by thi coupled mode solution and the FT

approximation-all gratings modeled with L = 5cn

Page 43: Mechanics and Experiments

0.0 0.2 0.4 0.6 0.8 1 .O Grating Strength [KL]

Figure 2.7: Ratio of maximum reflectivity as predicted by the coupled mode solution and

the FT approximation.

Page 44: Mechanics and Experiments

Figure 2.8: Quality of coupling distribution recovery using Born (FT) approximation.

Page 45: Mechanics and Experiments

into the structure, for stronger coupling, the interaction between higher perturbations begins

to build. These observations suggest that gratings with peak reflectivities of approximately

10% are best suited to deployment as distributed sensors using the FT inversion strategy

2 -5 Hypothetical infinitely long and infinitely short grat-

ing discussion

Two extreme examples are given to further an intuitive understanding of the relationship

between a grating's structure and its optical response. Consider an infinitely long, infinitely

weak sating: with a coupling distribution described by, K (2) = KO. Equation 2.61 will

appear as

Completing the integral, it is found that r (6,B) is proportional to a delta function, 6 (6P).

Thus, an infinitely long grating would reflect only a single wavelength at the phase-matched

condition.

On the other extreme, consider a measurement by which the experimenter found a re-

flectivity, constant in wavelength, with a constant accumulation of phase as one tuned the

interrogation source,

where z0 is introduced to indicate the rate of phase accumulation. Inserting equation 2.67

it is found that the coupling distribution is proportional to another delta function, this

time in length space, 5 (G - 2). This narrow band reflector is in fact a mirror, reflecting

all wavelengths. The fringe density seen by the experimenter is related to the mismatch

between interferometer arms, this mismatch distance is 2,.

These two admittedly extreme examples of uniform gratings can be useful in providing

some intuitive guidance in the more complex structures discussed later in this document.

Following the above examples, it is noticed that spatial features over a short length tend to

Page 46: Mechanics and Experiments

have wideband repercussions in the reflectivity, and narrowband features are generated from

long spatial features.

Page 47: Mechanics and Experiments

Chapter 3

Optical Response of fiber Bragg

grating to strain and temperature

Page 48: Mechanics and Experiments

3.1

Section

given a

Phase sensitivity of fiber optic sensors

2-2 describes how the optical properties of a periodic structure can be predicted

knowledge of the physical structure itself. This section outlines how environmental

changes in the fiber properties, namely mechanical strains and temperature alter the grating

structure.

Butter and Hocker demonstrated the first optical-fiber strain sensor in 1978[4]. The

authors constructed an optical fiber Mach-Zehnder interferometer, shown in figure 3.1 in

which both arms of the interferometer are bonded to opposite sides of a cantilever beam

loaded in bending. One arm is loaded in tension and the other in compression. The formula

describing the intensity issuing from the Mach-Zehnder given b ~ ' ,

((£1 + £2) (£1 + &)) = E: + E: + 2&£2 cos ((Y + 2A4) . (3-1)

The parameter y represents the constant phase offset resulting from m y initial mismatch

between the optical path lengths in the two arms, while 44, the term of interest, is the strain

induced relative phase shift between the optical path lengths of the two interferometer arms.

In the confi,wation of Butter and Hocker, the two arms are subject to strains, thus phase

changes: of equal magnitude but opposite sign. The mounting on either side of the beam

explains the sign inversion as well as the factor of 2 in equation 3.1.

The phase change of the optical fiber can be expanded as the following

In equation 3.2, P denotes the propagation constant, and df the fiber core diameter. Equa-

tion 3-2 describes the phase change of a fiber subject to uni-axial mechanical stress as the

sum of three effects: the physical elongation of the path length, the index change induced by

the mechanical strain field, and a change in the propagation constant due to contraction of

the fiber core through the Poisson effect. The latter mode dispersion effect is several orders

of magnitude less important than the former two terms and is dropped from the expansion.

Dividing equation 3.2 with the nominal phase, 4, = PL, a normalized phase change is found

as

Page 49: Mechanics and Experiments

Be- Lenses He-Ne Source

L

Load

Screen

Figure 3.1: Butter and Hocker Experiment

As will be seen, this sum of an induced index change and a physical length change is also

appropriate to describe the response of Bragg gratings to mechanical and thermal loads.

The complete phase/strain relationship for the Bragg grating is derived in the following

paragraphs .

First: a simplification of section 2.2 is reconsidered. Specifically, equation 2.9 limits

the susceptibility to an isotropic constant, X, where more generally it is a tensor quantity

dependent on the polarization of the optical field. Preserving the assumption that the sus-

cep tibility is linear and non-dispersive, and adopting tensor notation: equation 2.9 describing

the polarization density induced by the field is recast as

Expressed in terms of permittivities, equation 3.4 becomes

where lij denotes the identity tensor. -4nalogous to the simplification performed in section 2.2

the permittivity tensor is rewritten as a sum of two contributions

Page 50: Mechanics and Experiments

The first contribution is the permittivity tensor of the fiber. This quantity will reflect any

birefringence, either nominal or induced, in the optical fiber. A birefringence free fiber (lo-

bi fiber) would be modeled by dii with like permittivities seen by any polarization. The

general case remains a full tensor. The second term includes the spatial variation of index

of the Bragg grating. As well, it is appropriate to leave the tensor subscript in the second

term. This provides for the possibility of strength birefringence of the Bragg grating in which

one polarization is preferentially reflected. Equation 3-6 is normalized with the free space

Finallyz the tensor wave equation is written as

For any tensor quantity a coordinate choice can be made such that E,, becomes diagonal.

This set of axes is called the principal axes of the medium. In the present context, the

principal axes of the fiber will be chosen as the principal axes of the problem. Once cast in

a principal coordinate system, the permittivity tensor of the fiber can be written in terms

of effective index,

It is noted that there is no assurance that the principal axes of a grating strength coincide

with the principal axes of the fiber, thus while the tensor & is diagonalized, the tensor $ (zl) may continue to have non-zero off diagonal entries. The correct expression of the grating's

permittivity tensor would require a tensor transformation to convert the principal grating

axes to the axes of the fiber. In this document, the grating is assumed to have coincident

principal axis with the fiber. Analogously to section 2.2 the index modulation is expressed

Page 51: Mechanics and Experiments

Thus: again invoking the weakly guided assumption for both polarizations it is found that

equation 3.8 reduces to two polarization uncoupled scalar equations of the following form

where p = 2 ,3 indicates the transverse polarization index.

Solving equation 3.11 for either polarization, a Bragg condition will be found as

When such a grating is subject to mechanical or thermal loading, uniform over the e x t e ~ t of

the grating, the center wavelength will undergo a shift. Expanding equation 3.12 in a Taylor

series about the nominal center wavelength, the normalized wavelength shift is found to be

Ax, AE.. AA -=- - +h- (3.13) x, n p

The relationship in equation 3.13 will be seen to be linear with axial strain. It is this

effect that drew attention to the Bragg grating as an effective strain sensor[3l]. As only

the reflected center wavelength of the sensor is of interest-i-e. the loading information is

wavelength and not intensity encoded-the Bragg grating sensor can be made insensitive to

optical intensity effects such as power fluctuations of the interrogation source and coupling

changes in input and output leads.

Naturally the ratio is equivalent to the total strain component along the fiber axis,

el; as is the ratio of equation 3.2. Thus the phase change derived by Butter and Hocker

is identical to the normalized wavelength shift of the Bragg grating,

For a nominally uniform grating subject to a load field that is spatially varying along the

grating, equation 2.58 is examined. 11 (xl) is the term indicating the accumulated excursion

in detuning away from the nominal detuning 6P,. The p subscript is added to qb to indicate

the different detuning in each polarization axis. The spatial rate of change of accumulated

detuning is the local perturbation of the detuning fiom the unloaded condition, specifically,

Page 52: Mechanics and Experiments

A (6& (XI)) is first expanded to

The propagation constant of polarization p can be expanded as

With equation 3 - 17, equation 3.16 becomes

2xAKp (xl) 2rfiPAA (xl) A ( 6 P p (4) = -I-

A0 A:

Recalling the Bragg condition 3.12, equation 3.18 is recast as

Thus, for an axially varying load across an originally uniform Bragg grating, the recovered

detuning excursion is related to perturbations in the optical fiber sensor as

Again, the RHS of equation 3.20 is a distributed representation of the simple formula p r e

sented by Butter and Hocker for the Mach-Zenhder interferometer,

Equations 3.13, and 3.21 indicate that two phenomena determine the optical response

of either a point or distributed Bragg grating sensor: the physical length change and the

induced index change. In the upcoming paragraphs, the dependence of index of refraction

and of pitch on strain and temperature will be indicated. It is emphasized that the strain

sensitivities of the index and pitch presented in this document, refer to the sensitivities to

total strains, ei , in the sense that total strains indicate the total deformation of the fiber,

regardless of the cause of the deformation. Mechanical strains relate to stresses by Hooke's

Page 53: Mechanics and Experiments

where Di,- is the stiffness tensor. As mechanical strains are induced by mechanical stresses,

the total strains are the sum of the mechanical and thermally apparent contributions

where crf is the tensor of thenno-elastic constants of the optical fiber. It is important to

appreciate this distinction in deriving the phase/strain/temperature sensitivities.

The variation of the pitch length is naturally

recalIing that the total axial strain, e:d-'f, includes the possibility of a combination of me-

chanical and thermal apparent strains, as indicated in equation 3.23.

To describe the change in effective index, the relationship between the electric field and

the electric flux density

'oi = eijEj = eE (I*, + xij ) = c o v j z2-E (3.25)

is re-expressed as 1

for purposes of constructing the index ellipsoid. In equation 3.26, vij denotes the imper-

meability tensor. The sensitivity of the impermeability tensor to strain is described by the

For the most general case: it appears that 81 individual strain optic coefficients are

required. Symmetry arguments[32] reduce the number of Pockels coefficients needed to

completely describe the isotropic fiber core to only two values: pll and pl2. Along the

principal axes

thus

Page 54: Mechanics and Experiments

The constant ,F denotes the normalized sensitivity of the index to temperature,

Thus the index change as a function of environmental perturbations can be written as the

following sum -4 6

A% = -2 xmefd- + E ~ ~ A T . (3.31) i=1

Equations 3.24 and 3.31 are substituted into either equation 3.13 for the point sensing

phase sensitivity or equation 3.21 for the distributed phase sensitivity. The expansion of the

point sensitivity appears as

In terms of distributed sensing

An illustration of the deformation of the grating's pitch and effective index over the length

of the sensor subject to an axial strain distribution is shown in figure 3.2.

The sensitivity of several common loading conditions are now outlined. In deriving these

relationships, it is useful to adopt a common procedure:

Figure the total strains: ejd.lf, usually functions of both strain and temperature, using

equation 3.23.

a Substitute the total strains into the appropriate phase sensitivity expression, either

equation 3.32 or 3.33.

The simplest (and most common) strain situation is the that of uniaxial tension with no

temperature variation, ol ( x l ) = qo. The strain field is

Page 55: Mechanics and Experiments

Axial strain

I I I 1

0 L Xl

n + Deflected

Figure 3.2: Change in effective index and periodicity of a grating subject to strain field.

Page 56: Mechanics and Experiments

Substituting 3.34 into equation 3.32

In a point sensing modality, the bracketed term of equation 3.35 is referred to as the gauge

factor of the sensor, S,,

In practice the gauge factor is evaluated by a one step strain versus wavelength shift cali-

bration of the sensor instead of a calculation based on the estimated value of the individual

parmeters.

A temperature perturbation on the optical fiber results in a total strain field of ey*~~ = etot-,f - et%f - 2 - - af AT. Substituting into equation 3.32 the sensitivity is

Equation of 3.37 can be written in the form of 3.36 as

where -2

(3 -39)

A fiber subject to simultaneous uni-axial stress and temperature sees a total strain field of

= 8. Substituting into where e;"echyf is related to the lone axial stress component by el

equation 3.32 the sum of individual sensitivities is found,

Page 57: Mechanics and Experiments

Where a fiber is bonded to or embedded into a host material of different thermal expansion

constant ah the sensitivity is computed by recobgizing that the total strain in the host

material is equal to the in-fiber strain

Assuming that it is the host's mechanical strain and temperature that are of interest, in

detail: the phase sensitivity is

Simplif_iring

tivities free

equation 3.43, the sensitivity of the surface mounted fiber in terms of the sensi-

fiber is written as

in which sensitivity sTfed is

Given that the host material thermal expansion is known, equation 3.45 is useM because it

permits the evaluation of the thermal sensitivity of the embedded or surface mounted sensor

in terms of the calibration of the free fiber sensor.

There are many more configurations for which sensitivities can be derived. In all cases the

procedure remains the same- Sirkis has derived the sensitivities in a unified manner for Mach-

Zehnder, IvIichelson, Fabry-Perot , polarimetric, dual-mode and Bragg grating sensors [1 l] [l2].

The latter references detail six thermo-elastic loading cases. Kim, KollL and Springer[33]

and Kim, Ismail, and Springer[33] use the method of Lekhnitskii[34] to describe the strain

transfer from a general anisotropic host to the fiber core of a Fabry-Perot sensor due to

a general six component far field material stress field and a spatially inmriant thermal

perturbation.

Page 58: Mechanics and Experiments

3.2 Experiment a1 evaluation of sensitivities

In the course of this work, experiments are devised to recover the Pockels strain optic coef-

ficients that appear in the sensitivity expressions. Much of the literature uses values of the

PockeIs const ants as reported by Bertholds and Diindliker [35]. The latter reference evaluates

a fiber having a pure silica core and a 8 2 0 3 doped cladding at 633nm interrogation. The

strain measurements that constitute a large part of this thesis are performed in range of

1500n.m to 1560nm. Thus the Pockels constants are evaluated in this range.

Borrelli and Miller[36] provide a procedure to measure the Pockels constants in bulk

silica and evaluate the coefficients pll and p12 simultaneously. The authors construct a

Mach-Zehnder interferometer which includes: in one arm, an element of bulk fused silica.

The silica sample is rnechdcally attached to a piezo-electric crystai which is driven to

induce a mechanical resonance. The mechanical resonance induces an acoustic wave in

the silica sample for which the resulting strain field is computed. Upon interrogation with

Linearly- polarized Light the output intensity of the of the interferometer is dependent on the

birefringence induced in the crystal and the path length through the crystal. Separation of

the two coeEcients depends on the different sensitivities of these effects resulting in different

frequencies of the time response for each of the two effects for a given piezo driving frequency.

Barlow and Payne[37] propose an experiment to determine the stress-optic constant, C,

in optical fibers. The stress-optic coefficient is a combination of the Pockels coefficients

emressed as

where vf is the fiber's Poisson ratio and Ef the Young's modulus of the fiber. The constant

is normalized with the acceleration due to gravity, g. Barlow and Payne [37] apply a twist

to the fiber and measure the birefringence with a polarizer and an analyzer. The authors

also fit the stress optic constant over a wavelength range, l O O O n m to 1600nm assuming the

relative shape of wavelength curve of optical fiber is the same as that of bulk silica, save for

a constant; offset.

Namihira[38] measures the stress-optic coefficient by impressing a birefringent strain state

on the fiber. The birefringence is generated by mechanically loading the fiber with a unidirec-

Page 59: Mechanics and Experiments

tional transverse line load over a length of fiber. Using several wavelengths of interrogation-

different lasers-t he authors generate a curve of the st ress-opt ic constmt over a wavelength

range spanning 633nm to 1580nm. The wavelength dependence is fit to a fourth order

polynomial and compared to[37].

An evaluation of the individual Pockels strain-optic coefficients in an optical fiber is done

by Bertholds and Diindliker[39]. Two experiments are required to lrncouple the specific

dependence of pll and p12- A mechanical stretching of the fiber resulks in a phase change

with one dependence on pll and plz , while a twisting of the fiber, as used in[37], results in a

phase change with a merent strain-optic dependence. These two experiments are performed

on fused silica fiber. described as LT-F1506B, a t 633n.m interrogation.

The majority of grating experiments of the FOSS group employ gratings with center

wavelengths in the approximate range of 1500nm to 1560nm. As such, a n evaluation of

the Pockels coefficients is performed in the wavelength range of interest. To decouple the

independent Pockels coefficients, two experiments need to be employed for which the strain-

optic dependence in uncoupled. Drawing from the literature reviewed above, the stretch test

and the pinch test are selected.

The pull test is applied in the configuration of an all-fiber Michelson, figure 3.3. One arm

of the Michelson is stretched between two stages. The translation is done by an actuator

with a displacement resolution of a single micron. There exists some uncertainly due to

deformation of the fiber/epoxy bond on the translation stage. Thus, it is not true that the

fiber extension is precisely the translation of the stage. A long length of diber, approximately

one meter, is suspended between the translation stages to minimize this effect. A modest

strain extension of 0.5% strain is measured with resolution of I part in approximately 5000.

