slow light in fiber bragg gratings for sensing … › file › druid:gd710fc6670 › ... · i...

124
SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING THERMAL PHASE NOISE, ATTOSTRAINS AND OTHER APPLICATIONS A DISSERTATION SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Georgios Skolianos March 2016

Upload: others

Post on 07-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING

THERMAL PHASE NOISE, ATTOSTRAINS

AND OTHER APPLICATIONS

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF

ELECTRICAL ENGINEERING

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Georgios Skolianos

March 2016

Page 2: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/gd710fc6670

© 2016 by Georgios Skolianos. All Rights Reserved.

Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-Noncommercial 3.0 United States License.

ii

Page 3: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Michel Digonnet, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Olav Solgaard

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Jelena Vuckovic

Approved for the Stanford University Committee on Graduate Studies.

Patricia J. Gumport, Vice Provost for Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file inUniversity Archives.

iii

Page 4: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  iv  

Abstract

Strain sensors have many applications in structural health monitoring, civil engineering,

geoscience, and gravitational wave detection. To address the needs of high-end applications,

and to push the resolution of fiber strain sensors well beyond their current picostrain/√Hz

level, we have explored the potential of utilizing the narrow slow-light resonances that exist in

strong fiber Bragg gratings (FBGs) and the noise limitations of this kind of sensor.

Prior to this work the strain resolution of slow-light FBG strain sensors were limited by the

frequency noise of the laser used to interrogate them. The main goals of this thesis were first

to reduce this source of noise by utilizing a probe laser with a much narrower linewidth, which

necessitated the design of new gratings matched to the new laser; and second, to study

theoretically and experimentally the next noise source below the laser frequency noise, which

is thermodynamic phase noise.

This work was broadly divided into four major tasks. First, we studied theoretically, from

basic thermodynamic principles, the magnitude and frequency dependence of the thermal

phase-noise in Fabry-Perot-like resonances similar to the slow-light resonances available in

strong FBGs. This study showed that the thermal phase noise is proportional to the group

index of the resonance, and that when expressed in units of strain it is proportional to 1/√L,

where L is the length of the sensor. Thus the shorter the FBG the higher the thermal phase

noise in units of strain. Second, we improved the design and the fabrication of our FBGs in

order to achieve very high group delays, and hence very high sensitivities in short fibers. This

step was crucial because to measure thermal phase noise in units of strain (normalized output

power noise to input power times the sensitivity), a short FBG with high sensitivity is needed

for the sensor to be limited by thermal phase noise, otherwise it would be limited from other

noise source, i.e. laser intensity noise. We achieved this result by using strongly apodized

FBGs written with a femtosecond laser in deuterium-loaded fiber, and thermally annealing the

FBG optimum sensitivity was achieved. Using this technique, we were able to achieve a 42-ns

group delay, an eight-fold improvement compared to what was reported previously in similar

FBGs. Third, we had to modify our experimental setup to improve its stability and reduce the

Page 5: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  v  

dominant laser frequency noise of the sensor. To this end we probed our sensors with a new

low-noise laser from OrbitsLightwave with a 200-Hz linewidth, we placed our sensor in an

anechoic enclosure, and we used a low-noise photo-detector. The last effort was to use these

various developments and improvements to design, fabricate, and test two FBGs, one to

measure the thermal phase noise in an FBG for first time, and the other to observe the smallest

strain ever measured in an FBG-based sensor (an minimum detectable strain (MDS) of

110 fε/√Hz at 2 kHz and 30 fε/√Hz at 30 kHz). This sensor was so stable that it exhibited no

drift in its Allan variance after a four-day measurement. By integrating a 4-day output trace

with an 8-hour integration time, we were able to measure an absolute MDS of 250 attostrains,

the lowest value ever measured in an FBG.

While we were aiming to measure thermal phase noise and reduce the MDS, some other

applications presented themselves. Because these devices confine light not only temporally

but also spatially, they can be used in applications that benefit from extremely high intensities

and confinement, in particular in quantum electrodynamic experiments and nonlinear optic

applications such as optical signal processing. As a proof of concept, I report in the end of this

thesis the performance of two FBGs optimized for maximum field enhancement, maximum

Purcell factor, maximum group delay, and minimum group velocity (a record of 300 km/s).

The measured values for these parameters are the highest reported in an all-fiber device. These

properties enable robust novel devices that are simple to fabricate and in which light can be

coupled easily and efficiently.