As one arm of the interferometer is extended, a power fluctuation dependent on the

relative phase difference between the two path lengths is expected as

where V is the fringe visibility. The phase sensitivity to extension is derived to be

Page 60: Mechanics and Experiments

Coherent Source

Coupler r L

Photodetector

Mirror

Fixed Stage w Translating Stage

Figure 3.3: Pull test experiment

Equation 3.48 is different from equation 3-2 by a factor of 2 because of the return trip

of the Michelson interferometer not present in the Mach-Zehnder. As the photodetector's

photocurrent is amplified and measured on a digital oscilloscope. The number of fringes is

manually counted for a measured excursion of the actuator to acquire 3. The second experiment, shown in figure 3.4, consists of a linearly polarized source, a

polarization rotater, a length of fiber on which a birefringence is mechanicalIy impressed,

a polarizer and finally a photodetector. The polarization rotator consists of two paddles

each having a loop of optical fiber. Rotating the paddles relative to each other provides

torsional birefringence to each of the fiber loops, thus controlling the state of polarization.

The input polarization is not i n t idy known. However, the response of the sytem can be

sampled by viewing the transmitted intensity variation on the oscilloscope. The polarization

controller is manipulated so that the fringe visibility is maximized. This adjustment ensures

equal lauch power into each to the principle polarization axis of the portion of fiber beneath

the pressing block. The birefringence of the fiber is ihduced by a mechanical load. The

apparatus employed is shown in figure 3.5. A glass bottomed pressing block diametrically

loads two fibers against a glass plate. Only one fiber is interrogated, the other is a dummy-

A mass a t the end of a lever arm provides force. To restrict forces to the downward direction

only, a ball bearing is used between the lever arm and the top of the pressing block. This

Page 61: Mechanics and Experiments

Polarization

Figure 3.4: Pinch test experiment

loading apparatus is described by Sirkis[40]. The degree of induced birefiingence alters the

distribution of power between polarizations. The polarization distribution is monitored by

insertion of the polarizer preceeding the photodetector.

The Jones calculus is used to derive the intensity variation at the photodetector as a

function of impressed transverse load. A representation of the electric field is made as

where E,; indicates the electric field strength, and cr indicates the axis of interrogation of

the linearly polarized source relative to principal axis of the fiber. The electric fields, Era

and Ef3, beyond the birefringent element are found by mdtipiplying equation 3.49 with an

appropriate transfer matrix describing the phase difference between polarizations given by

the birefringent element. This is written as

where A0 = ,B2 - ,B3 indicates the mechanically induced birefringence. It is useful to re-

express primed field strengths into the principal coordinates of the polarizer (indicated with

Page 62: Mechanics and Experiments

Load ures Downward

Fibers Under Diametric Load

Figure 3.5: Pinch test load apparatus

a double prime). This is done with a conventional rotation of angle 0,

The polarizer allows only one polarization component pass through to the detector, E1'2.

Thus, multiplying Elf2 with its complex conjugate gives the power transmitted to the pho-

todetector. Performing this operation yields

= cos2 a sin2 9 + sin2 cr cos2 B - 2 cos B sin B cos cr sin a cos (APL) . E O 2

(3.52)

Both €J and n are adjusted to maximize the sensitivity of the experiment. This occurs when

8 = a = 45", and equation 3.52, simplifies to

E ' I ~ E ~ 1 = - (1 - cos (A@)) a2 2

in which

4 4 = a p L

is introduced as the phase response of the experiment. In terms of strains, equation 3.54

becomes

Page 63: Mechanics and Experiments

The strain components of the fiber core can be related to P, the load in force per unit length

applied diametrically to the fiber as [U],

Combining equations 3.55 and 3.56 The final phase sensitivity is written as

Load is applied in increments of roughly 10 grams, measured on a balance with a res-

olution of better than a milligram. Output intensity is recorded manually. The choice of

weight increment yields approximately 180 points per 27i phase excursion of the experiment

to assess the sensitivity g. 6

Equations 3.48 and 3.55 are cast as to extract the Pockels coefficients as

The experiment is performed using SMF-28 fiber at two wavelengths, 1506n.m and 1560nm,

using a tunable laser, manufactured by New Focus as a high coherence Linearly polarized

source. The experiment is repeated using a helium-neon source to assess a York fiber (Sm600)

with a numerical aperture suitable to be single moded at 633nm. The results are presented

in table 3.1.

Three parameters are required to perform the computation of equation 3.58, the fibers

Young's modulus of the fiber: Ef , Poisson's ratio, uf , and index of refraction, n. An attempt

is made to evaluate the Young's modulus of SMF fiber. Figure 3.6 shows the experiment in

which a fiber is suspended vertically from a mandrel. The fiber hangs through two guiding

devices to constrain the swinging of the fiber. The guiding devices are simply small holes

drilled into metal plates. Two microscopes are mounted adjacent to each of the guiding

Page 64: Mechanics and Experiments

Microscope

Fiber Suspended fkom Mandrel on Translation Stage

T Guide Hole and Refence Pin

Fiber Gauge Length - lm.

- 4 Guide Hole and Refence Pin on Translation Stage

Clamp

Figure 3.6: Apparatus to measure fiber stiffness

Page 65: Mechanics and Experiments

Table 3.1: Strain and Stress Optic Constant Strmmary

Authour

Borrelli & MilIer[36] Bertholds & Diindliker(39l

Namihira [3 81 Barlow & Payne[37J

This work This work This work

device. The microscopes are-used to determine the fiber deflection under load. *4t the lower

end of the fiber, a specially designed clamp allows a mass to be suspended from the fiber. A

receptacle is attached to this clasp to load the fiber.

The measurement procedure is as follows: Two reference positions are defined, visible

through the upper and lower microscopes respectively. For a coated fiber, two small inden-

tations in the coating are made with a knife. For an uncoated fiber, the coated/uncoated

interfaces are used as the reference positions. The fiber is then loaded. The upper reference

mark on the fiber translated downward due to extension of the fiber portion above the top

mark, and some deformation of the mandrel material. To compensate, the mandrel is trans-

lated upwards until the reference mark coincides with the original position relative to the

guide hole. The lower guide hole is translated downwards to find the lower reference position

on the fiber. The translation of the lower guide pin defines the fiber extension under load.

The resolution of the micrometers that translates the lower guide pin is 5 microns. Coated

and uncoated SMF fibers are tested in this way. Each set of measurements includes ten

individual loading and unloading. The standard deviations of the normalized extensions are

respectively &0.47%; f 0.7 9%, and f 0.37%. The constant of proportionality between strain

and load for the uncoated fiber is 10.75pc/gm &0.06&grn; for the stripped length of fiber

a value of 1 1 . 2 6 ~ ~ * O.OGp~/grn is found. Through Hooke's law,

P11

0.126 0.113

N/A N/A 0.127 0.146 0.150

P12

0.26 0.252

N/A N/A 0.243 0.269 0.272

C bm2/kl

N/A N/A 3.34 3.17 3.09 3.05 3.03

X [=I 633 633 633 1300 633 1506 1560

Material

Code 7940 Silica Silica FiberfLT-FI506Bj GeOz Doped Fiber GeOa Doped Fibre York Sm600 Fibre

SMF-28 SMF-28

Page 66: Mechanics and Experiments

the Young's modulus of the uncoated fiber is found as 70.96GPa f 0-SGPa. Comparing this

result with that of the coated fiber, an estimate of the coating rnoddus , Ec: can be found.

The distribution of force carried in each region can be written as

Assuming that equal strains exist in the coating and fiber, an expression for the coating

modulus can be found as

where A, is the area of the coating. The coating modulus is foun& to be Ec = L18GPa

f 0-5GPa. The error in the coating modulus is, proportionately, much higher than that of

the fiber and can be considered only as an order of magnitude estimate.

In the computation of the Pockels constants, the fiber's Poisson ratio, vf = 0.17[42] and

the index, ii = 1.4682[43], are taken from literature.

Page 67: Mechanics and Experiments

Chapter 4

Interpret at ion of mechanical

interactions: emphasis on strain

transfer between host and fiber sensor

Page 68: Mechanics and Experiments

4.1 Strain transfer considerations for point sensors

The role of the optical fiber sensor is to extract thermc+mechanical information from a

structural host. As indicated in the previous chapter, the optical fiber's spectral and phase

variation from some nominal condition are determined by the thermo-mechanical field vari-

ation inside the sensor. For a measurement to be meaningfully interpreted, the field values

in the sensor must be properly related to the underlying host field values.

4.1.1 Shear lag formulation

The simplest implementation of the Bragg sensor is its deployment as a point sensing tool-

in other words, a sensor reporthg an average axial strain value over the typically short

sensor length. For this configuration to be valid the axial strain in the sensor must closely

approximate the st rain in the material under interrogation. A derivation, originally presented

by Cox[44] is useful as a starting point. This Cox article is originally presented in the context

of the study of the strength of fibrous materials such as paper. The geometry of Cox's

elasticity model is appropriate so that the model can be transferred to the field of optical

fiber strain sensing. Cox indicates the axial stress and strain fields that are expected when

a fiber of finite length is surrounded by an infinite material subject to uni-axial far field

tension: figure 4.1. Both materials are considered isotropic. Figure 4.2 shows a summation

of axial acting forces over an incremental element of the optical fiber sensor. c i indicates in-

fiber normal stress. Over the fiber cross-section this quantity is averaged and taken radially

constant, T{ (z). The in-fiber radial and azimuthal stresses are assumed to be negligible,

Summing the forces acting on the element and s impwng , the following relationship is

found between the mean normal stress and the shear acting on the outer surface,

Figure 4.3 shows the forces acting in the z direction on an infinitesimally thin element of

the surrounding material. Summing yields the following equilibrium expression

Page 69: Mechanics and Experiments

Figure 4.1: Fiber segment in infinite host under uni-axial tension.

Figure 4.2: Sum of axial acting forces in the fiber

Page 70: Mechanics and Experiments

Figure 4.3: Sum of axial acting forces in the surrounding material

in which a,-, again, denotes the mean normal stress over the cross section.

Equations 4-1 and 4.2 are related through the equivalence of shear forces at the sen-

sor/surrounding material boundary. Specifically. TZ (rf, z) = T:= (rf, z)_ The interfacial

shear can be eliminated, and the expression recast as

m 3 a: ( z ) ( z ) r2 - r; rTZ (T: z) = -- -

27- dz dz 2r . Assuming negligible transverse stresses, F: = F{ = 0, equation 4.3 can be written directly

in terms of mean axial strains as:

where Em and Ef indicate the stiffnesses of the surrounding material and the fiber sensor

respectively.

At this point a simplifymg assumption is invoked. Consider the bracketed terms of

equation 4.4. The axial strain gradients are anticipated to be of the same order,

Page 71: Mechanics and Experiments

thus the strongest factor determining the relative magnitude of the two bracketed terms is

the ratio of stihesses: present in the second term. Optical fiber coatings, in the case of a

coated fiber, or typical epoxies: in the instance of a bare embedded fiber are between one

and two orders of magnitude less stiff than the fiber itself, thus

Therefore, the second bracketed term of equation 4.4 can be neglected.

The shear stress is then expressed in terms of the shear modulus and material displace-

Equation 4.7 can be recast &

At this point it is necessary to make a second simplifying assumption. The term is

neglected in comparison with the term. The u-displacements are due, principally: to

Poisson contraction and are less significant than the w-displacements,

The resulting expression is integrated, over r: from the fiber/material boundary, to the

surrounding material radius r,.

The shear modulus is written in terms of the stiffness, G, = 2(5;m1 , resulting in the following

expression.

Equation 4.11 is differentiated with respect to z yielding

Ef e y (rm, Z) - ef (z) = - (1 + v,) - In Em

Page 72: Mechanics and Experiments

The shear lag parameter, AT, encapsulates both the effects of geometry and relative s t i fbas

of the system components [45]. N is written as

E m @=- 1

J% (1 + urn) Zn (T) ' and thus equation 4.12 simplifies to

For the case of a host subject to uniform axial tension, (R, z) = e , . m . The stress

transfer to the fiber happens through shear transfer at the radial interface. Any tensile axial

effects at the fiber host interface can be neglected[46]. Equation 4.14 admits the general

solution

i$ = Aexp [A,r] + B exp [X.r] + co. The eigenvalue of the solution, A, is determined by inspection of equation 4.14

Boundary conditiolzs are invoked that reflect the assumption that there is no axial strain

transfer on the end surfaces of the fiber: Z: (-L) = F: (-L) = 0. Solving for A and B the

strain distribution along the fiber is

Fiewe 4.4

host to fiber.

-m ez (4 = ey',

cosh (7) shows a family of curves indicating the transfer

N values of 0.1, 0.2, and 0.3 are considered.

behaviour of axial strain from

These numbers are chosen to

represent realistic values. (For example, a geometric ratio of = 2: and a stiffness ratio rf

of 3 = % G P ~ / G P ~ corresponds to N = 0.29: while a geometric ratio of = 5 , and a rf

stifhess ratio of E~ = ~ G P ~ / G P ~ 70 corresponds to N = 0.12.) It is clear that the spatial

rate of transfer increases with stiffer surrounding materials. In the worst case presented,

N = 0.1: after an adhesion distance of 50 fiber radii, the ratio of host axial strain to fiber

axial strain is better than 99%. For a 62.5prn radius fiber the required build up distance,

to overcome shear lag, is approximately 3mm. Thus, it is reasonable to assume that the

Page 73: Mechanics and Experiments

10 20 3 0 40

Axial Position zlr,

Figure 4.4: Shear lag plots for N = 0.1, N = 0.2, and N = 0.3.

Page 74: Mechanics and Experiments

Figure 4.5: Common optical fiber sensor installations

axial strains inside the fiber sensor are sufkiently close to host strains if the sensor bonding

configuration accommodates a modest development length the order of several &meters.

Also, figure 4.4 suggests an important phenomenon: the study of which constitutes a

significant portion of this document. The in-fiber strain does not necessarily equal the host

strain. In the case of a finite length sensor: the effect is seen at the end of the sensor? as

a certain development length is required for the sensor to accurately approximate the host

strain- Later in this document, the effect of strain mismatch between mechanical host and

fiber sensor due a spatially varying field is investigated in some depth.

For the case of the distributed sensor: the strain transfer relationship requires rethinking.

Equation 4.17 indicates that the strain field in an embedded sensor does not necessarily

replicate the strain field in the material host. The distributed sensor has been shown to

be capable of measuring arbitrary, sharply varying spatial strain distributions[lO] [47] [16] [17]

[18] [I] [48] [49] [14] [50] with high resolution. The host/sensor interactions are the most relevant

for this type of sensor when high resolution strain measurements are attempted. With this

in mind, the following question naturally arises: What type of strain variations can be

transferred mechanically from host to sensor? And once transferred to the fiber, what are

the optical limits on the extraction of sharp in-fiber elasticity field variations? This chapter

addresses the former part of this question by pursuing three objectives:

Page 75: Mechanics and Experiments

1. establish criteria to define the quality of axial strain transfer between a substrate and

an optical fiber sensor,

2. describe and quantify the parameters influencing this traasfer, and,

3. for both the embedded and surface bonded case, present the expected performance of

the sensor.

4.2 Axial strain transfer from host to fiber: Finite el-

ement -models

The initial work in this thesis towards addressing the question of the quality of strain transfer

is performed with two finite element models; one for the surface bonded case and one for the

embedded case. The two models will be described in turn.

4.2.1 Surface bonded sensor model

The surface bonded model is a hybrid model of Reissner/Mindlin and quasi-three-dimensional

elements. Reissner/Mindlin elements are typically used to model plates of composite ma-

terials having several laminates of different material properties. Conventionally, different

material properties for the different laminates are due to different ply orientations of the

same anisotropic composite material. Figure 4.6 shows the position of the Reissner/Mindlin

elements in the surface bonded model. In this implementation, different laminates, within

elements are assigned different mechanical properties, not due to ply orientation, but due

to individual laminates representing either optical fiber sensor or surrounding adhesive. It

is common for composite materials to be anisotropic such that the stiffness in the direction

parallel to the reinforcing fibers is approximately two orders of magnitude greater than the

stiffness perpendicular to the reinforcements. When comparing the optkd fiber sensor stiff-

ness (70GPa) with that of typical structural epoxies; (range of 2 to 5 GPa[51] [52]) a similar

disparity of mechanical properties is found. Reissner/Mindlin elements also model the un-

derlying substrate. Figure 4.6 illustrates the co-ordinate system used for Reissner/Mindlin

Page 76: Mechanics and Experiments

Reissner1Midli.n plate elements model stacking of fiber and adhesive

Reis sner/Midlin plate elements model surrounding adhesive

Quasi-three-dimensional elements model thin

I Reissner/Midlin plate elements model underlying sub strate

Figure 4.6: Model for surface mounted fiber.

analysis. A quasi-three-dimensional element is used to represent the thin layer of adhesive

that lies between the fiber and the substrate l.