 

 

Page 6: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  vi  

Acknowledgement  

First of all I would like to thank Professor Michel Digonnet for his help and support

throughout all the years I spent at Stanford University and for letting me join his research

group from the moment when I arrived at Stanford University. From this first moment, until

the very end of writing this thesis he was always available for me when I needed his help and

guidance. In our interactions in our meetings, which were often several per week, he patiently

showed me how to thoroughly think about a problem from every angle, how to solve it, how

to pay attention to the details and then step back and explain my work to someone else without

taking for granted that this person would know what I already knew. I am really thankful for

all his efforts, without his support in every aspect, my Ph.D. journey would not have been

possible.

I would also like to thank Professor Jelena Vuckovic and Professor Olav Solgaard for being

on my reading committee. They were the instructors in my first courses at Stanford, and both

my interactions with them and the knowledge I gained through their courses helped me with

my research and enabled me to realize other possible applications of my research. Also I

would like to thank Professor Nick Bambos for being the chair of my oral defense committee

and Amir Safavi-Naeini for being on my oral committee.

I would like to thank Professor Martin Bernier at University Laval for the close interaction we

had, and for fabricating all the fiber Bragg gratings I studied in this thesis. Without his help

we wouldn’t be able to realize our gratings and make this research a reality.

I would also like to thank my professors at the Aristotle University of Thessaloniki in Greece,

where I did my undergraduate studies, and especially my undergraduate advisor Professor

Emmanouil Kriezis who introduced me to optical fibers and provided me with the right skills

and knowledge to attend Stanford University.

I would like to specially thank He Wen, who started this work of slow-light in fiber Bragg

gratings and guided me in my first steps in this work. I would also to thank Arushi Arora for

her help in the last years of my Ph.D. journey, and especially for taking over some time-

Page 7: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  vii  

consuming but very important aspects of this work. Her help sped up my progress a lot. Also I

would like to thank the rest of our research group (Josh, Kiarash, Olive, Jacob, Therice).

I would like to thank Northrop Grumman for their financial support for my research and the

interactions we had in our biannual meetings. I also thank Ingrid Tarien for her administrative

help.

I would like to thank all my friends and the Hellenic community at Stanford that made the

years here a very enjoyable experience.

Last but not least I would like to specially thank my parents and my brother for all their

support, not only during my time at Stanford University but throughout all my studies and my

life.

Page 8: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  viii  

 Table of contents

Abstract iv

Acknowledgement vi

Table of contents viii

List of Figures x

Chapter 1: Introduction 1 1.1 Background and motivation 1 1.2 Fiber Bragg gratings as a strain sensor 1 1.3 Evolution of FBG strain sensing and their limitations 2 1.4 Summary of this thesis work 9 References 11

Chapter 2: Basic theory of strain sensing using a slow-light FBG 13 2.1 Fiber Bragg Grating 13 2.2 Strain sensing using FBGs 17

2.2.1 General principle of strain sensing using an FBG 17 2.2.2 Using a sharp resonance to increase the strain sensitivity 20 2.2.3 Noise sources in an ultra-high sensitive strain sensor 21

2.3 Temperature sensitivity of slow-light FBGs 23 2.4 Slow-light resonances in an FBG 24

2.4.1 Slow-light resonances in a Fabry-Perot interferometer 25 2.4.2 Forming an FP and highly sensitive slow-light resonances 27

References 30

Chapter 3: Modeling the thermal phase noise in a passive Fabry-Perot resonator 32 3.1 Modeling phase noise in a Fabry-Perot interferometer 33

3.1.1 Background 33 3.1.2 Intuitive picture 34 3.1.3 Phase noise in the transmitted signal 35 3.1.4 Phase noise on resonance 41 3.1.5 Phase noise off resonance 45 3.1.6 Phase noise in the reflected signal 50

Page 9: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  ix  

3.1.7 Applicability to FBGs and FBG-based FPs 52 3.1.8 Phase noise in experimental fiber FP sensors 53

3.2 Conclusions 54 References 56

Chapter 4: Improving the performance of slow-light FBGs 59 4.1 Apodized FBGs 59 4.2 Realization of improved FBGs 64

4.2.1 Fabrication of slow-light FBGs with femtosecond lasers in deuterium-loaded fibers

65 4.2.2 FBG annealing 69

4.3 Modeling the index profile of FBGs written with a femtosecond laser 71 4.4 Conclusions 76 References 77

Chapter 5: Measuring the intrinsic thermal phase noise and 250 attostrains using slow-

light FBGs 79 5.1 Experimental Setup 80 5.2 Measuring the strain sensitivity and MDS of FBG sensors 83 5.3 Measuring thermal phase-noise in a 5-mm FBG 85 5.4 Measuring an absolute strain of 250 attostrains 88 5.5 Conclusions 92 References 93

Chapter 6: Other applications of slow-light FBGs 94 6.1 Optimizing slow-light FBGs for high Purcell factor and intensity enhancement 96 6.2 Fabrication and characterization of two FBGs optimized for high Purcell factor and

intensity enhancement 98 6.3 Conclusions 102 References 102

Chapter 7: Conclusions and future work. 105

Appendix: Power spectral density on resonance 109    

 

Page 10: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  x  

List of Figures

Figure 1.1. a) Generic transmission spectrum of a uniform FBG. b) Shift of the transmission

spectrum of an FBG due to an applied strain. In solid line is the initial transmission and

in dashed line the spectrum after a strain has been applied to the FBG. Probing the edge

of the FBG bandgap with a laser results in a large transmission change when strain is

applied on the FBG. ........................................................................................................... 2

Figure 1.2. Evolution of the MDS of selected strain sensors using FBGs reported over the

years. .................................................................................................................................. 4

Figure 1.3. Conversion of frequency noise into power noise through a device with a steep

transmission slope. The resulting power noise is proportional to the laser frequency noise

and to the sensitivity (i.e., the slope). ................................................................................. 5

Figure 1.4. Noise dependence at the output on detected power at 23 kHz [8]. .......................... 6

Figure 2.1. Periodically varying refractive index in a uniform FBG. ..................................... 14

Figure 2.2. Simulated power (a) reflection and (b) transmission spectrum for a uniform FBG

of length L = 2 mm, index modulation Δnac = Δndc = 5 x 10-4, period Λ= 534.3 nm, and

loss γ = 0.1 m-1. ................................................................................................................. 16

Figure 2.3. Shift of the Bragg wavelength and the whole spectrum due to an applied strain.

The initial transmission is indicated with the solid line while the spectrum after a strain

has been applied to the FBG is shown with the dashed line. ........................................... 18

Figure 2.4. Using a sharp peak for strain sensing. ................................................................... 20

Figure 2.5. Schematic of a generic Fabry-Perot interferometer ............................................... 25

Figure 2.6. Calculated reflection spectrum of a representative π -shifted grating (the grating’s

index profile is shown schematically in the inset). .......................................................... 28

Figure 2.7. a) For wavelengths near the band edge of a strong FBG, multiple reflections from

both ends of the FBG cause recirculation of light inside the FBG. The situation is

analogous to a Fabry-Perot (FP) interferometer. b) Similar to a FP, sharp resonances are

formed in the transmission spectrum. .............................................................................. 29

Figure 3.1. Schematic of the computation of the phase noise in the signal transmitted by a

Fabry-Perot interferometer ............................................................................................... 36

Figure 3.2. Phase-noise PSD of the signal on resonance transmitted by a 1-cm fiber Fabry-

Perot interferometer (see text for details). ....................................................................... 43

Page 11: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  xi  

Figure 3.3. Phase-noise PSD spectrum at the transmission port of a silica-fiber FP with

reflectivities r1 = r2 = √0.99, normalized to the single-pass PSD spectrum. Solid curves

are the numerical solutions for different detuning δΦ0 from resonance. ......................... 47

Figure 3.4. PSD spectrum on resonance (dashed curve) and off resonance for δΦ0 = 0.05

(solid curve) illustrating the splitting that occurs in the phase-noise resonances when the

optical signal is detuned from a Fabry-Perot resonance. ................................................. 49

Figure 3.5. PSD spectra in reflection (black curves) for the same FP as in Fig. 3.3, calculated

for two values of the detuning from resonance. The PSD spectra in transmission for the

same two detunings (solid red curves, reproduced from Fig. 3.3) are also shown for

comparison. ...................................................................................................................... 51

Figure 3.6. Theoretical predictions of the phase noise dependence on group index for three

highly sensitive strain sensors utilizing FBGs, and experimental noise measured for each

of them. ............................................................................................................................. 54

Figure 4.1. Creating strong slow-light resonances in an apodized FBG. a) ac and dc index-

modulation profiles of an apodized fiber Bragg grating. b) Dependence of the Bragg

wavelength on position along the grating. c) Effective mirrors in the apodized FBG

forming equivalent FPs. d) The transmission slow-light resonances formed as a result of

these multiple equivalent FPs. .......................................................................................... 60

Figure 4.2. a) Index profile of a uniform FBG (period Λ not to scale) and b) simulated

transmission and group index spectra of this FBG. ......................................................... 63

Figure 4.3. a) Index profile of a Gaussian-apodized FBG (period Λ not to scale), and b)

simulated transmission and group index spectra of this FBG. ......................................... 64