The following assumptions describe Reissner-Mindlin plate theory:

The model can accommodate structures of moderate thickness

The midplane displacements ul: up; and u3 are small compared to structure thickness.

The in-plane strains el, e2, and el*, strains are small compared with unity.

The transverse normal strain, es, is neglected.

h he computer code used to generate results for the surface bonded model was written by colleague and

co-authour[53J Guillaume Renaud.

Page 77: Mechanics and Experiments

The transverse shear stresses, a13 and 023, me parabolic in x3.

The Reissner/Mindlin elements are two dimensional, having nodes along the mid-plane

only. Five degrees of freedom are sought in the Reissner/Mindlin formulation: mid-plane dis-

placements, ul (xl, x2), u2 (51, x2), and 113 (xi, x2), and two variables 91 (xi, ~ 2 ) and 02 ( ~ 1 ~ x 2 )

that provide an indication of rotation. 6 indicates the vector of midplane variables

Displacements as functions of three dimensions, (indicated with an overbar, .Ltl (xl, x2, x3)), - u2 (xl. 52: x3) and us (xl: x2, x3) are recovered from the mid-plane strains and the two

rotation-like terms, and 02, by the following relationships:

Strains, in turn are derived from the displacements of equation 4.19

To facilitate the notation of the stress resultants it is useful to introduce the following

not ation:

Page 78: Mechanics and Experiments

Equation 4.21 is further condensed to

The strain vector, {el, can be expressed in terms of the displacements, 6,

The matrix of operators in equation 4.23 is labeled [L]. As will be seen, the matrix [L] acts

on the interpolation functions of the discretized dement.

Relevant stress components appear as

I 4 =

Page 79: Mechanics and Experiments

Stress, N , moment, M, and transverse shear, Q, resultants are obtained by sllmming stresses

over the faces of the element,

The transverse integration of equation 4.26 is expressed in terms of the stresses of equa-

tion 4.22 as:

The inner sub-matrices, G,, bij: 6, and Sj are representation of the aggregate behaviour

of a multi-layered composite material element. The constructions of Gj; bij, dij7 and Eij

accommodate an element having an arbitrary number of plies, each with generally anisotropic

mechanical properties. Explicit descriptions of how to compute the entries in these matrices

can be found in composite materials tex*s such as[54]. Equation 4.27 is rewritten as

The governing expression for a single element is found by considering a virtual displace-

ment field, d (61, imposed on the element. The work done against the element boundary by

external forces, moments? and transverse shears must be stored inside the element as internal

strain energy. This equivalency, in terms of the virtual displacement, is written as

where {r} represents the stresses acting on the element boundary. {r} includes the same

content of normal stresses, moments: and transverse shears as the vector, {a). Inserting

Page 80: Mechanics and Experiments

Figure 4.7: Node distribution across Reissner/Mindlin element.

the stress resultants of equation 4.26 into equation 4.29 reduces the right h a d side of the

equation to a two dimensional integral,

Then, expressing the stress resultants in terms of strains, and strains in terms of displace-

ments, the following is found

At this point it is necessary to discretize the element. This is done by assigning discrete

nodes to the element. Consider first the square element in the E-v plane bounded by -1 < < < l a n d - 1 < q < 1 .

Figure 4.7 shows the distribution of nodes across this element. The coordinates XI and

x2 are mapped into the <-9 plane through interpolation functions &

Page 81: Mechanics and Experiments

The continuous displacements are approximated by a discrete set of vector of nodal displace-

ments, {6Ie and a matrix of interpolation functions, [@]. Specifically:

where,

for which each of the sixteen interpolations functions is given as

The k-th interpolation function has the property that it has the value of unity at the k-th

node and zero at all other nodes-

The physical problem requires casting from the XI-xa plane to the J-r) plane. This is

accomplished with the coordinate transform,

The two by two matrix of derivatives in equation 4.36 is called the Jacobian of the transfor-

mation, [J] . Area transformations scale between the two coordinate systems as

Line integrals

coordinates is

are required to sum boundary loads. An infinitesimal length ds in Cartesian

expressed as d s = J ( d z l ) 2 + ( d ~ ~ ) ~ . In the <-) space this length becomes

Thus, combining equations 4.29 through 4.36, and introducing [B] = [L] [a], a discrete

system of equations for the single element is written as

Page 82: Mechanics and Experiments

A c

a Reis sner/Mindlin Element

2 @ Quasi-three- ,E dimensional

3 , element

ReissnerMindlin Element

Fiowe 4.8: Quasi-t hree-dimensional adhesive element node positions.

Canceling the virtual displacement from either side of equation 4.40 the governing equation

for the single Reissner/Edlin is found as

with

and

{F}' = /' [@lT {/f -- {r) dZ3} d ~ . 2

The adhesive layer is modeled as a quasi-threedimensional element with virtual nodes

as proposed by Renaud and Hansen[55] [56]. Such an element is shown in figure 4.9. The

nodes are described as 'virtual' because their displacements can be fully determined with

Page 83: Mechanics and Experiments

Fiewe 4.9 : Quasi- three-dimensional adhesive element.

knowledge of the nodal degrees of fkeedom of The Reissner/Mindlin elements both above and

below the adhesive layer elements. Ultimately, the addition of the adhesive element does not

increase the size of the global stifhess matrix. Assuring continuity, displacements on the

top surface of the adhesive element are assigned values corresponding to the displacement

of the bottom surface of the upper Reissner/Mindlin element. Analogously, the adhesive's

lower surface displacements are assigned values determined by the lower Reissner/&hdlin

element. Figure 4.8 illustrates the relationship.

Three degrees of heedom are accommodated in the quasi-three-dimensional formulation;

displacements, u1: u2; and 213,

The relationship between displacements on the top surface of the adhesive element are ex-

pressed in terms of the degrees of freedom of the upper Reissner/Mindlin element through

Page 84: Mechanics and Experiments

Analogously, the displacements of the bottom surface of the adhesive element are related to

the underlyimg Reissner/Mindlin element through

and

Expressioms 4.45 and 4-46 are re-cast in matrix form as

respectively.

Strains ape expressed in terms of displacements in the conventional manner:

Again, a mat ~ i x of operators relates st rains and displacements,

The stress vector includes five relevant components,

Page 85: Mechanics and Experiments

Stresses relate to s t r e through an isotropic stiffness matrix,

To determine the governing equation for the element, a virtual displacement is imagined

to act on the element and the work done at the element boundary is equated to the internal

energy,

(-2.53)

With a view to discretization: the element is transformed into the <: q, and < space

as indicated in figure 4.8. The three dimensional represent at ion requires three-dimensional

interpolation functions. An interpolation function is assigned to each of the 33 nodes of the

adhesive element. The interpolation functions are constructed as Lagrange polynomials and

are cubic in the J-77 plane and linear through the thickness c. The interpolation functions

appear as

The interpolation functions relate the discrete to the continuous representation of mriables.

For example, dispiacements are converted through

Equation 4.53 is discretized to read:

It is desired to write the contributions of the adhesive nodes in terms of the Reissner/Mindlin

nodes. This is performed by invoking the transformations of equations 4.45 and 4.46. Equa-

tion 4.56 is rewritten as

Page 86: Mechanics and Experiments
Page 87: Mechanics and Experiments

4 7 3 Eightnode,isoparametric - axi-symmetric elements

1

Figure 4.10: Axi-symmetric finite element model.

Page 88: Mechanics and Experiments

The

and

displacements are shortened in vector form to

the strains are related through the operator matrix [L] to the displacements, {e} =

[L] (6). [L] is written as

[Ll

Stresses in axisymmetric coordinates appear as

The stresses are related to strains through an isotropic stiffness matrix; [Dl; analogous to

equation 4-52.

The governing element equation is derived: as for the surface bonded model, by consid-

ering the effect of a virtual displacement. The redefinitions of [L]; [Dl; { T } : m d (6) for

the cylindrical element permit the governing energy balance to be written analogously to

equations 4.29 and 4-53.

The discretization is done with an eight node isoparametric serendipity element. The

nodal enumeration is also show in figure 4.10. r and z are mapped to the J-7 space as

As before, the k-th interpolation function, pk, equals unity at the k-th node and is zero at

a,ll other nodes. The individual interpolation functions are

Page 89: Mechanics and Experiments

Again ( 6 ) = [a] The coordinate transform between the two systems is described by

where

Line integrals are transformed in the same manner as indicated in equation 4.38.

Rewriting ?? in discretized form with [B] = [L] [O]:

is obtained. The element stiffness matrix is computed as

and the element force vector is resolved as

{Fie = { [a] {r} rds. s

As before, a global set of equations is found by summing the nodal contributions in the

global indexing scheme

[KIg {5)g =

Page 90: Mechanics and Experiments

in which

and

4.2.3 Finite element analysis of linear strain gradient

To gain preliminary appreciation for the strain transfer problem, the behaviour of a fiber/host

structure to which a linear strain gradient is applied is investigated with the two finite element

configurations described in section 4.2 and the degree

sensor reflects the strain gradient in the substrate is

described mathematically as-

where ey indicates the maximum strain

to which the strain gradient in the

assessed. This axial strain field is

attained by the gradient, and L denotes the length

of the gradient. The axial dimension, XI: is normalized with the fiber radius, rf.

For both embedded and surface bonded descriptions, boundary conditions are most easily

imposed in terms of displacements in the substrate or surrounding material. Such displace-

ments, uy (xl), can be found by integrating the strains of equation 4.78. In this instance

where the axial displacement has been non-dimensionalized with rf. The strain and dis-

placement relationships are illustrated in figure 4.11. -

Given an axial strain gradient in the substrate or-surrounding material, analysis can

recover the axial strain profile in the fiber optic sensor. A typical example of the strain

transfer is presented qualitatively in figure 4.12. Compared are the impressed axial strain

field as a function of axial position in the substrate, ey (xl) and the resultant axial strain

Page 91: Mechanics and Experiments

Axial Strain Imposed in Substrate

Axial Displacement Imposed in Substrate

Figure 4.11 : Imposed st rains and displacements

Page 92: Mechanics and Experiments

Figure 4-12: Equivalent transferred linear strain gradient

field in the sensor core; e{ (~1). The impressed gradient in the host; following equation 4.78,

is sharp and well defined; the strain distribution in the sensor core cannot be assumed to

have the same degree of 'sharpness.'

As axial strain in the sensor core does not necessarily vary linearly with position, it

is useful to describe an equivalent linear slope for purposes of comparing the slope of the

substrate gradient to the gradient transmitted to the sensor core. The equivalent slope

defined by the line in figure 4.12 is selected by demanding that the area beneath the line - AB and the area beneath the ef (zl) curve, from -oo < rf < - L. 2rf are equal. Mathematically,

On a plot in which the maximum strain excursion is normalized with ey = 1, the d u e of 6

will be *. Further; a can be extracted from the solution of the elasticity field in the following

manner. The left hand side of equation 4.80 can be related to the axial displacements of the

Page 93: Mechanics and Experiments

sensor core, u{ (r l). Specifically,

Far to the left of the gradient,

So the function of integration is zero, 3(x2, x3) = 0. Therefore, the value of a can be written

The quality of transfer will be described as follows. The length by which the transferred

equivalent gradient in the fiber core exceeds the length of the imposed gradient in the host

structure, normalized with the fiber radius is indicated as Referring again to figure 4-12? rf

it is seen that

which: considering equation 4.83, reduces to

Equation 4.85 proposes a simple formula to estimate the gradient 'smoothing' in terms of

parameters readily available from the finite element solution.

Edge displacements as described by equation 4.79 are set at r,. r, is taken to be the

location at which the host strain field can be considered unaffected by the presence of the

optical fiber sensor.

At the left face of the bod% the asial displacements are held at zero,

and the axial displacement on the right face is held constant in the 22-23 plane at XI,-,

No other loads or displacements are imposed. Thus, the appropriate Poisson contraction of

the specimen is expected.

Page 94: Mechanics and Experiments

The length of either the surface mounted or embedded model, delimited between xlvmin

and x1,-, is chosen such that the elasticity state a t either end of the body is uncoupled

kern the presence of the gradient. In particular, it is asserted that axial strain in the fiber

core matches axial strain in the substrate boundary to better than 0.1%. SpecScally.

The condition expressed in equation 4.88 ensures that the region which is included in the

model is sufficient to properly describe the entire effect of the imposed strain gradient.

The surface mounted cod?guration, as described in sectian 4.2.1, is viewed as two plates

bonded together with a thin adhesive layer. The top section comprising the optical fiber

sensor and some surrounding adhesive is modeled to vary linearly &om its maximum height

to zero at a distance r, from the a of symmetry. This group of elements is laminated in

order to approximate the properties of the fiber/adhesive set. The circular cross-section is

approximated by different stackings of fiber and bonding materials for each element. The

relative thickness of each ply is chosen in order to best represent the cross-sectional area of

the fiber relative to the cross-sectional area of adhesive on which it sits. For a given element,

the thickness of the adhesive layer is selected so that the cross-sectional area of the adhesive

matches the cross sectional area of the adhesive that would have been present below a circular

fiber over the width of the element. The lateral tapering of t he bond is modeled with varying

element thicknesses imposed at the Gauss points. The thin adhesive layer is assigned a

thickness of l p m . The underlying substrate, in which the boundary conditions are enforced

is modeled as another group of Reissner/bIindlin plate elements. The fiber, considered

isotropic, has radius rf, Young's modulus Ef and, Poisson's ratio, vf . Equivalently, the

adhesive material has Em, and v,. Recalling equation 4.46, the displacements imposed

in the substrate define fully the displacements of the bottom surface of the adhesive, the

material properties of the lower Reissner/Mindlin plate are not parameters of the problem

and are assigned arbitrary values.

Given the thickness distribution of the adhesive material, the most likely breach of the

linear elastic and perfectly adhered assumptions will occur as shear failure in the adhesive

as high intensity shear strains develop in the region where contact is made between the

Page 95: Mechanics and Experiments

Lamina assigned zero structural stiffness

Figure 4.13: Location of lamina modeled as having no stiffness

substrate and the fiber in the presence of a strain gradient. This potentially jeopardizes the

integrity of the bond and the assumption that perfect adhesion is present. A high gradient

test case is evaluated to assess the assumption. The substrate is typically a structural epoxy L having a range of shear stress limit between 10 and 30MPa[51] [52j. A step gradient, ;T = 0

climbing from 0 to 1000 pe, is imposed in the substrate. This load case results in an adhesive

shear stress above its strength limit. As such, for the surface model, it is assumed that an

adhesive failure of the width of the fiber occurs over the whole length of the model. Figure 4.13

indicates the location of the lamina assigned no structural stifiess. The load transfer from

the substrate to the fiber is therefore exclusively through the lateral portion of the bond

cross-sections. This is incorporated into the model by assuming no strength of the adhesive

in the region directly beneath the fiber. This modification helps assure that the quality of

transfer is conservatively estimated.

The embedded geometry is modeled as an axisymmetric system consisting of two concen-

tric regions, representing the optical fiber sensor surrounded by the substrate material. The

fiber, considered isotropic, has radius r f : Young's modulus Ef and, Poisson's ratio, vf. The

substrate is described by rm, Em, and v,. The finite element formulation of section 4.2.2 is

employed.

Page 96: Mechanics and Experiments

Figure 4.14: Shear stress distribution at fiber/surrounding material interface for embedded

installation. = 1.05, 2 = 70: $ = 0

To evaluate the potential shear failure of the surrounding material for the embedded

model, the same high gradient load case is tested as was done for the surface bonded case.

-4 l 0 0 0 p ~ step gradient is imposed in close proximity.to the fiber, = 1.05. The same

compliant surrounding material is chosen with = 70. Figure 4.14 depicts the shear

stress, in the substrate, at the substrate fiber/host interface for the previously described

case. It can be seen that the highest stress found in -:that substrate is 8MPa. This case

indicates that shear failure for the embedded installation requires high gradients, compliant

bonds, and very small values: and is unlikely with typical structural strain gradients.

The parameters that impact the quality of transfer are

Page 97: Mechanics and Experiments

a the relative stifbesses of the fiber and the surrounding material,

the geometry of the surrounding materials relative to the fiber, and

the degree to which the gradient can be called 'steep.'