Figure 4.4. Power loss coefficients versus ac index modulation for different writing

techniques. ........................................................................................................................ 65

Figure 4.5. Setup and exposure conditions used to write the deuterium-loaded fiber Bragg

gratings using a femtosecond laser. (Courtesy Martin Bernier). ...................................... 67

Figure 4.6. Measured evolution with annealing temperature of (a) the transmission, (b) the

group delay, and (c) their product, for the slow-light resonances of a particular FBG. ... 71

Figure 4.7. Index modulation profile example of a fiber Bragg grating written with a

femtosecond laser; (a) convoluted profile, half-peak-to-peak ac index profile, and dc

index profile; (b) upper and lower envelopes of the index modulation. .......................... 72

Page 12: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

  xii  

Figure 4.8. a) Measured and simulated transmission spectra of the FBG. b) Magnified portion

of slow-light peak region of the transmission spectra shown in a). c) Measured and

simulated group-delay spectra in the same wavelength range as in b). ........................... 73

Figure 4.9. Experimental setup used to measure the group delay and transmission of the FBG

(see text for details). ......................................................................................................... 74

Figure 5.1. Experimental setup used to characterize the noise spectra and the sensitivity of

FBG strain sensors. The PZT plate excited by the function generator induces a known

strain on the FBG to calibrate its sensitivity. The lock-in amplifier measures the sensor’s

response ............................................................................................................................ 80

Figure 5.2. Noise contributions in a 5-mm FBG slow-light sensor. ........................................ 85

Figure 5.3. Transmission spectrum of the FBG used for measuring thermal phase noise,

measured after fabrication (before any annealing took place). ........................................ 86

Figure 5.4. Evolution of Τ0τg/c versus annealing temperature for the FBG that was used to

measure phase noise. ........................................................................................................ 87

Figure 5.5. Measured and calculated noise spectrum contributions in a 20-mm slow-light FBG

strain sensor in units of strain. .......................................................................................... 89

Figure 5.6. Time trace of the 20-mm sensor’s response at 30 kHz. ......................................... 90

Figure 5.7. Generic Allan deviation curve. .............................................................................. 91

Figure 5.8. Allan deviation in units of strain calculated from the data of Fig. 5.6. ................. 91

Figure 6.1. Simulated dependence on the laser beam width of (a) the Purcell factor and

intensity enhancement, and (b) the transmission of the best slow-light resonance of a

saturated FBG, evaluated at the peak of the slowest resonance. Solid curves simulate an

FBG with the parameters of [11], and dashed curves the experimental FBGs. ............... 98

Figure 6.2. Measured and fitted (a) transmission and (b) group-delay spectra of the first FBG.

Inset shows its inferred index-modulation profiles. ......................................................... 99

Figure 6.3. Calculated distribution of the intensity distribution along the FBG for the three

lowest modes. ................................................................................................................. 100

Figure 7.1 Evolution of the MDS of selected strain sensors using FBGs reported over the

years. .............................................................................................................................. 106  

Page 13: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

1  

Chapter 1: Introduction

1.1 Background and motivation

Sensing strain (ε = ΔL/L), namely a strain-induced change in length ΔL relative to some initial

length L, is useful in several applications, including structural health monitoring in aerospace

[1] and civil engineering [2], monitoring volcanoes and sensing earthquakes in the geosciences

[3], and gravitational wave sensing [4]. The requirement for the minimum detectable strain

(MDS), or equivalently the strain resolution, is different for different applications. In civil

engineering, an MDS of a few µε (10-6) is required, while in aerospace and in geosciences an

MDS of a few pε (10-12) up to a few nε (10-9) is necessary. For higher end applications such as

gravitational-waves detection, the resolution requirement is considerably lower, around 10-22 ε.

Optical fibers, especially fiber Bragg gratings (FBGs), have been used for decades as strain

sensors because they can be very sensitive, small, and they can be used in harsh environments,

they are immune to electromagnetic interference, and they operate equally well at low and

high temperatures, among other advantages common to most fiber sensors. In order to achieve

very low MDS, in the sub-femtostrain/√Hz range, very long interferometric optical sensors, on

the order of a few kilometers, are used [4], which is obviously exceedingly costly and of no

practical use for most applications. Thus it is important to investigate methods to reach this

range of strain resolution in a small and simple sensor. Such practical ultra-sensitive sensors

can enable new applications and fundamental observations that were not attainable before.

1.2 Fiber Bragg gratings as a strain sensor

A conventional FBG is a short length of single-mode fiber with a spatially periodic

modulation of the refractive index in the core region. At each of these index modulations

multiple Fresnel reflections take place. Over a limited wavelength range, centered around

what is known as the Bragg wavelength, most of these partial reflections are in phase, and

they add essentially constructively to create a strong reflected signal. Thus an FBG acts as a

reflector, and little to no light is transmitted (see Fig. 1.1a). This wavelength range is called

the bandgap. Outside the bandgap, the partial reflections are out of phase, leading to

Page 14: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

2  

essentially no reflection: the incident light is fully transmitted. But at some specific

wavelengths close to the bandgap some of the partial reflections happen to be in phase,

leading to a partial reflection and a dip in the transmission. If the wavelength is slightly

detuned from these weaker resonances, the partial reflections are abruptly out of phase, and all

the light is transmitted. This phenomenon leads to the ripples shown in Fig. 1.1a. Large index

modulations and long FBGs, which cause stronger and more Fresnel reflections, respectively,

lead to higher reflections in this wavelength range. Also, a longer FBG exhibits a narrower

bandgap because the reflections can only be in phase over a smaller range of wavelengths.

When a strain is applied to an FBG, the whole spectrum shifts (see Fig. 1.1b). By measuring

this wavelength shift, the applied strain can be inferred. An easy way to measure this shift is

with an optical spectrum analyzer (OSA). But because of the limited resolution of even high-

end OSAs (~0.01 nm), this approach produces strain resolutions only down to a few

microstrains/√Hz. To reduce the MDS, significantly smaller wavelength shifts must be

detected. This can become possible by increasing the sensitivity of the sensor and by reducing

the noise in the detected signal, which was one of the objectives of this thesis work.

Figure 1.1. a) Generic transmission spectrum of a uniform FBG. b) Shift of the transmission spectrum

of an FBG due to an applied strain. In solid line is the initial transmission and in dashed line the

spectrum after a strain has been applied to the FBG. Probing the edge of the FBG bandgap with a laser

results in a large transmission change when strain is applied on the FBG.

1.3 Evolution of FBG strain sensing and their limitations

In the early 1990s Kersey et al. used an imbalanced Mach-Zehnder interferometer (MZI) to

sense the strain-induced shift in the bandgap if an FBG [5] with a far greater precision that can

Page 15: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

3  

be done even today with a commercial OSA. The FBG was interrogated with a broadband

light source, and only the portion of the spectrum of the light that overlaps with the FBG’s

narrow bandgap was reflected. The reflected light was sent through an MZI with an arm

imbalance. When a strain is applied to the FBG, the bandgap shifts and so does the center

wavelength of the reflected light spectrum. The MZI converts this wavelength change into a

power change at its output port, which is measured with a power meter. This power change is

proportional to the wavelength shift and to the length mismatch between the two arms of the

MZI. Thus a smaller wavelength shift can be detected by using an MZI with a large length

mismatch between its two arms. But by increasing the length mismatch, two deleterious

effects take place. First, the thermal stability of the MZI deteriorates. Second, the linewidth of

the reflected signal entering the MZI is approximately equal to the bandwidth of the FBG

bandgap. If the length mismatch is larger than the coherence length of the reflected signal

(which is inversely proportion to this linewidth), the light signals at the ends of the two arms

of the MZI are no longer mutually coherent and no longer interfere at the output of the MZI,

and the latter no longer functions as a wavemeter. This limitation can be pushed back by

reducing the linewidth of the FBG (i.e., increasing the coherence length of the reflected

signal). However, this goes hand in hand with extremely long FBGs that have limited use as

sensors, or with a reduction in reflected power: this improvement can only be taken so far

before the reflected signal is so weak that the signal-to-noise ratio (SNR) at the detector

increases, resulting in a reduced MDS. By using an MZI with a 10-mm length mismatch and

an FBG with a ~0.4-nm reflection linewidth, Kersey et al. achieved an MDS of ~600 pε/√Hz

[5], which was a record at that time. This data point is shown in Fig. 1.1, which illustrates the

evolution of the resolution of selected high-end FBG strain sensors reported over the past 25

years.

In 1998 Lissak et al. used the edge of the bandgap to make a significantly more sensitive FBG

sensor [6]. At the edge of the bandgap, the spectrum has a steep slope, i.e., the transmission

goes from ~100% to ~0% over a small change of wavelength (see Fig. 1.1b). Thus, if the

interrogating laser wavelength is selected to fall at the point with the steepest slope, as the

bandgap shifts relative to the (fixed) laser wavelength in response to a small applied strain, the

output power changes by an amount proportional to the strain (see Fig. 1.1b). With a steep

slope, even a slight strain-induced shift will produce a very large change in the output power.