For the embedded installation, the geometric parameter of interest is simply the ratio of

the radii of the cylindrical host region and fiber regions. In the surface bonded installation

the geometry is characterized by the half-bond length, divided by the fiber radius. In both

cases this ratio is expressed as k. Material parameters are also easily described by the ratio rf

of sensor s t i5ess to bond stifhess 2. The gradient is given some consideration. The degree to which a gradient can be consid-

ered 'steep', at first, would seem a combination of two factors: the maximum strain achieved

by the gradient and the length, relative to some characteristic length, over which the max-

imum strain is achieved. The rate of change of axial strain with axial position, normalized

using the fiber radius is rf 2. Perhaps surprisingly! for the assessment of the strain transfer

the value of the parameter ey has no impact on the degree of strain quality of transfer as

defined by equation 4.85.

This can be argued by considering the general finite element formulation. In both surface

mounted and embedded formulations, a finite element analysis generates a global system of

equations such as equation 4.61 and 4.75,

where [KIg is dependent on material properties and geometry, (6)' is a vector of nodal

displacements for which equation 4.89 is solved, and { F j g is a load vector containing the

nodal forces acting on the system.

Only displacement boundary conditions are imposed in either the surface mounted or

embedded study, no explicit nodal forces are applied. So initially a system such as

is found. To leave the expression as such is clearly inadequate because some of the dis-

placements have been explicitly prescribed. The prescribed displacements at the boundary

Page 98: Mechanics and Experiments

are removed from the (6)' vector and shifted to the right han~d side of equation 4.90 after

having been multiplied by the appropriate entry of the stiffness matrix. This will result in a

system with fewer degrees of freedom and with a new force vectoor, derived from the imposed

displacements

[K'l{6') = {F'} * (4.9 1)

where [Kr] is the reduced s t f i e s s matrix, {6'} represents the vector of unknown displace-

ments and {F') denotes the force vector equivalent to the prescriibed displacement conditions

at the nodes.

The force vector {Ft ) is constructed from imposed displacements and corresponding

stiffness matrix elements. For example, in a system with N degrees of freedom with boundary

displacements imposed at nodes ni, a typical force vector would be written as

It can be seen from equation 4.92 that the implicit load vectorr is directly proportional to

the magnitude of the imposed displacements. Recalling equatiom 4.79 the displacements are

directly proportional to the ey term of the prescribed gradient. Thus, equation 4.91 can be

arbitrarily scaled, by choice of e$ Thus, a potent i d y counter-imtuitive conclusion is found:

The maximum height of the gradient does not affect the quality- of transfer. A larger ey will

scale the vertical axis of figure 4.12 without changing the qualitzy of transfer. Thus, for two

situations both with a gradient of length L and a fiber radius rf, and different maximum

strain excursions: both will show the same degree of quality of -transfer, to the elastic limit

of the structure.

The treatment of all further gradients in this study will be forr unit e;. Thus, the measure

of gradient steepness is simply T. The finite element analysis is performed for a range of values of the parameters z, p,

and 3. For each case, a value of is recovered to assess the equality of transfer. Typical

epoxy stifhesses range between 2GPa and 5GPa[52] [56]. Thus a t the two extremes, two

Page 99: Mechanics and Experiments

values of 2 are analyzed, 14 and 35. The range of gradients considered are

For a typical 62.5pm radius fiber this represents a range of gradient lengths between 0 (a

step function) and lcm. The range of bond sizes is

Fiewe 4-15 shows an indication of quality of gradient transfer over the parameter space.

The plot is given in terms of rf / 4L so that higher values indicate better quality of transfer.

It is seen that stiffer substrate materials, less steep gradients, and larger wider bond profiles

lead to better quality of trimsfer. The effect of increasing bond width is limited. For

geometric ratios larger than % > 16 the quality of transfer does not increase for further

increases in geometric ratio. This result indicates that, when surface bonding optical fibers

to a substrate: there is little advantage to increasing the epoxy coverage over the fiber to a

larger ratio than > 16. rf

Individual contour plots for the surface mounted installation are presented in fi,o;ures 4.16

and 4.17 for stiffness ratios of = 114 and = 35 respectively.

The quality of transfer over the parameter space for the embedded model is plotted in

figure 4.18. The trend is different than seen with the surface mounted model. The transfer

quality decreases monotonically with geometric ratio. This is expected considering that

larger and larger geometric ratios indicate that the imposed gradient exists at farther and

farther distances from the sensor itself. Similarly to the surface mounted configuration,

stiffer substrate materials and less steep gradients lead to better strain transfer.

The two-dimensional contour plots corresponding to the two stiffness ratios are presented

in figures 4.19 and 4.20. These charts can be used as design tools for determining an adequate

bond width for an expected transfer efficiency. For example, consider a situation in which a

gradient of length 5mm exists in an epoxy of stifFness 2GPa 0.5mm away from an embedded

fiber sensor of radius 62.5pm and stifhess 7OGpa. This corresponds to values of = 35

and = 8- Figure 4.20 indicates a normalized elogation value of approximately 12. Thus, a

gradient of length 5mm in the substrate is reflected in the fiber as a gradient of length 5.8mm.

Page 100: Mechanics and Experiments

Figure 4.15: Quality of transfer for surface mounted installation (Vertical axis represents

5 x 1000)

Page 101: Mechanics and Experiments

5.0 10.0 15.0 20.0 25.0 30.0 35.0 Geometric Ratio r ,/rf

Figure 4.16: Gradient elongation, e, for stiffness ratio of 2 = 11 (Surface mounted)

Page 102: Mechanics and Experiments

5.0 10.0 15.0 20.0 25.0 30.0 35.0 Geometric Ratio r ,/r,

Figure 4.17: Gradient elongation, u, for stiffness ratio of = 35 (Surface mounted) rf

Page 103: Mechanics and Experiments

Figure 4.18: Quality of transfer for embedded installation (Vertical axis represents & x 1000)

Page 104: Mechanics and Experiments

5.0 10.0 15.0 20.0 25.0 30-0 35.0 Geometric Ratio r ,/r,

Figure 4.19: Gradient elongation, 5, for stiffness ratio of 2 = 14 (Embedded)

Page 105: Mechanics and Experiments

5.0 10.0 15.0 20.0 25.0 30.0 35.0 Geometric Ratio r ,/rf

Figure 4.20: Gradient elongation, e; for stiffness ratio of = 35 (Embedded)

Page 106: Mechanics and Experiments

ez

Figure 4.21: Illustration of problem: axial strain transfer from host to sensor

If a stiffer epoxy is selected, 2 = 14; it is found that the gradient is better transfered such

that an equivalent gradient of 5.4- is found in the fiber.

4.3 Transfer of arbitrary strain distribution

The finite element method of section 4.2 is a stable md proven method. However, pro-

gramming the finite element code requires recoding boundary conditions on a case by case

basis. The matrix inverison required to implement the solution is computationally extensive.

Naturally. an analytical approach would be desirable to. illuminate parameter dependencies

and to provide a 'feel' as to the behaviour of the system. Figure 4.21 illustrates the problem

to which an analytical description is sought. In cylindrical co-ordinates, (r: z, 0) an infinite

fiber is bounded by an axi-symmetric region of surrounding material. At some radius, r,, in

the surrounding material and arbitrary strain distribution is specified. In view of the results

Page 107: Mechanics and Experiments

of section 4.2.3 it is expected that the fiber will be subject to different axial field than the

surrounding material.

Consider the original derivation of Cox, presented in section 4.1.1. Continuing &om

equation 4.14, the . At r = rm an axially varying strain field exists, e? (r,, 2 ) . In the instance

of an optical fiber sensor with a relatively compliant coating? rm naturally corresponds to the

outer coating radius, where ep (r,; z) denotes the strain field present at that boundary[l3].

For the more general embedded situation, r, is interpreted as the distance from the fiber

core at which er (r,, z ) is known. T o determine r, it is necessary to first estimate the

effective shear lag parameter N of the fibre/surrounding material geometry and stfiesses.

This can be performed by comparison to finite element results or experimental data[45].

Equation 4.14, for the case of an arbitrary far field strain is solved by Fourier decompe

sition. The solution of equation 4.14, E: ( r ) , will be composed of a weighted distribution of

spatial wavelengths: A,. The spatial wavenumber is introduced to describe a single Fourier

component of the solution where k = $. The Fourier transform relates the strain descrip-

tions horn the z to the k space as

00

E! ( k ) = /__ < ( 2 ) exp ( -2r ikz ) d z , and roo

a: (rm, k ) = J- e y (T,, Z ) exp ( -2r ikz ) dz- 00

The inverse relationships appear as

00 < ( 2 ) = 1- e! (k) exp (2rrikr) dk, and - -

e: (T,, Z) = /-_ &r (r,, k ) exp ( k i k r ) dk.

The Fourier transform of equation 4.14 yields the expression

From equation 4.97, the transfer function of the system, H ( k ) , is recovered,

6: (k) - - 1 H (k) = e y ( r m j k ) I+ ( .Y , )~ '

Page 108: Mechanics and Experiments

Using equation 4.13 to

The transfer function,

distribution, of spatial

expand 4.98 in t e n of the shear lag parameter yields 1 I H (k) =

1 + 47r2k2r; ((I +- L..) 1D. (T) (z) X ( k ) , indicates the proportion by which a purely sinusoidal strain

frequency k, will be transferred horn surrounding material to fiber

as function of N and k. As k increases, indicating shorter spatial wavelength content, the

transfer gets progressively weaker. This supports the intuitive notion that quickly varying

strain fields with high wavenumber content, are more poorly transfered than slowly varying

strain fields. At the extreme limits, a constant strain field, ey (r,, z ) = c, is perfectly trans-

ferred, H (k = 0) =-I, while a very quickly varying harmonic strain field in the surrounding

material (high k) is fully attenuated H (k --+ oo) = 0. Further, sharp strain variations in

the surrounding material will be 'smoothed7 as they are transferred to the fiber sensor. The

high wavenumber components will be strongly attenuated relative to the low wavenumber

components.

In the section 4-2.3 individual cases are made to suggest the parameters on which lin-

ear gradient transfer is dependent. It is argued that they are material stiffness ratio, g; geometric ratio; and an indication of gradient steepness, s. For the arbitrary case the

rf

parametric dependence is now revealed in equation 4.99. The stiffness and geometric ratios

appear explicitly, and the spatial wavelength content is used to more precisely define the

degree to which a strain field varies in space.

Finally, the expression by which the in-fiber strain function can be found is written as

( z ) = Jm H (k) C (rm, k) exp (2x ik t ) dk, -M

where Ey (r,, k) is found from equation 4.95. Equation 4.100 is written in detail as a function

of the e," (r,, z),

x exp (27rikz) dk. (4.101)

Figure 4.22 shows the transfer function H (k) . Also included in figure 4.22 are the results

from a the conventional FEM analysis of the same physical configuration. FEM calculations

Page 109: Mechanics and Experiments

Figure 4.22: Transfer function, H ( k ) dependence on spatial frequency k

Page 110: Mechanics and Experiments

were made using the

boundary conditions

T m 7

geometry of figure 4.21 and the model of section 4.2.2. DispIacement

are imposed at r = r,. A strain of the following form is specified a t

e," ( z ) = eEo cos (27rkz) .

Integrating equation 4.102 the displacements to impose at the model boundaq are found as

z m

w m ( z ) = / e y (2') dzr = % sin (27rkz) . 27ik

In this study, N values of 0.1: 0.2, and 0.3 are considered. The trend towards poor

transfer with increasing wavenumber is clear, and the plots also indicate that higher N values increase the quality of transfer.

There exists significant agreement between this work and the FEM results, but limitations

originating from the simplifjying assumptions are noted. Specifically, disagreement increases

as N increases at high wavenumber. For example, a t k = 1000m-', and N = 0.3 an

attenuation of 0.37 is predicted by this work, while the finite element analysis predicts

0.31. At the same k, for the smaller Af = 0.1: the predicted peak suppression is 0.06

(equation 4.101) compared to 0.05 (FEM) . A s the wavenumber increases, the transfered

profile is considerably attenuated, compromising the assertion that the strain gradients are

of the same order (Recall assumption of equation 4.5). As N increases, the assumption that

the material stifiess ratio is much smaller than unity becomes less robust.

A family of curves is presented in figure 4.23. This set of plots compares the transfer of

a single period of the function e? (R, z ) = cos(27ikz) for different k with the induced strain

field in the fiber core E{ ( 2 ) . Seven plots appear in each graph: the plot of er., (R, Z ) and

the solution of equation 4.100 for the three N and the three corresponding FEM solutions.

Over the z range, there is agreement between this work and FEM. The weak transfer of high

k surrounding material strain fields (short As) is apparent. The tendency of higher N to

improve the transfer is also seen. The attenuation in peak to peak magnitudes of e F (R, Z )

compared with E: ( z ) is precisely the physical meaning of the transfer function, H (k) : a t a

specific wavenumber k.

Figure 4.24 compares the transfer of a family of Gaussian strain profiles of e? (R, z). The

Gaussian shape is studied because many realistic applications of fiber optic sensing attempt

Page 111: Mechanics and Experiments

Fibwe 4.23: Transfer of far field sinusoids 97

Page 112: Mechanics and Experiments

to extract information from strain distributions that, Like Gaussians, can be characterized

in terms of a peak strain and a full width at half maximum ( F W B ) , 4zm. The strain

distributions due to holes: advancing cracks, or other structural imperfections can be de-

scribed in terms of a maximum strain and a E m . A range of F W M are investigated

from 4 z m = 0.lm.m to 42" = 5mm. The imposed strain field at the r = r, boundary is,

I d

In this case, a numerical integration of equation 4.104 is employed to compute the required

boundary displacements, z

w m (rm,z) = / ey (r,, 2') dr'.

4 and the The transfer can be described in terms of the degree of peak suppression eF(rm,c proportion by which the FWHM broadens, g.

For very narrow Arm, the peak suppression is significant. For example at Azm = 0. lmm

the peak is suppressed from unity to 0-02, while the in-fiber FWHM is estimated to be

~ z f = Smm, an expansion by a factor of 90. The latter result indicates that sharp, narrow,

strain 'spikes' due to structurd imperfections axe poorly transferred to the fiber sensor core.

The results of figure 4.24 are synthesized in figure 4.25. The peak suppression and Azm

increase are shown as functions of the imposed FWHM. for the three values of Af. The solution of equation 4.100 is performed with a fast Fourier transform (FFT) alge

rithm. The FFT is considerably faster than an analogous FEM solution, the latter requiring

a matrix inversion. Arbitrary functions of eF (r,, r ) , either periodic and aperiodic are cal-

culated with the same computer code. The proposed f6rmulation allows the calculation of

the axial strain field in the fiber from an arbitrary sti& profile in the host provided the

shear-lag parameter, N, can be determined.

4.4 Equivalent embedded geometric ratio for surface

bonded sensor

The mathematical representation of equation 4.101 is useful for the embedded case. An

extension to the case of the surface mounted fiber is attempted in this section. As was

Page 113: Mechanics and Experiments

Figure 4.24: Transfer of far field Gaussians 99

Page 114: Mechanics and Experiments

Profile F'WHM Azm (mm)

0.0 0.5 1 .O 1.5 2.0 2.5 ProfiIe FWHM A Z ~ (mm)

Figure 4.25: Peak suppression and broadening of Gaussian far field strain profile

Page 115: Mechanics and Experiments

Adhesive layer of thickness "t"

Figure 4.26: Surface bonded installation modelled as embebbed installation with equivalent

radius

argued in section 4.2 and explicitly derived in section 4.3 the shear transfer from host to

fiber in the embedded case is dependent on three factors-stiffness ratio between host and

fiber, the geometric ratio indicating the location of the far field relative to the fiber, and the

degree to which. the field varies rapidly in space. The first two factors, stifbess and geometry

are included in the shear lag parameter N. The quality of transfer is expected to vary with

the degree of far field variation. To determine an estimate for the behaviour of the surface

mounted sensor, the following approach is used: The model of section 4.2.1 is employed to

model the transfer of periodic strain variations (specifk k values) from the substrate to the

optical fiber. Then, the attenuation incured during mechanical transfer is extracted from

the finite element solution. Recalling that the transfer function H (k) of equation 4.99 is

precisely the ratio of the maximum amplitudes of the far field and transferred profiles, the

transfer function is directly computed kom the results of the surface bonded model. Further

an equivalent embedded far field radius is estimated such that the surface mounted behaviour

emulates the behaviour of an embedded situation with a fictitious equivalent far-field radius.

Thus in equation 4.99 the only unknown variable is the geometric ratio % which is

estimated from the transfer efficiency. Table 4.1 indicates the results obtained in this analysis.

For a sensible range of stiffness ratios, a range of transfer efficiencies between near unity and

50% peak suppression is found. Figure 4.26 illustrates the idea of an equivalent geometric

ratio. The surface bonded model is considered to be equivalent in strain transfer performance

to an embedded sensor with the equivalent geometric ratio. The results of table 4.1 report

an equivalent geometric ratio of between 1.16 and 1.26. For a fiber of radius 62.5pm the

Page 116: Mechanics and Experiments

ratio period transfer efficiency Estimated equivalent

geometric ratio 5

rf

Table 4.1: Equaivalent radius estimate for surface mounted situation

equivalent radius is estimated to be between 73pm and 79pm-a distance of between 10pm

and 17pm from the outer surface of the fiber. The finite element model of the surface bonded

configuration considers the fiber to sit on a layer of adhesive one micron thick. The equivalent

embedded radius is estimated to be between ten and seventeen times this value.