Using this technique, the steeper the slope the larger the output change for a given wavelength

Page 16: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

4  

shift (and equivalently a given strain). The ultimate limitation of this general scheme is noise

sources that affect the relative position of the spectrum and the frequency of the interrogating

laser, i.e., laser frequency noise, thermal phase noise, and environmental noise: these noise

sources are proportional to the slope of the transmission, and for a steep slope they become

dominant. In this case MDS is limited by the noise in the relative position of the spectrum and

the frequency of the interrogating laser which is independent of the slope and the spectrum.

Thus increase of the slope does not affect the MDS. This technique led to a higher sensitivity

and decreased the MDS down to ~50 pε/√Hz [6], one order of magnitude smaller than

achieved by Kersey et al. in [5].

Figure 1.2. Evolution of the MDS of selected strain sensors using FBGs reported over the years.

In 2008 Gatti [7] used the sharp transmission resonance that exists in the reflection spectrum

of a π-shifted FBG to further increase the slope of the spectrum and hence the strain

sensitivity. The sensing principle is exactly the same as in [6]: the steep slope of this

resonance converts a small strain-induced wavelength shift into a large power change. The

narrower this resonance is, the steeper its slope is, or equivalently the lower the group velocity

of light traveling at the frequency of this resonance. A higher sensitivity then requires a

resonance with a narrow linewidth, or equivalently a low group velocity vg, or equivalently

still a high group index, which is by definition ng = c/vg, where c is the velocity of light in

vacuum. A higher sensitivity also requires a high resonance transmission (since obviously a

resonance with near zero transmission, even if extremely narrow, would transmit nearly no

light, leading to no detected signal and hence zero sensitivity). In [7] the MDS was limited by

the frequency noise of the laser. The latter scales like the laser linewidth. This linewidth

introduces an uncertainty in the laser wavelength. When the laser is transmitted through a

Page 17: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

5  

steep slope, this uncertainty is converted into an uncertainty in the output power, namely a

power noise. Thus the power noise due to laser frequency noise is proportional to the

sensitivity (see Fig. 1.3). The laser used in [7] had a linewidth of 2 MHz. The fact that Gatti et

al. used a sharp resonance to increase the sensitivity led to a reduction of the MDS down to

5 pε/√Hz [7], another order of magnitude compared to Lissak’s measured MDS in [6]. So in a

period of about 15 years, the MDS of FBG strain sensors was reduced by roughly two orders

of magnitude, to the few pε/√Hz level.

In 2011 He Wen, a former graduate student of our research group, was first to introduce a new

principle to greatly improve on these results [8]. When the index modulation of an FBG is

high (high reflectivity) and the propagation loss is low, the FBG does not behave like a simple

reflector but like a Fabry-Perot (FP) interferometer, and it exhibits sharp resonances either just

outside or inside the bandgap, as described in relation to Fig. 1.1a. Although the existence of

these resonances was known [9,10], He Wen was first to design FBGs with very sharp

resonances (very steep slopes) and to exploit them to demonstrate very low strain resolutions.

This method produced resonances much sharper than any reported until then in FBGs, and

therefore FBGs with unprecedented group indices (hence the term of slow-light resonances)

and strain sensors with record sensitivities. By interrogating such an FBG with a laser with a

much lower frequency noise than used by Gatti et al. (~8-kHz linewidth [8] compared to

2 MHz in [7]) He Wen was able to measure a record MDS of 280 fε/√Hz at 23 kHz [8], more

than one order of magnitude lower than the previous record in a single FBG [7].

Figure 1.3. Conversion of frequency noise into power noise through a device with a steep transmission

slope. The resulting power noise is proportional to the laser frequency noise and to the sensitivity (i.e.,

the slope).

Page 18: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

6  

Figure 1.4. Noise dependence at the output on detected power at 23 kHz [8].

In Fig. 1.4 the noise dependence on the output power of Wen’s FBG sensor at 23 kHz is

shown [8]. The filled blue circles indicate the experimental points and the curves indicate the

predicted or measured main noise sources present in the detected signal. The solid black curve

is the total predicted noise, calculated by adding geometrically all the noise sources. At low

output power the noise is limited by the photodetector thermal noise. This noise source

depends solely on the photodetector and is completely independent of the incident light power

and of the sensor. At high power the noise is limited by the laser frequency noise. As already

discussed, the laser frequency noise depends on (increases with) the laser linewidth and it is

proportional to the sensor’s sensitivity. All the other noise sources shown in Fig. 1.4, namely

laser intensity noise, electrical shot noise, and optical shot noise, are negligible, because the

high sensitivity reduces their impact relative to the sources of noise that depend on sensitivity.

The laser intensity noise depends only on the laser type and on its operating power, whereas

the electrical shot noise and optical shot noise are due to the quantum nature of the light and

(for a given detector) depend only on detected power. If the sensor were shot-noise limited, by

increasing the operating power the MDS could be reduced, since the noise would increase

with increasing power (as the square root of power) slower than the sensitivity (proportional to

power, as plotted in green in Fig. 1.4). Unfortunately, this limit is hard to reach because other

noise sources are usually higher. But even if a sensor is shot-noise limited in a specific power

range, we could not indefinitely decrease the MDS by increasing the detected power because

at high enough power (above ~80 µW in the specific example of Fig. 1.4) other noise sources

Page 19: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

7  

proportional to power, such as laser intensity noise and frequency noise, would become

dominant.

In 2010, around the same time as Wen’s experiment, Gagliardi et al. [11] reported a fiber

strain sensor utilizing one of the slow-light resonances of a 10-cm FP made of two FBGs

fabricated on the same fiber. Using the same general principle as described in Fig. 1.1a, the

authors were able to measure an MDS of ~220 fε/√Hz at 1.5 kHz and ~700 fε/√Hz at 2 Hz,

which are the lowest MDS reported in a passive, relatively short fiber sensor prior this work.

These figures are remarkably low, especially considering the low frequencies of these

measurements. One downside of this sensor was that it was 5 times longer than Wen’s sensor,

which reduced its ability to detect a strain confined to a small area. Furthermore, to reduce the

laser frequency noise a relative complicated approach was used that involved locking the

interrogating laser to an ultrastable frequency comb. That publication [11] also claimed that

the sensor noise was limited by the fiber thermal phase noise, namely noise arising from

thermodynamic temperature fluctuations of the fiber. Specifically, when light propagates

through a medium, thermodynamic temperature fluctuations induce random fluctuations of

both the refractive index and length of the medium, which in turn impart noise to the phase of

the light traveling through it. In 1992 Wanser published a simple formula for the power

spectral density (PSD) of the thermal phase noise of a signal that has traveled once through an

optical fiber [12]. In this case the thermal phase noise is proportional to the square root of the

length of the fiber, making it extremely difficult to measure it and to verify this expression in

short fibers. Nevertheless, this expression can be conveniently used to predict the phase-noise-

limited minimum detectable phase shift in the large number of fiber sensors in which light

travels through a fiber only once, as in a Mach-Zehnder interferometer [13]. Wanser

subsequently extended this study to the phase noise of a signal that has traveled twice through

the same fiber [13], as occurs for example in a Michelson interferometer. It was, however, not

straightforward to extend it to the case of a multi-pass interferometer like an FP.

Gagliardi’s claim of a sensor limited by phase noise limited was subsequently challenged and

proven to be incorrect by Cranch and Foster [14], who showed via theoretical considerations

that in [11] the phase noise was calculated incorrectly and was in reality negligible. These

results raised a controversy in the community regarding how the phase noise scales with group

index in a resonant cavity, a calculation that had not been published at the time. Gagliardi et

Page 20: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

8  

al. assumed that the phase noise is proportional to the square root of the group index.

Therefore they calculated the phase noise of their sensor by replacing the length of the fiber in

Wanser’s formula with an effective length equal to the length of their FP times the group

index. This calculation inherently assumed that the phase noise accumulated by the light

during different passes through the FP were uncorrelated, but this is not true. As we showed

subsequently in a detailed study of the phase noise in FP interferometers [15], these

contributions are correlated, and as a result the total phase noise at the output of an FP is

proportional to the group index. On the other hand, Cranch and Foster [14] assumed that the

phase noise doesn’t scale with group index at all, because they used a different definition of

phase noise without explicitly stating it, as explained in [16, 17]. When the work reported in

this thesis started, the dependence of the phase noise at the output of a resonant cavity was

therefore an open question that needed to be answered carefully in order to first understand the

magnitude of the contribution of phase noise in Wen’s slow-light sensors, and second to

design a short passive strain sensor limited by phase noise, which had not been done before.

This was another major objective of this thesis work.