For the balance of this work, the surface bonded case is assigned the most conserva-

tive equivalent geometric ratio found kom the finite element simulations, = 1.26. The +f

Poisson's ratio of the adhesive is taken to be 0.25. The modulus of the adhesive is also as-

signed a conservative value of 1GPa. Thus the shear lag parameter for the surface mounted

configuration is computed to be

For the remainder of this document, N is taken as 0.22.

It is noted that the analysis of this section is limited to axial strain transfer. There will

be an amount of bi-axial stress induced in the fiber due to the asymmetry of the bonding

configuration. For example, when the substrate/fiber sensor system is under axial tension,

transverse Poisson contraction is expected. The fiber admits a low Poisson ratio and has

small a5ni ty for contraction. The host/adhesive set has a stronger desire for contraction. As

Page 117: Mechanics and Experiments

displacements must be continuous, the relative contractions are equilibrated by an internal

bi-axial transverse stress field-

4.5 Conclusions of mechanical strain transfer analysis

This section presents several approaches by which one can estimate the ability of a sensor

to mechanically transfer spatially varying strains from a substrate to an optical fiber. Sec-

tion 4.2 details finite element methodologies by which both an embedded fiber and a surface

mounted fiber can be analyzed in terms of strain transfer behaviour. The embedded analysis

is straightforward, using cyclindrical elements. The surface mounted approach is novel, using

Reissner/Mindlin plate elements to model the tapered profile of an optical fiber &ed to

the surface of a substrate. The parameters contributing to the strain transfer behaviour are

considered. Geometry of the fiber/bond set, ratio of stifhesses clearly contribute. In the

limit of elastic behaviour, strain gradients are defined in terms only of their spatial extent:

not the strain excursion. Design charts are provided to make an evaluation of the strain

transfer response of either an embedded or surface mounted optical fiber sensor subject to a

linear gradient.

In section 4.3 a more general and analytical approach to the arbitrary strain transfer

problem is explored. For an embedded fiber, an analytical derivation predicts strain transfer

behaviour when the substrate strain field is expanded into its spatial harmonic components.

In this description, a transfer function determining the attenuation of each harmonic com-

ponent is derived as a function of the shear lag parameter N. The results of this technique

are compared against conventional finite element solutions of the same physical problem and

agreement is found to be strong. As well, the parameters argued in section 4.2 to determine

the strain transfer appear explicitly in the transfer function.

Finally, to extend the transfer function analysis to the surface mounted situation, an

equivalent N is estimated. The estimation is done by comparing the surface bonded sen-

sor's transfer function (evaluated by the finite element method) to the transfer function of

the embedded sensor and inferring an equivalent shear lag parameter. The latter analysis

provides a method by which the surface mounted sensor's ability to transfer a strain field

Page 118: Mechanics and Experiments

can be quickly evaluated.

Page 119: Mechanics and Experiments

Chapter 5

Demonstration of method and

measurements

Page 120: Mechanics and Experiments

This section presents optical measurements of B r a g gratings subject to a variety of load

conditions. From this optical data, the physical structure of the grating is reconstructed

using the theory of chapters 2 and 3. The strain transfer mismatch issues of chapter 4 are

considered in the comparison to the resolution of the sensor. The processing of the recovered

optical data requires some consideration which is also discussed in this section-

5.1 Review of methods for interrogating gratings with

non-uniform profiles

Several methodologies exist for recovering distributed information from a non-uniform grat-

ing. Two similar strategies-, phase spectmm based (PSB) and intensity spectrum based

(ISB) interrogation were developed in parallel at the University of Toronto Fiber Optic

Smart Structures (FOSS) group in the mid to late 1990's. The strategies are simila because

of their dependence on an estimate of penetration depth of a single wavelength into the

sensor grating. In both cases an explicit wavelength-position relationship is derived- These

relationships hold for the case of monotonically increasing or decreasing in-fiber strain fields.

5.1.1 Phase-based spectrum (PSB) strategy

The concept of local resonant wavelength is introduced,

Equation 5.1 is simply 2.46 expressed as a function of position. In other words: at position

z: wavelength X(z) is locally reflected. As well, equation 3.35 of chapter 3 is used to specify

the relationship between local wavelength shift and local axid strain

The PSB is derived by considering the phase accumulation of a wavelength, A, reflected at

z; over an interaction length dz7 figure 5.1. In terms of the wavelength of the propagating

wave, and the local index of refraction: the incremental phase accumulation is written as

Page 121: Mechanics and Experiments

Figure 5.1: Phase accumulation due to reflection at z of length dz.

The rate of change of expression 5.3 with wavelength is

Summing 5.4, d+P (A,) -- - 47mp (2) z

~ X P % (5.5)

is obtained. Equation 5.5 is re-arranged to solve for the axial position of the specific wave-

length A,; as a function of the grating's phase, @p,

Equation 5.6 indicates the position of each local wavelength, A,. Thus, using equation 5.2

the local Bragg wavelength shift can be used to find the local strain contribution at every

2. The si,dficance of expression 5.6 can be understood with an analogy to an interferom-

eter. Imagine two wavelengths, between which a large phase excursion is found. The large

phase excursion suggests a large physical distance between the penetration depths of the

two wavelengths. The converse analogy can also be imagined: two wavelengths with a small

Page 122: Mechanics and Experiments

phase excursion between them are reflected fiom physically proximate reflection centers. The

shorter the physical distance between subsequent wavelengths, the larger the spatial chirp,

thus the strain gradient. PSB measurements are reported in[15] and[47] - This type of ex-

periment requires a high coherence continuously tunable source for narrow band wavelength

discrimination. The local coupling strength of the grating is of no consequence provided that

there is significant reflectivity to recover the phase.

5.1.2 Intensity based spectrum (ISB) strategy

The reflected intensity approach, again, derives a specific one-to-one relationship between

physical space and local reflected wavelength. Matsuhara and Hill[57] study a monotonically

chirped grating. The authors describe the chirped grating as having a distribution of local

wavelengths as

where A, is the central wavelength of the spectrum. The aforementioned reference provides

a relationship between the reflectivity at the center wavelength as a function of the chirp

and grating strengths. In the context of sensing, the relationship between the reflectivity of

a certain wavelength, Ap is extended to the case where the gradient is not necessarily linear,

and the grating coupling is not necessarily constant,

R (Ap (2)) = 1 - exp X' (An (2))

Equation 5.8 can be rearranged, as the spatial rate of change of local B r a g wavelength is

sought,

Equation 5.9 can be integrated as it appears, or, if a grating of uniform coupling strength is

employed, An (I) = An,, an explicit expression for z can be found,

The last equation indicates the specific one-to-one relation between spatial position and local

wavelength. As can be seen born equation 5.9 the reflectivity contribution from a specific

Page 123: Mechanics and Experiments

physical location is dependent not only on the local chirp, but also the local coupling strength.

This method provides an advantage as there is no requirement that a high coherence tunable

source need be employed-a broadband light emitting diode is sficient-but an estimate of

the original grating strength distribution must be made to discriminate between the effects

of strong local coupling and weak local chirp. Published accounts of results of the ISB cite

spatial resolution of approximately 3mm[47] and[l5].

For the case of an arbitraq strain field, the single penetration depth approximation ceases

to hold. The physical reason for this may be seen by considering the following example: A

sensor subject to a field in which a strain field increases to some value then, with increasing

position, decreases mew to the original nominal strain value. In practice, this type of field

is common, as strain concentrations in the vicinity of holes or advancing crack tend to show

a high localized strain field surrounded by a nominal far field strain value on either side

of the imperfection. Thus, the same local wavelength exists at two distinct points in both

the increasing and decreasing branches of the strain field. The intuitive ascription of a

single physical location to a single reflected wavelength fails to describe the field. Which

location should be chosen to represent the reflected wavelength? As well, multiple reflection

centers for the same wavelength produce Fabry-Perot cavity interactions incompatible with

a one-to-one relationship between position and reflected wavelength.

To overcome this problem two solutions are proposed. The first suggests the continued

use of the ISB method, but with a p r e - w e d grating as the sensing device. The sensor

is designed such that the nominal chirp is sufficiently strong so that with the additional

imposition of the non-monotonic field due to structural perturbations of the substrate, the

resulting in-fiber field remains monotonically increasing or decreasing and the interrogation

methods for a monotonically strained sensor can be used. The substrate measurement is

recovered by subtracting the nominal pre-chq from the total in-fiber chirp- An improvement

in spatial resolution to lmm is reported using a pre-chirped grating[58].

The second strategy, demonstrated in the late nineteen nineties a t the University of

Toronto FOSS group, can be seen as the confluence of the individual phase and reflectivity

strategies- As derived in section 2.4, for low reflectivity cases, a Fourier relationship relates

the complex reflectivity (of which constituent parts are the reflectivity and optical phase)

Page 124: Mechanics and Experiments

ELED Delay line Bragg Grating

? stretcher

VCO I

[ counter , I

Figwe 5.2: Closed loop schematic of[l]

and the physical structure of the grating for the arbitrary strain field. Beginning with both

reflectivity and phase, the inverse transform technique uses two pieces of information in

one inverting algorithm to recover arbitraxy distributed in-fiber strain fields. The inverse

transform method is used in this work.

5.1.3 Low coherence interferometry .

Before outlining in detail the inverse transform experiment, an alternative for interrogating

non-uniform Bragg gratings is discussed. The University of Southampton's Opteoelectronic

Research Center has proposed a low coherence strategy fcr performing this measurement [l8] [17]

111 PSI The proposed approach uses a low coherence source to interrogate the Bragg grating in

two configurations. Both employ Michelson interferometers of which one arm leads to the

Page 125: Mechanics and Experiments

VCO

m 1 ELED

Optical Amplifier \ I

Lens coupling to Gree space mirror

/ Acquisition

Figure 5.3: Open loop schematic of[l]

Page 126: Mechanics and Experiments

grating under interrogation. The key to the low coherence approach is that interference

h g e s appear only when the interferometric path mismatch of the two Michelson arms

are balanced to within the sub-millimeter coherence length of t b e ELED source. Thus, by

displacement of the reference arm either by stretching the fiber or by displacing a bulk

mirror, a short resonant position is interrogated in the Bragg grating. All other positions

being outside the source coherence length-

In the closed loop configuration, a feedback loop tracks the local Bragg wavelength. The

feedback algorithm maximizes the power through an acoust~optic tunable filter (AOTF) by

continuously varying its transmission window. The reference arm is 50m long. A portion

is wrapped around-an oscillating PZT cylinder for purposes of generating several fringes

within the coherence length. Two posts 1.7m apart can be actuated relative to one another

to provide fiber extension. In the open loop configuration, position is scanned by displacing

an actuated hee space mirror coupling into the end of the reference arm. The open loop

configuration characterizes devices as functions of time of flight from the reference mirror.

The results of this technique are strong. Spatial resolution of 342pm in conjuction with

strain resolution of 5 . 4 p ~ / m is reported.

5.2 Implementation of interferometric phase recovery

experiment

O h et al. [16] first describe the interferometric Fourier transform experiment for distributed

strain sensing. A schematic of the experiment is illustrated in figure 5.4

The tunable source is a semiconductor distributed Bragg reflector (DBR) laser. The laser

includes three sections: a Bragg structure lies at one extremity of the lasing cavity, a passive

waveguide provides phase control and gain is achieved in an active section. Three input

currents are required to control the device--excluding thermal stabilization. A first injection

current controls lasing power as it is supplied to the active region. The other two currents are

used together to determine the lasing wavelength. With a goal of preserving single moded

operation a coarse tuning current flows through the Bragg section of the semiconductor. The

center Bragg wavelength, specified by the same Bragg condition seen in the context of fiber

Page 127: Mechanics and Experiments

Reflectivity Photodetector Bragg

Grating

Tunable DBR Source

Photodetector A

! Lock-in Ramp Signal / Amplifier

Fiewe 5.4: DBR Source interferometric experiment with lock-in phase.

Page 128: Mechanics and Experiments

devices, is determined by period and current (which influences E): X = BiA. The grating

is tuned by varying the carrier density though the Bragg section, which manipulates the

section's index. To maintain single h e behaviour, current is supplied to the phase section

to compensate for the change in optical path in the Bragg section. The DBR is able to tune

continuously over a 3 . 5 ~ range.

Section 2.4 details the derivation of the recovery of in-fiber strain from in-fiber parameters.

For the balance of the experiments, low birefringence fiber is used, thus the polarization

specifying p subscript is dropped from all nomenclature. It is required to construct the

complex electric-field proportional reflectivity r (A) as a function of the tuning wavelength.

It must be assembled from independent measurements of reflected intensity and phase. In

the first set of distributed experiments, the DBR is tuned somewhat slowly, taking nearly

two minutes to scan over the tuning range. A dither and lock in amplifier is used to better

extract phase measurements-

Results reported in reference[l6] cite a theoreticd strain resolution of approximately

25pm and a spatial resolution of 3mm. The latter reference does not include a comparison

with analytical elasticity solutions for purposes of verifying the strain and spatial resolutions.

It is at this point that the current thesis seeks to continue to contribute to the field, following

the developments of the ISB, PSB; and the initial reporting of the potential of the inverse

transform method.

5.2.1 DFB tunable source

The source employed in the bulk of this work is generously provided by Nortel Networks.

The specific fabrication of the device itself is proprietary to Nortel Networks and is not

known to this author. It is known that the device is a distributed feedback (DFB) laser.

A general overview of a DFB laser is given. This type of semiconductor has a single active

section. Inside the active section a phase shifted Bragg grating is impressed[59]. Viewed in

transmission, a notch exists in the center of the grating stop band-where typically reflectivity

is highest. The lasing cavity exists about the mid-point of the T-shifted grating. Figure 5.5

indicates that the phase shifted grating can be imagined as two juxtaposed gratings separated

by the phase shift. Imagine the wave fkom the center traveling to the grating to the right.

Page 129: Mechanics and Experiments

Lightwave resonates between the

i penetration depths of the x-shifted

Resonator formed by "two" gratings separated by a spatial mphase shift in DFB structure.

Figure 5.5: Phase shifted grating as DFB

The wave penetrates to a certain effective depth and is reflected. As it passes the phase shift,

the wave, coherent with the right side grating is maximally out of phase with the left side

grating. Directly after the phase shift the coupling between the wave and the LHS grating is

minimal due to the phase mismatch between them. The wave penetrates and is reflected and

repeats the cycle on the other side. Because the Bragg reflectivity is strongly wavelength

dependent, cavity modes that axe detuned from the Bragg condition are strongly suppressed

permitting stable single mode operation.

Tuning is achieved by altering the effective index of the substrate. The physical spacing

between the penetration depths and the index define the cavity modes. Cavity spacing

depends on this an optical path, ZL- As well, the Bragg reflections are wavelength dependent

with the peak reflectivity defined by the Bragg condition. This is an optical path-like term

EA. With injection current, the Bragg condition moves in parallel with the change in optical

path, preserving single line behaviour over a range of lasing wavelengths. Lasing power is

not independently controlled and some variation in output intensity over the tuning range

is expected.

The Nortel device in question is driven with a variable current source capable of ramping

bstween 30mA and 300mA in scans as fast as lms. Temperature control is accomplished

Page 130: Mechanics and Experiments

with a commercially available Pelletier thermoelectric cooler. The calibration of t h e device

is performed with a simple all fiber Michelson interferometer- The output intensity of a

Michelson interferometer as a function of wavelength is

in which Il and I2 are dependent on the splitting ratio of the coupler and 4 L is the mismatch

length between interferometer arms. As the wavelength of the laser is tuned a wawelength

dependent fringe pattern is received at a photodetector. The injection current is sirmultane-

ously monitored as the laser is tuned giving phase, as a function of drive current, @ I (2). The

total phase excursion can be expressed in tenns of the beginning and end wavelenmhs: XI

and X2 respectively, as

The group of variables 47rEA L can be solved in terms of the total phase excursion, pelrmitting

to write the calibration of lasing wavelength in terms of injection current, i,

X (i) = A@~otcrlxl A2 (5.13) 4 ( 2 ) (A1 - A2) + &'Tddx2 '

The tuning behaviour of the DFB is plotted in figure 5 -6. A continuous tuning range at nearly

lOnm is characterized. The end wavelengths are measured using an HP optical spectrum

analyzer (OSA) with a resolution of 80 pm. This error may lead to an absolute offset of

the same order. The relative accuracy of the calibration is much better than 80pm -because

any error scales over the lOnm tuning range. Figure 5.6 shows two experimental calibration

curves derived kom two differently mismatched interferameters and a best fit plot. T h e high

wavelength portion of the calibration obtained from the interferometer of several milllimeters

of mismatch deviates slightly horn smooth behaviour. This is corrected by lengtherning the

mismatch of the interferometer and increasing the fringe density at the detector. Higher

fringe density facilitates the fringe counting algorithni as the fringe density increases to

the point where it was no longer required to seek intra-hinge information. A 10-th degree

polynomial is used to fit the calibration of the long interferometer calibration in cmmputer

code used to process experimental data

Page 131: Mechanics and Experiments

Calibration of Nortel DFB Laser

. . . . . . . Long Interferometer --.- -.- Curve Fit Used for Code

0.0 0 5 1.0 15 2.0 2 5 3 -0

Ramp Signal [V]

Figure 5.6: DFB laser wavelength calibration

Page 132: Mechanics and Experiments

Tuneable DFB Laser (1536nm to 1545nm range) Mirror

(Cleaved end is suflicient) -

A* k -1 Reference Power

. Coupler

Phase Photodetector

Reflectivity Photodetector

Figure 5.7: Interferometric experiment

The interferometric experiment is shown in figure 5.7. Three New Focus transimpedance

photodetectors rated to a bandwidth of IMHz are used to capture the three optical signals.

Data acquisition is done with two PGbased DAC cards. The DAC cards, manufactured by

Gage Corporation, can track data samples at rates of up to 1 million readings per second.

Each DAC card has two analog inputs. Four sets of variables are measured over the course of

a single scan, a ramp voltage-directly related to injection current, the reflectivity response,

the phase response, and a reference power. The scan time is lms. This fast sweep minimizes

the vulnerability of the experiment to thermal crosstalk. The reference power is captured to

track the expected source power fluctuations as the source tunes. A trigger signal issued horn

the current supply synchronizes the data capture of the two boards. The code to control the

boards is written in C++. Figure 5.7 also shows qualitative drawings of the data recovered

horn each of the photodetectors.

Page 133: Mechanics and Experiments

5.2 -2 Representative grating interrogation

The goal of the data recovery is to f ist represent the wavelength domain information with

a complex electric field proportional reflectivity, r (A), recalling equation 2.54 of chapter 2 .

The magnitude of the electric field proportional reflectivity is found by taking the square root

of the reflected intensity, accounting for the reference power. The optical phase extraction is

more difiicult. An interference pattern is recovered a t the phase photodetector from which

the phase must be recovered. A formal phase expression for the intensity received at the

phase photodetector can be written as

I, (A) indicates the reflectivity of the silvered mirror termination of the Michelson arm.

I, (A) is a weak function of wavelength. RFBo (A) indicates the power proportional reflec-

tivity originally introduced in equation 2.43. RFBG (A) is a strong function of wavelength.

Equation 5.14 has little practical use. To invert and recover phase explicitly following equa-

tion 5.14, the reflectivity of the mirror would need to be calibrated and the reflectivity of

the grating would need to be divided out. Dividing out the reflectivity envelope proved

unprofitable as the small but fmite noise floor did not permit exact knowledge of the reflec-

tivity envelope. Attempts to self normalize the interferometric response resulted in many

numerical 'division-by-zero7 type problems.

As well, the fringe density is dependent on two factors: the strain content of the sensor

(high strain content increases the fringe density) and the optical path mismatch between the

interferometer arms (large optical path mismatch induces high hinge density). Empirically,

it is found that the experiment performs best when receiving a moderate Mnge density- The

phase accumulation of an interferometer is derived in equation 5.12. Replacing one mirrored

end of the interferometer with a Bragg grating will give a slightly merent phase response,

A - A, 27rc t$ (A) = 45~riaL- + - r (A) d i -

A20 A20

The first term indicates the Michelson phase excursion due to the interferometric mismatch

between the mirrored arm and a distance at the fioont of the Bragg grating. The second

term accounts for differential penetration depths into the grating. Recalling the comments

Page 134: Mechanics and Experiments

of section 2.3 relating delay to penetration depth, the grating arm can be considered a

mirrored arm: but with a wavelength dependent arm length. SpecXcally: the delay charac-

teristic defiaes the penetration depth for every wavelength. To manipulate the rate of phase

accumulation, AL can be lengthened or shortened.

As hinge density increases the wavelength resolution required to discriminate phase also

increases. Too high a fringe density can exceed the source's ability to resolve individual

fringes resulting in lost phase information. As well? too low a fringe density requires intra-

fringe phase interpolation. This too is undesirable because the envelope of reflectivity can

vary with the same order as the interferometric phase, making interpretation difEicult. In

other words, if interferometric phase varies too slowly, both the RFBG (A) term and the

cos ( 4 (A)) vary with similar rates making separation of their respective contributions to

I (A) problematic.

In practice, a numerical routine determines the phase from the int erferometric reflectivity-

but a check is performed on every data set to ensure that the recovered phase makes sense

when viewed against the reflected intensity. An example is shown in figure 5.8 where the

cosine of the inferred phase is plotted with the interferometric reflectivity.

The complex reflectivity is assembled fiom the phase and reflected power as

in which the computation of R (A) includes the effect of the varying source power over the

scan. Equation 2.63 relates the spatial distribution of the coupling strength to the complex

reflectivity as a function of wavelength,

2 = K (z) = --

A2 /--- r (A) exp

A uniform grating of length 8mm and peak reflectivity of near 0.6 is interrogated with the

interferometric system. For comparison, the grating behaviour is simulated by the T-Matrix

method of chapter 2.4. The coupling distribution is computed fiom both the experimentally

recovered reflectivity and the T-matrix simulated reflectivity. Knowing the grating length

and its coupling distribution allows accurate synthesis of the of the reflectivity. Figure 5.9

shows the experimental and simulated reflectivities in the upper plot. The lower shows three

Page 135: Mechanics and Experiments

6000 6500 7000 7500 8000

Sample Number

Figure 5.8: Cosine of inferred phase against interferometric reflectivity

Page 136: Mechanics and Experiments

-0.4 -0.2 0.0 0.2 0.4 Detuning (A-Lo) [nm]

=120 - - I - C 0 1 80 - 3 L Comprud fmm simulued reflectivity

r on which simvlalion is basal

Interferometric Mismatch [m]

Figure 5.9: Reflectivity and coupling distribution experiment and simulation comparison

Page 137: Mechanics and Experiments

curves of coupling distribution. The perfectly rectangular c w e is the coupling distribution

used to simulate the reflectivity shown in the upper plot. The two other plots are the results

of equation 5.17 using both the simulated and experiment a1 reflectivities as reflectivity inputs.

Both coupling distributions show a similm behaviour, climbing in magnitude from left to

right. The computed non-uniformity, predicted by simulation and seen by experiment, is

due to the violation of the low-reflectivity assumption used in deriving equation 5.17. The

grating has a peak reflectivity of 0.6 due to a coupling distribution/length product, KL = 0.8.

Figure 2.8 of section 2.4 explores the limitations of the low reflectivity simplification.

Notice as well, that the coupling distribution computed from the experimental data shows

a different magnitude from the coupling distribution computed from the simulation. The

simulated result also shows a ripple on the coupling distribution. The reason for these effects

lies in the data treatment. The integral of equation 5.17 needs to be discretized for numerical

computation. Naturally, the function r ( A ) is represented by a vector of complex numbers

r,: each at A,, and the integral is approximated by a discrete, finite sum:

in which the coupling distribution is also discretized as Kl at y. The I-th z is found in terms

of wavelength by, y

-Z[ = 277 (A* - XI) - (5.19)

In practice equation 5.18 is implemented as a discrete digital Fourier transform. The simu-

lated coupling distribution shown in figure 5.9 is computed using relationship 5.18.

Reconsider equation 5.18 in the following light: Equation 5.18 expands the coupling

distribution in terms of spatial wavelength components. Specifically, it may be rewritten as

where the new variable A,, represents the spatial period of the components used to construct

the coupling distribution. In terms of wavelength span, the m-th spatial period, A,,, is

Naturally, to achieve higher spatial resolution, more and more spatial periods need be in-

cluded in the sum of equation 5.20. In theory, one can specify an arbitrary spatial resolution

Page 138: Mechanics and Experiments

by extending the range of the wavelength scan. For example, if one desires a spatial resolution

of l p m the wavelength range required by the scan is f 8OOnm about the center wavelength

of the scan. More reasonably, a 100pm resolution requires a scan width of 8nm- The first

limitation to implementing such a strategy is the Limited bandwidth of source and detectors-

A second experimental limitation is more important: signal to noise ratio. Consider a scan

over ten nanometers. The spectrum of a the Bragg grating is concentrated in a narrow

bandwidth surrounded by a finite noise floor. As equation 5.18 is an integral over the scan

range, the longer the scan, the more background noise will be included in the sum. Thus as

the scan range is increased arbitrarily, an advantage in principle of more spatial resolution

is gained at the expense of including more background noise in the integral.

To moderate this effect, the technique of vtzndowzng is used. Windowing selectively

suppresses the reflectivity data in an attempt to reduce the contribution of background

noise, while enabling a longer scan range. The windowing envelope is superposed on the

reflectivity data, suppressing the end points of the scan to zero. A basic Hanning window is

used in this work. The actual computer coded implementation of equation 5.17 appears as

where the Hanning window, centered

written as

between the beginning and end points of the scan is

+ cos ( 2r ;X - 7r (A2 - XI) A2 - A1 )I -

In figure 5 -9, the experimental reflectivity is inverted using the data window. The success of

the window of reducing the ripple is seen. However, windowing changes the reflectivity data.

In particular, it preferentially suppresses the reflectivity in the extremities of the scan. This

reduction of the reflectivity induces an underreporting of the grating's coupling strength.

The consequences of experimental data processing determine the spatial resolution. For a

typical m~asurement as wide as possible a wavelength range is used. Performing the inversion

of equation 5.18 coupling distributions are recovered with significant high hequency noise.

The high frequency spatial noise is clearly because of the inclusion of a significant amount

of background noise in the wide wavelength range scan. The window function is applied to

the data to reduce the spatial oscillations. The tightening of the window explicitly specifies

Page 139: Mechanics and Experiments

Bragg grating, partially

Fibwe 5.10: Grating of which one half is bonded to an aluminum substrate

the spatial frequencies that are included in the reconstruction of the coupling distribution,

and as such, determines the spatial resolution. This is a critical observation: the resolution

of the sensor is determined by the lowest period spatial wavelength that can be included in

the construction of the coaplzng distribution.

5 -3 Interferometry-based phase recovery distributed sens-

ing strategy

The following section describes some experimental results obtained with the interferometric

phase recovery distributed sensing experiment.

5.3.1 Uneven thermal apparent loading: experimental demonstra-

tion of spatial resolution

Figure 5.10 shows a schematic of an experiment designed to test the spatial resolution of

the interferometric distributed sensing technique. An 8mm long uniform Bragg grating is

Page 140: Mechanics and Experiments

bonded to an aluminum substrate. The grating is bonded such that approximately one half

of the length of the grating is affixed to the aluminum, while the remaining portion hangs

free, beyond the aluminum substrate. An initial measurement is taken at room temperature

after allowing a sufficient cure time for the epoxy. Then, the specimen is heated b i d e

a laboratory oven. The temperature is increased to 39OC above ambient conditions. The

specimen is held at this temperature for two hours at which point it is removed from the

oven. After cooling, it is interrogated again.

It is expected that the portion of the grating bonded to the alllminum substrate will

behave ditferently than the fkee portion of the grating. The response of either side of the

grating has been derived in chapter 3. Equation 3.39 describes the response of the free portion

of the fiber; a sensitivity of 9.66pm/OC, is predicted. The bonded side sensitivity, taking the

thermal expansion coefficient of the T2024 aluminum as 22.5p~/OC[60]~ is computed with

equation 3.45. The expected bonded response is 35.6pm/OC. The apparent strain is found

as the ratio of thermal and strain sensitivities

For the free side of the grating, an apparent strain sensitivity of 8.2p~/OC is expected. On the

bonded side, a higher sensitivity is expected, 30.1p~/OC, as the thermal expansion of the alu-

minum dominates the response. Figure 5.11 depicts the reflectivity of the partially adhered

grating as it is cycled thermally. As the grating is heated to 14OC above room temperature,

the power distribution of the reflectivity shows two distinct features, one centered roughly

0 . 2 m above the nominal wavelength, the other 0.5nm above A,. These distinct features

correspond to the two halves of the grating. The higher wavelength (higher apparent strain)

portion of the reflectivity corresponds to the bonded portion of the grating with its larger

thermal sensitivity. Figure 5.12 illustrates the apparent strains recovered from the experi-

ment. The apparent strain for the specimen heated to 14OC above room temperature shows

the spatial distribution across the grating. Two distinct regions of different strain are seen

as well, again corresponding to the two halves of the sensor. The grating is then held at 39°C

above room temperature where it undergoes a post cure. Initially, there is relative motion

between the fiber and the epoxy as the epoxy softens. After some time-approximately two

hours-the epoxy re-cures at the higher temperature and the optical fiber is fixed relative

Page 141: Mechanics and Experiments

Figure 5.11: Measured reflectivities fkom sensor partially bonded to an aluminum substrate

subject to thermal load.

Page 142: Mechanics and Experiments

-

-0.003 -0.002 -0.001 0.000 0.001 0.002 0.003

Position [m]

Figure 5.12: Apparent strains reported from sensor bonded on dllminum substrate subject

to thermal load

Page 143: Mechanics and Experiments

to the aluminum at this new position- When the specimen is permitted to return to room

temperature the apparent strain on the bonded side of the fiber is seen to have dropped

appreciably with the thermal contraction of the alllminum. The sensitivity of 35.6pm/*C

predicts an apparent strain drop of 1180y~. This decrease, illustrated in figure 5.12, is very

close to the recorded strain drop of the bonded side of the grating. The reflectivity plot

for the post cured specimen shows the group of wavelengths that, when heated appear on

the long wavelength side of the spectrum, have fallen to the short wavelength side of the

spectrum.

The scans are processed as described in section 5.2.2. Initial inversions of the data reveal

the expected high spatial frequency oscillations in the recovered coupling distributions. The

window function is applied to the data to suppress off band background noise. The results

presented in figure 5.12 are windowed such that spatial periods more narrow than 200pm are

excluded from the inversion. For the group of the first three scans, this level of windowing

is acceptable. The final scan reveals some oscillations in the region of the high spatial

gradient. This result indicates that the spatial resolution of the sensor is being approached.

As the inversion attempts to recreate the narrow spatial feature, more spatial h-equencies

are required. The experimenter may be tempted to include more bandwidth in the inversion

to further increase the spatial resolution, however a point of diminishing returns has been

reached. The experiment is not capable of extracting further real information from the

scan. Increasing the scan range will simply include more of the off band noise floor, not

real reflectivity information from which the resolution of the coupling distribution can be

improved. For the final example, the window is tightened to restrict spatial features to

less than 500pm and the high kequency oscillation is suppressed at the expense of spatial

resolution.

This series of results proves that it is possible to resolve in-fiber spatial features with

high resolution, in this example 500pm- As well: a variation of 6 6 0 p ~ is recovered over the

space of l2mm-a gradient of 550p~lmrn

Page 144: Mechanics and Experiments

Figure 5.13: Grating bonded to an alllminllm plate in the vicinity of a circular hole

5.3.2 Grating bonded in the vicinity of a circular hole

A longer grating of 50mm is. affixed to an allrmin~~ll substrate into which a circular hole of

radius 5.4m.m is drilled, figure 5.13. The fiber lies 7.7- from the center of the hole. Axial

tension is applied using a Tinius-Olsen tensile testing machine. The specimen is strained to

several different loads. The circular hole is a desirable test because it admits an analytical

solution with which experimental results can compared[61]. The longer grating has a peak

undeflected reflectivity of 17% from which the coupling strength-length product can be found

as EL = 0.43, which is well within the low reflectivity assumption. The expected reflectivity

for each of the load cases is simulated with the T-matrix formulation.

Individual reflectivities are plotted for the four loadings in figure 5.14. The experimental

reflectivities are consistent with simulation. Some salient features of the spectra are noted.

Common to all load cases, the bulk of the reflectivity is concentrated at the lower wavelength

portion of the spectrum. This is consistent with the shape of the strain profile in the

neighbourhood of a circular hole where a significant portion of the physical length of the

grating is close to the far-field strain level. Oscillations in the reflectivity are seen in all

cases. A trend appears as the oscillations become less rapid at higher wavelengths. This

feature is explained by a Fabry-Perot like reflection inside the strained grating. Lower

wavelength regions exist at two points in the grating, on either side of the hole-perhaps a

few centimeters-and a resonance between these two physical points manifests itself as the

rapid oscillations in the low wavelength portion of the spectrum. The Fabry-Perot cavity

for the higher wavelength group, is physically shorter in length-perhaps sub centimeter-

Page 145: Mechanics and Experiments

Figure 5.14: Experimental and simulated reflectivities for different loadings.

Page 146: Mechanics and Experiments

Table 5 -1: Summary of measurements at different far field strains

Far Field Tension

e:

k 4 172 534 818 1115 ,

located on either side of the high strain peak.

Using the results from section 4.4, with N = 0.22 it is found that the strain field trans-

ferred into the fiber will be replicated inside the fiber with less than 1 percent attenuation.

Thus, the performance of the measurement is limited only by sensor performance.

The strain results are plotted in figure 5.15 in the upper plot, with the deviation from

Maximum Measured

er" (4 b 1 267 804 1256 1695 ,

elasticity theory with distance along the grating is presented in the figure's lower plot. The

results are compared in terms of the root mean square error as a function of axial position,

% Maximum Error

eFPI ey0' (z1

27 8 7 I1 ,

Maximum Error

eF" (4 [PEI 71 64 90 189

defined as I

Table 5.1 summarizes the results of the measurements. The maximum error over the mea-

RMS Errar

ez RMS

[ P I 9 11 17 3 3 ,

surement in absolute terms increases with the magnitude of the strain content. At high far

% RMS Error

eF- efLMS(x)

3 2

< 2 < 2 .

field strains: the strain content of the region near the hole increases. In turn, the wavelength

content of the reflectivity is increased and a subsequent signal to noise reduction occurs as a

similar total optical power is spread across a larger bandwidth. As well, the phase becomes

increasingly rapidly varying with increasing strain content.

For a single loading repeated measurements are performed to gain an appreciation for

the reproducibility of the experiment. Several scans are made, the specimen is removed horn

the holding jaws of the tensioning apparatus, remounted, and reloaded to the same nominal

tension. Figure 5.16 shows the results of both mountings. Measurements are compared with

elasticity solutions in the left hand side plots and the deviation &om theory is presented in the

right hand side plots. The nature of the error is similar between measurements at the same

Page 147: Mechanics and Experiments

Figure 5.15: Measured and analytical strain profiles for grating bonded near circular hole

for different axial loads (upper plot). Deviation between theory and experiment (lower plot)

Page 148: Mechanics and Experiments

Table 5.2: Summ;trv of repeated measurements

Far Field Tension

e [wI 765

(First Mounting)

765 (Second Mounting)

Mounting)

loading conditions, suggesting rceasonable repeatability of the measurement system. The

first mounting results consistently over-report strain values on the left side of the maximum,

while the right side is systematiocdy under-reported. This feature is less prominent in the

second mounting attempt. The second mounting more closely reports the theoretical strain

field. Table 5.2 summarizes b o t h the absolute and RVIS errors of the measurements.

The long grating is less successful than its shorter counterpart. Strong windowing is

required to eliminate erroneous ~ g h spatial frequency oscillations in the measurements. No

spatial frequencies are included wi th spatial periods shorter than 2mm. .Attempts to extend

the window to larger scan ranges resulted increased spatial noise and no improvement in

spatial resolution. The reason f a r the variation in sensor performance in spatial resolution

is discussed in section 5.5. The stteepest gradient reported by the longer sensor is 58pelmm.

M b u m Measured

eZ"" (4 [ ~ . 4

1353

1271

5.4 Delay-based

%Y

5.4.1 Measurement

phase recovery distributed sensing strat-

Maximum Error

eF"- (4 [wI 141

91

Experimentally, a drawback of t h e interferometric technique is the requirement that a Michel-

son interferometric path must be included in the optical arrangement to recover phase infor-

% M a w Error r \z) eFor

10

7

WIS Error

RMS e z

[PEI

40

25

% R M S Error

efMdS (2)

eF-

3

2

Page 149: Mechanics and Experiments

0 0 0 0 m o m

Figure 5.16: Repeated measurements of two different mountings of the same load (left) and

deviations from elasticity solution (right).

Page 150: Mechanics and Experiments

Tunable Source Modulator Grating

Figure 5.17: Phase shift modulation method

mation from an interferornetric kinge pattern. The phase of the reflectivity must be found by

a hinge counting algorithm. From measurement to measurement, the fringe density changes

depending on factors such as the initial bandwidth of the grating, the strain content in the

sensor, and the mirror/grating optical path mismatch. That the fringe density is measure-

ment dependent limits the potential of the interferometric strategy to be implemented as

a field deployable system. An arbitrarily high fringe density requires an arbitrarily high

wavelength resolution of the scanning source to ensure hinge discrimination.

This section proposes a method by which the interferometric arm can be eliminated horn

the experiment: the group delay characteristic of the grating sensor is measured directly.

A proof of concept test of this new strategy involves a more simple optical arrangement

that can be implemented for distributed sensing.

The elimination of the interferometric arm is realized by recognizing the relationship

between the delay of a wavelength component and its phase. The relationship between delay

Page 151: Mechanics and Experiments

Tunable Electro-Optic cpoWff Meter Bragg Source Isolator MZ Modulator Coupler Grating

Coupler Signal Amplifier Photodetector

Figure 5.18: Delay-based phase recovery experiment

and optical phase is given as[62] d4 (N

7 (A) = --. a!#

Equation 5.26 is recast as

Equation 5.27 provides the phase for the construction of the function r (A). Measurement

of the group delay is performed by the phase shzft modulation method first proposed by

Rhu[63]. Figure 5.17 helps to explain the phase shift modulation method principle. A

tunable source is modulated at an RF frequency, whose period determines the available

range of continuously measurable group delay. The modulation also affixes an indication of

time of flight to the wave. The time of flight of the wave is continuously compared to a

fixed phase reference (such as the modulator drive signal) . The reflection &om the grating

carries with it the modulation. As certain wavelengths suffer different penetration depths

(i-e. different delays) the relative phase between the fixed reference and the reflected wave

varies. The phase difference is determined by a high speed network analyzer. The phase

difference is translated into time delay through knowledge of the modulation frequency.

An alternative derivation can be performed in the frequency domain. Writing the mod-

ulated signal as the sum of two optical frequencies separated by the modulation frequency,

f2 - fl = fmod., the reflection horn the grating is properly seen as the superposition of reflec-

tions originating from X (fi ) and X ( f2). The phase of the superposition of the reflected waves

Page 152: Mechanics and Experiments

is a function of the differential delay between the two closely spaced wavelengths. The mod-

ulation frequency is selected such that the separation of the two probe wavelengths matches

the step resolution of the reflection measurement: effectively matching the resolution of the

delay and amplitude measurements.

The optical arrangement is shown in figure 5.18- An Anritsu tunable source is externally

RF modulated by a JDS-Uniphase Mach-Zehnder modulator. The Bragg grating sensor is

then interrogated with the moddated wave. A low bandwidth power meter detects reflected

optical power, R (A). The modulated signal reflected fiom the Bragg sensor is compared to a

reference signal by an EfP network analyzer to acquire the group delay as the laser is tuned.

The reference signal is obtained electrically. an optical reference, a t the expense of a second

high speed photo-detector, will provide a measure of thermal compensation if the optical

reference arm is proximate to the sensing arm.

The RF network analyzer and the Tinius-Olsen tension testing machine were not simul-

taneously available, thus an alternative loading apparatus is devised based on beam bending,

in lieu of pure tensioo, figure 5-19. Two aluminum beam samples are prepared to demon-

strate the measurement. Into each specimen a circular hole is drilled a smaller hole in one

specimen of diameter 5mm, and a larger hole in the other: with diameter 10m.m. An in-fiber

Bragg grating sensor, of 50mm length, is bonded to each beam, parallel to the long axis of

the sample, at a distance of lmm fiom the hole perimeter. A frequency of 400MHz is used to

modulate the interrogation lightwave. The available differential phase resolution was &0.l0.

This corresponds to a group delay resolution of *0.7ps.

To assess the strain resolution: the strain field was theoretically estimated. A combination

of two simple models is proposed. The elasticity solution in the neighbourhood of a circular

hole[4l] due to a uni-axial tension is linearly scaled to account for the bending beam. In the

absence of aoy hole, the upper surface of the beam on which the sensor is mounted, would be

in linearly varying tension as dictated by conventional beam theory. The strain prediction

is made by scaling the solution of the hole in a plate under uniform tension, a t each point,

by a ratio of the local strain of the equivalent uniform beam (without a hole) to the beam

tensile strain value at the sensor midpoint.

The specimen is loaded by suspending a mass from the end of the beam. The reflected

Page 153: Mechanics and Experiments

Bragg grating senso bonded near hole.

Figure 5.19: Beam loaded in bending

Page 154: Mechanics and Experiments

Figure 5.20: Reflectivity and delay of unloaded and loaded samples.

Page 155: Mechanics and Experiments

optical power and the group delay are plotted in figure 5.20.

The analytical estimate and the experimental measurements are plotted in fiewe 5.21

for the two samples. The evaluation of the strain error is performed by computing the root

mean square over the range of measurement, yielding an error of f 2 0 p ~ for the smaller hole

and & 2 4 p ~ for the larger hole. Normalized with the maximum strain excursion, the RkfS

error is, for both loadings, 7%. Spatial resolution is again limited by the maximum amount

of scan range that can be included in the coupiing distribution integral. The window giving

the best results eliminated spatial frequencies having periods less than 1 -65mm.

The demonstration of the proposed method etiminates the undesirable interferometric

mirrored arm necessary in previous interferometric distributed sensing systems.

5.5 Conclusions of measurements

In this section experiments are undertaken to evaluate the performance of fiber Bragg grating

as a distributed sensor and to attempt to understand factors relating to its performance.

The ability of the sensor to resolve spatial feat-ures-the spatial resolution-is of primary

importance.

Section 5.2.2 shows that the coupling distriburtion computed &om experimental data is in

fact a Fourier sum of spatial frequencies. High spatial resolution depends on the inclusion of

high spatial fkequencies in this summation. More. high spatial frequencies in the z domain

are related to reflectivities at large wavelength excursions horn the center of the scan. Thus

to achieve high spatial resolution, it is required to be able to recover accurate information

over a large bandwidth in reflectiviv

The degree to which a grating offers recoverable reflectivity over a large bandwidth de-

termines the potential spatial resolution. Fibwe 5.22 shows two gratings of equal peak

reflectivi~ 30%; undeflected then subject to a localized strain perturbation. The strain

perturbation is a lmm Gaussian shape with a peak strain of l o o p . One grating is modeled

as having a 5cm length and the other icm. To achieve similar peak reflectivity the lcm

grating requires an index modulation, 672, five times stronger than the 5cm, 5.87 x

versus 29.3 x respectively. The long grating has a much more narrow bandwidth than

Page 156: Mechanics and Experiments

, # I * * , , . , Ncar Large Holc (Theordical) Ncar Small Holc (Expcrimcnlal) ..-.---

0.00 Position [m]

Page 157: Mechanics and Experiments

- lcm Grating Undefl ected --- With localized permrbarion

-0.2 -0.1 0.0 0.1 0.2

Wavefength Offset fnm]

-0.2 -0.1 0.0 0.1 0.2

Wavelength Offset [nm]

0.4

A

6 0.3 er: h 4 -- 1 0.2 - U 0

C 4 0-1

Fi,we 5.22: Comparison of two gratings lcm and 5cm of like peak reflectivity subject to

Like localized strain perturbation

-

5cm Grating - Undeflected --- With Iocalized perturbation

Z I \ 1 1 I t I I I I I \

0.0 -

Page 158: Mechanics and Experiments

the short sensor. When deflected: both gratings behave in a similar manner qualitatively:

the central shape shifts to a higher wavelength and a side lobe grows on the low wavelength

side of the spectrum. The longer, low bandwidth grating is much more difficult to interre

gate. While scanning with a tunable source of fixed step size, the longer, low bandwidth

grating offers fewer points of reflectivity available for interrogation. For example, with a

bandwidth of approximately lOOpm the short grating provides an order of magnitude more

points available for interrogation. Recalling that each point in wavelength space represents

a spatial frequency in the z space, it is clear that more spatial resolution i s available using

the short grating.

For example, using equation 5.21, a wavelength offset of O . l n m corresponds to the spatial

period of 7.5mm. The magnitude of the reflectivity at this position for the short grating is

appro-dmately 1.5%. The corresponding reflectivity of the longer grating is 0.065%. The

reflectivity of the longer grating, at this position, is more than twenty times smaller than

the analogous reflectivity of the short grating. Thus, the spatial kequency of the coupling

distribution represented by this wavelength position is twenty times more difficult to extract

for the case of the narrowband grating. Should the noise floor of the experiment exceed the

reflectivity of the sensor, the information relating to that spatial frequency i s lost. Including

this erroneous information in the inverse transform will lead to a high spatial frequency noise

on the recovered coupling distribution. Windowing is required to suppress these hequencies.

The limit where the recoverable reflectivity falls below the noise floor determines the spatial

resolution.

Experiments support this observation. Section 5.3.1 presents the results of the deploy-

ment of a short high bandwidth sensor. Spatial resolution is very strong, between 200pm and

500pm. The spatial resolution of the sensor determines its ability to extract gradients. With

the high spatial resolution, the recovery of a gradient of more than 500p~fmm is demon-

strated. A short sensor can be deployed for extremely high spatial resolution measurements.

The experiments of section 5.3.2 deploy long, low nominal bandwidth gratings. The ad-

vantage of a using a longer sensor is compromised by the reduction in spatial resolution.

Spatial resolution is limited by the sensor's narrow bandwidth. Only harmonic components

of spatial periods exceeding one or two millimeters are relevantly extracted from the reflec-

Page 159: Mechanics and Experiments

tivities. A maximum gradient of 6 0 p ~ / m m is recovered in this configuration. Percentage

errors in recovered strain profiles are presented. Typical strain errors are of the order 10%

when one considers the maximum excursion between theory and measurement. Taken as an

m4S value, the errors are typically of the order of 5% to 10%.

A second, novel, experimental approach is reported in section 5.4. The interferometer

arm is replaced with a delay based strategy for phase recovery. The performance of this

approach is equivalent to the interferometric approach and offers a more direct method by

which to extract the grating phase.

In s u m m a r y 7 the following conclusion are drawn from the experiments of this work.

0 Spatial resolution is determined by the shortest spatial wavelength used in the Fourier

series representation of the coupling distribution.

Wider scan ranges in wavelength enable the recovery of better spatial resolution.

Practical limits due to the finite noise floor limit the degree to which the scan range can

be arbitrarily expanded. The technique of windowing the reflectivity data is required

to moderate the effect of the finite noise floor.

Spatial resolutions of sub millimeter are achieved with a lcm long, high nominal band-

width sensor.

Spatial resolutions the order of millimeters are achieved with a 5cm long, narrow

nominal bandwidth sensor.

The interferometric arm of the Michelson can be replaced with a delay based measure-

ment approach with no performance penalty

The bandwidth of the sensor is the most important factor in determining the spatial

resolution of the sensor.

Page 160: Mechanics and Experiments

Chapter 6

Conclusions and contribution

summary

Page 161: Mechanics and Experiments

6.1 Summary of contributions and conclusions

The content of this thesis is intended to contribute to the fiber optic sensing community.

in particular, in the area of Bragg grating distributed sensing. The content of chapter 2 is

largely background derivations necessary to describe the optical response of a grating. Tools

employed heavily in later chapters such as the T-Matrix formulation to simulate a grating's

optical response, and the approximate Fourier inversion technique for reconstructing the

coupling distribution based on optical data are reviewed.

6.1.1 Contributions in understanding the optical response

-4 responsible investigation of the Bragg grating's deployment as a distributed sensing device

mandates the investigation of the sensor's response to thermo-mechanical variables. The

phase response is detailed for the distributed Bragg sensor for strain and temperature in

chapter 3. Experimental evaluations of the strain optic coefficients, the fiber stiffness, the

Bragg grating thermal and strain sensitivities are performed.

6.1.2 Contributions towards understanding the strain transfer prob-

lem

Significant work is done to illustrate the relationship between the in-fiber strain field and the

strain field in the host or substrate material in the presence of a spatially varying field in the

host. Chapter 4 details this work. The initial study considers the transfer of a linear strain

gradient from host or substrate material to fiber. This is the &st reporting of the problem

of relating the directly measured in-fiber strain to the desired host or surrounding material

strain in the context of distributed sensing. Results for both the embedded and the surface

mounted case are presented. The finite element modeling of the surface mounted optical

fiber, using appropriate stackings of Reisner/hlindlin plate elements is presented as novel.

A parameter study of the factors affecting axial strain transfer is undertaken. A first author

journal article in Smart Materials and Stmctures[53] supports the latter contributions.

The investigation into the strain transfer problem is further pursued. A novel analytical

derivation is presented, proposing a methodology to relate the in-fiber strain to the host

Page 162: Mechanics and Experiments

surrounding material strain field for the case of an arbitrarily varying strain field. A unified

description of the strain transfer problem is outlined, with emphasis on spatial frequency

content of the strain field and the transfer of individual components- Agreement is found

between the proposed theory and finite element analysis. A second, first author article in

Smart Materials and Structures[64] details this derivation. As well, a new strategy for simple

modeling of the three dimensional case of a surface mounted sensor is proposed through the

use of the equivalent geometric ratio.

6.1.3 Contributions in the area of distributed strain measure-

ment s -

Chapter 5 reports distributed Bragg grating measurements. This portion of the thesis con-

tinues in the same direction established by the work of colleagues at the FOSS laboratory.

At the inception of- this thesis, the potential for distributed Bragg measurements was known,

but the resolution of the technique was not strictly defined and presentation of definitive

measurements was not complete. In this work, e-xperiments are undertaken to carefully eval-

uate the performance of the sensor. Comparisons with analytical solutions are employed to

report the experimental error.

The relationships between sensor bandwidth and spatial resolution are defined. Sub

millimeter spatial features are recovered using a short length broad band grating. Longer

g a t ings are used to demonstrate measurements. The best strain resolution obtained with

the longer sensors shows an mC3 error of 25 microstrain over a measurement with a strain

content of more than 1000 microstrain. These results are included in the proceedings of the

1999 CLEO rneeting[49].

Spatial resolution is discussed in terms of grating bandwidth. The data treatment tech-

nique of windowing is proposed and implemented to improve the measurement. The param-

eters determining strain spatial resolution are clearly specified.

A contribution is made in terms of the distributed sensing experiment design. A novel

sensing scheme based on the recovery of group delay in lieu of interferometric phase is

proposed. Similar resolution of twenty or so microstrain is demonstrated. This contribution

is supported by a first author Optics Letters publication[50].

Page 163: Mechanics and Experiments

6.2 Suggested future work

There are several departure points for future study implicit in the content of this thesis.

The Fourier transform inverse scattering method and its limitations are outlined in chap

ter 2. The principal limitation of the FT is the low reflectivity assumption included in its

derivation. Poladian [65] and Feced et. al. [66] have recently proposed inversion algorithms

that are not restricted to low reflectivity devices. An attempt to apply improved inverse

scattering algorithms to distributed Bragg grating sensing is suggested.

The question of strain transfer quality to the optical fiber sensor from the mechanical

host holds several unexplored areas. The approach studied in this document is restricted to

the case of a perfectly adhered sensor, performing in the elastic regime. A comprehensive

evaluation of the strain trahsfer problem: including debonding, and non-linear effects is

relevant to the real world deployment of distributed Bragg sensors-

-Also useful, would be a fully three dimensional finite element model of the surface

mounted sensor. First, such an analysis would be expected to improve the estimate of

equivalent transfer performance of the surface mounted sensor. As well, in the telecommuni-

cations industry, the bi-axial strain induced by the surface mounting of a fiber is of interest.

Extension of the surface mounted analysis with a view to suppressing unwanted induced

birefringence would prove a valuable study.

Several experimental improvements require further assessment. A lock-in technique may

be attempted to better extract relevant sensor bandwidth from the noise floor--effectively

increasing the spatial resolution of the sensor. The results of such an experiment might be

useful in specifying a strict condition by which to apply the data window. Miniaturization

and cost reduction of the RF circuitry required to perform the delay measurement are key

to distributed Bragg grating sensing's widespread viability. A further increase in the scan

speed of a single measurement wodd make dynamic distributed measurements possible.

The results of this document argue that larger bandwidth sensors provide better spatial

resolution. A series of spa t idy cascaded shorter sensors of different wavelengths can be

imagined as a first step to optimizing the length of the sensor, while preserving sub-millimeter

spatial resolution. Such a progression may lead to the use of a chirped grating of moderate

strength. In general, the apodization of the sensor; in conjuction with the improved inverse

Page 164: Mechanics and Experiments

scattering methods, merits further study to optimize distributed sensing performance.

Page 165: Mechanics and Experiments

Bibliography

[ l j M. Volanthen, H. Geiger, and J. P. Dakin, "Distributed grating sensors using low coher-

ence reflectometry." Journal of Lightwave Technology, vol. 15, pp. 2076-2082: November

1997.

[2] E. Udd, ed., Fiber Optic Sensors: An Introduction for Engineers and Scientists. Toronto:

John Wfiey & Sons, 1991.

[3] R. J. V. Steenkiste and G. S. Springer, Strain and Temperature Measurements with

Fiber Optic Sensors. Basel, Switzerland: Technomic Publishing, 1997.

[4] C. D. Butter and G. B. Hocker, "Fiber optics strain gauge," Applied Optics, vol. 17:

pp. 2867-2869, Septemeber 1978.

[5] G. P. Agrawal: Fiber-optic Communications systems. Toronto: John Wiley & Sons,

1997.

[6] R. M. Measures, "Fiber optic sensing for composites smart structures:" Composites

Engineering, vol. 3, no. 7-8, pp. 715-750, 1993.

[7] R. M. Measures, S. M. Melle, and K. Lui. "Wavelength demodulated Bragg grating fiber

optic sensing systems for addressing smart structural critical issues," Smart Structures

and Materials, vol. 1; pp. 36-44, 1992.

8 S. M. Melle, K. Liu, and R. M. Measures, "A passive wavelength demodulation system

for guided wave bragg grating sensors," IEEE Photonics Technology Letters, vol. 4,

pp. 516-518, 1992.

Page 166: Mechanics and Experiments

[9] S. M. Melle, K. Liu, and R. M. Measures, 'Practical fiber-optic Bragg grating strain

gauge," Applied Optics, vol. 32, 19%.

[lo] S. Huang, M. LeBlanc, M. M. Ohn, and R. M. Measures, "Bragg intragrating structural

sensing," Applied Optics: vol. 34, pp. 5003-5009, August 1995.

[I11 J. S . Sirkis, "Unified approach to phase-strain-temperature models for smart structure

interferometric optical fiber sensors: part 1, development," Optical Engineering: vol. 32.

pp. 752-761, April 1993.

[12] J. S. Sirkis, "Unified approach to phase-strain-temperature models for smart structure

interferometric optical fiber sensors: part 2, applications," optical engineering, vol. 32,

pp- 762-773, April 1993.

[13] M. LeBlanc and R. M. Measures, "Micromechanical considerations for embedded single-

ended sensors," in Smart Sensing, Processing and Iwtmmentation, vol. 1918 of SPIE

Proc-, pp. 215-227, SPIE, Bellingham, WA: 1993.

[14] M. LeBlanc, S. Huang, M. Ohn, R. Measures, A. Guemes, and A. Othonos, "Distributed

strain measurement based on a fiber grating and its reflection spectrum analysis," Optics

Letters, vol. 21. no. 17: pp. 1405-1407, 1996.

[Is] S. Huang, M. Ohn, and R b1. Measures: "Phase based Bragg intragrating distributed

strain sensor," -4pplied Optics, vol. 35: pp. 1135-1142, 1996.

1161 M. M. Ohn, S. Y. Huang, R. M. Measures, and J. Chwang, "Arbitrary strain profile

measurement within fibre gratings using interferometric Fourier transform technique,"

Electronics Letters, vol. 33: pp. 1242-1243, July 1997.

[17] M. Volanthen, H. Geiger, M. J. Cole: and J. P. Dakin, "Measurement of arbitrary strain

profiles within fibre gratings ," Electronics Letters, vol. 32, pp. 1028-1029, May 1996.

[18] M. Volanthen, H. Geiger, M. J. Cole, R. I. Lamming, and J. P. Dakin, "Low coherence

technique to characterise reflectivity and time delay as a function of wavelength within

a long fibre grating," Electronics Letters, vol. 32, pp. 757-758, April 1996.

Page 167: Mechanics and Experiments

[19] K. 0. Hill, Y. Fujii, D. D. Johnson, and B. S. Kawasaki, "Photosensitivity in optical

fiber wave,edes: application to reflection filter fabrication," Applied Physics Letters,

vol. 32, pp. 647-649, May 1978.

[20] G. Meltz, W. W. Morey, and W. H. Glenn, "Formation of Bragg gratings in optical fibers

by a transverse holographic method," Optics Letters, vol. 14, pp. 823-825, August 1989.

[r?l] D. 2. Anderson, V. Mizrahi, T. Erdogan, and A. E. White, "Production of in-fibre

g a t ings using a difiactive optical element ," Electronics Letters, vol. 29 pp. 566-568,

March 1993.

!22] B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics. New York: John Wiley

and Sons, 1991.

[23] D. Gloge, "Weakly ,guiding fibres," Applied Optics, vol. 10: no. 10, pp. 2252-2258, 1971.

[24] A. Yariv and M. Nakamura, "Periodic structures in integrated optics," IEEE Journal

of Quantum Electronics, vol. QG13, pp. 233-253, -4pril 1977.

[25] T. 0 koshi, Optical Fibers. Tokyo: Academic Press, Harcourt-Brace- Jovanovich, 1982.

[26! J. E. Sipe and G. I. Stegeman, "Comparison of normal mode and total field analy-

sis techniques in planar integrated optics," Journal of the Optical Society of America:

vol. 69, pp. 1676-1683, December 1979.

[27] R. Kashyap, Fiber Bragg Gratings. Toronto: Academic Press, 1999.

[28] M. Yamada and K. Sakuda, "Analysis of almost-periodic distributed feedback slab

waveaOUide via a fundamental matrix approach," Applied Optics, vol. 26, pp. 3474-3478,

1987.

1291 H. Kogelnik, "Filter response of nonuniform almost periodic structures," The Bell Sys-

tem Technical Journal, vol. 55, pp. 109-126: Jmuary 1976.

[30] U. Eriksson: P. Blixt; and J. A. Tellefsen, Jr., "Design of fiber gratings for total disper-

sion compensation," Optics Letters, vol. 12, pp. 1728-1738, October 1994.

Page 168: Mechanics and Experiments

[31] W. W. Morey, G. Meltz, and W. H. Glenn, 'Fiber optic Bragg grating sensors," in

SPIE: pp. 98-10?; 1989.

[32j J . S. Sirkis, "Optical and mechanical isotropies in embedded fiber optic sensors:" Smart

Materials and Structures, vol. 2, pp, 255-259, August 1993.

[33] K . 4 Kim, L. KollL, and G. S. Springer, "A model of embedded fiber optic Fabry-

Perot temperature and strain sensors," Journal of Composite Materials, vol. 27, no. 17,

pp. 1618-1662, 1993-

[34 S. G. Lekhnitskii, Theory of Elasticity of an Anisotropic Elastic Body. San Francisco:

Holden-Dax 1963.

[35] A. Bertolds and R. Dandliker, "Determination of the individual strain-optic coefficients

in single-mode optical fibers," Journal of Lightwave Technology, vol. 6, pp. 17-20? Jan-

uary 1988.

[36] 3. F. Borelli and R. A. %filler, "Determination of the individual strain-optic coefficients

of glass by an ultrasonic technique,: Applzed Optics, vol. 7, pp. 745-750, May 1968.

[37] A. J. Barlow and D. N. Paqne, "The stress-optic effect in optical fibers," IEEE Journal

of Quantum Electronics, vol. QE-19, pp. 834-839: May 1983.

[3S] Y. Namihira, "Opto-elastic constant in single mode optical fibers," Journal of Lightwave

Technology, vol. LT-3, pp. 1078-1083, May 1985.

[39] A. Bertholds and D. Dlindliker, "Deformation of single-mode optical fibers under static

longitudinal stress," Journal of Lightwave Technology, vol. LT-5, pp. 895-900, July 1987.

(401 J. S. Sirkis; Y. L. Lo, and P. L. Neilsen, "Phase-strain model for polarirnetric strain

sensors based on fictitious residual strains," Journal of' Intelligent Material Systems and

Structures, vol. 5, pp. 494-500, July 1994.

[41] E. E. Sechler, Elasticity in Engineering. New York: Dover, 1952.

Page 169: Mechanics and Experiments

[42] K.-S. Kim, Y. Ismail, and G. S. Springer, "Measurement of strain and temperature with

embedded intrisic Fabry-Perot optical fiber sensors," Journal of Composite Materials:

vol. 27, no. 17: pp. 1663-1677, 1993-

[43] Corning Corporation, s ~ F - 2 8 ~ ~ Fiber Specifications www.corningfiber.com.

[ H. L. Cox, "The elasticity and strength of paper and other fibrous materials," British

JournaZ of Applied Physics, vol- 3, no. 72, pp. 72-79, 1952.

[45] &I. LeBlanc, Interaction mechanics of em bedded, single-ended optical fibre sensors using

novel in-situ measurement techniques. P hD thesis, University of Toronto, 1999.

[46j M. R. Piggott, Load Bearing Fiber Ccomposites. Toronto, Canada: Pergamon, 1980.

[47] S. Huang, M. LeBlanc, M. M. Ohn, and R. M. !vfeasures, 'Continuous and arbitrary

strain profile measurements with fiber B r a g gratings," Smart Materials and Structures:

vol. 7, pp . 248-256, 1998.

[48] bf. Volanthen, H. Geiger; M. J. Cole: and J. P. Dakin, "Distributed grating sensor for

real time monitoring of arbitrart strain fields," in Advanced Sensing Photonics, vol. 11

of OFS-Post Deadline, pp. 6-9: 1996.

[49] G. Duck and R. M. Measures, "High resolution and high speed distributed in-fiber Bragg

grating strain measurements," in CLEO '99: 1999 Technical Digest Series, pp. 295-296;

1999.

[50] G. Duck and M. M. Ohn, "Distributed Bragg grating sensing with a direct groupdelay

measurement technique," Optics Letters, vol. 25, pp. 90-93, January 2000.

[51] Ciba-Geigy Corporation, Ciba-Geigy Corporation Adhesive Documentation.

[52] Dexter Aerospace Materials Division, Hysol Structu~al Adhesives.

[53] G. I. Duck and G. Renaud, 'The mechanical load transfer into a distributed optical

fiber sensor due to a linear strain gradient: embedded and surface bonded cases," Smart

Materiats and Structures: vol. 8, pp. 175-181, 1999.

Page 170: Mechanics and Experiments

[54] R. M. Jones, Composite Maten'als. Taylor & Francis: 1975.

[55] G. Renaud and J. S. Hansen, "A quasi-three-dimensional for cracked structures repaired

with composite patches," Canadian Aeronautics and Space Journal, vole 45, pp. 90-98,

June 1999.

[56] G. Renaud, "Composite patch repair including thermal processing effects: a finite ele-

ment analysis capability," Master's thesis, University of Toronto, 1997.

[57] M. Matsuhara and K. 0. Hill, "Optical-waveguide band-rejection filters: design," Ap-

plied Optics, vol. 13, pp. 2886-2888, December 1974.

[58] S. Huang, M. Leblanc, M. Lowery, R. Maaskant, and R. M. Measures: "Distributed fiber

otpic strain sensing for anchorages and other applications," in Proceedzngs of the 2nd

International Conference on Advanced Composite Materials in Bridges and Structures,

pp. 991-998, 1996.

[59] G. P. Agrawal and S. Radic, "Phase-shifted fiber bragg gratings and their applications

for wavelength demultiplexing," IEEE Photonics Technology Letters, vol. 6 , pp. 995-997,

August 1994.

[60] W. F. Riley and L. Zachary, introduction to Mechanics of Materials. New York: John

Wiley & Sons, 1989.

1611 S. Timoshenko, Strength of Materials. New York: Van Nostrand Inc., 1940.

[62] F. Ouellette, "Dispersion cancellation using linearly chirped Bragg grating filters in

optical waveguides," Optics Letters, vol. 12, pp. 847-849, October 1987.

[63] S. Ryu. Y. Horiuchi, and K. Mochizuki, "Novel chromatic dispersion measurement

met hod over continuous gigahertz tuning range,? Journal of Lightwave Technology,

vol. 7, pp. 1177-1180, August 1989.

[64] G. Duck and M. LeBlanc, "Arbitrary strain transfer from a host to an embedded fiber-

optic sensor," Smart Materials and Structures, vol. 9; no. 4, pp. 492-497, 2000.

Page 171: Mechanics and Experiments

[65] L. Poladian, "Simple grating synthesis algorithm," Optics Letters, vol. 25, pp. 787-789,

June 2000.

[66] R. Feced, M. N. Zenms, and M. A. Muriel, "An efficient inverse scattering algorithmfor

the design of nonuniform fiber b r a g gratings," IEEE Journal of Quantum Electronics7

vol. 35, pp. 1105-1112, 1999.