Another elegant and highly sensitive method for detecting a very small strain with an FBG is

to write the FBG in an active fiber to create a fiber laser [18,19]. The lasing wavelength

depends on the Bragg wavelength. Thus this sensing technique is similar to what has been

described for passive slow-light FBGs. A strain applied to the fiber laser induces a shift in the

laser wavelength, and an interrogation system is set up to measure this shift. The MZI

approach described earlier [5] has been used extensively for interrogating active FBG strain

sensors [18,19]. As discussed earlier, in passive FBG strain sensors utilizing an MZI as a

wavelength readout the main limitation is that the bandgap of the FBG can only be made so

narrow before its reflectivity becomes too small for useful lengths. But this is not the case for

active fiber laser sensors. Fiber lasers can have very narrow linewidths (less than 200 Hz)

[20], thus this interrogation technique is not limited by the coherence length of the light input

into the MZI. As a result, an MZI with a large length imbalance can be used, which leads to

much larger sensitivities and therefore much smaller resolvable wavelength shifts and strains.

Using an active FBG-based laser as a strain sensor, a measured MDS of ~60 fε/√Hz at 7 kHz

was reported in [18], and ~120 fε/√Hz at 2 kHz in [19] (see Fig. 1.2). At the start of this work

these were the lowest reported strain resolutions in a fiber sensor.

Page 21: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

9  

1.4 Summary of this thesis work

At the end of Wen’s work in 2011, it was clear that in order to reduce the MDS of slow-light

FBG sensors and reach or even exceed the resolution level of active FBG-based sensors, two

objectives needed to be met. We needed to reduce the frequency noise of the laser, and to gain

a sound understanding of the next noise source that would potentially limit our MDS, which

was phase noise. Thus the main objective of this thesis was first to study theoretically, then to

measure, the thermal phase noise in a passive fiber resonator, with a view to reduce the

measured MDS in a slow-light FBG strain sensor. This was a four-pronged effort. First, a

theoretical model was develop to predict the thermal phase noise of an FP cavity and its

relationship to the FP length and group index. Second, we showed how to increase the group

index in very short FBGs in order to achieve high group delays and high strain sensitivities.

This study was necessary to observe phase noise in an FBG and to explore other possible

applications of these slow-light FBGs, for example as high-Q resonators for nonlinear optics

and as optical delay lines. The third prong was to improve the experimental sensor developed

by He Wen to (1) reduce the laser frequency noise; (2) reduce the environmental noise; and

(3) improve the long-term stability of the sensor. The fourth and final prong was to use these

various developments and improvements to design, fabricate, and test two FBGs, one to

measure the thermal phase noise in an FBG for first time, and the other to observe the smallest

strain ever measured in an FBG-based sensor.

The first part of this effort, the development of a theoretical model of phase noise in a Fabry-

Perot interferometer, is discussed in Chapter 3. We show that at least at low frequencies, the

phase noise in proportional to the group index, and that when expressed in units of strain it is

inversely proportional to the square root of the length of the FP cavity. Thus, longer cavities

have a lower phase noise expressed in units of strain. These results enabled us to ultimately

design and test FBG sensors with a noise limited by phase noise and record-high strain

sensitivities.

The second part of this work, the design of FBGs with very strong resonances, is discussed in

Chapter 4. Since a short FBG was required to measure phase noise and verify our theory, we

needed to maintain the same sensitivity as, or achieve a higher sensitivity than, in previous

work but in a much shorter FBG. As discussed in Chapter 2, where the basic theory of strain

Page 22: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

10  

sensing using FBG is presented, to achieve this result the group index in an FBG had to be

increased to increase the sensitivity and/or amplify the phase noise. We were able to achieve

high group delays and sensitivities in short FBGs by using strongly apodized FBGs that

exhibit a high index modulation while maintaining approximately the same low loss as in

earlier FBGs. These FBGs were realized in collaboration with Prof. Martin Bernier at

University Laval in Québec, who fabricated all the FBGs tested in this work according to our

specifications. The FBGs were written using an infrared femtosecond laser in a deuterium-

loaded fiber, a technique that has been shown, in part as an outcome of this work, to increase

the maximum achievable index modulation and reduce the loss for the same index modulation.

These FBGs were thermally annealed to further reduce their internal loss and optimize them

for our applications. These improvements led to a record high group delay of 42 ns [21] and

revealed other interesting properties of slow-light FBGs, in particular high Purcell factor and

high intensity enhancement, that are discussed in Chapter 6. These properties can enable new

applications of slow-light FBGs, such as low-threshold lasers, enhanced nonlinear effects, etc.,

which can be useful in telecommunications.

The last two efforts are reported in Chapter 5. Specifically, first we discuss the improvements

made in the experimental sensor, including the acquisition of a new laser with very low

frequency noise, isolating the sensor in an anechoic enclosure to reduce the environmental

noise and improve the overall stability and noise performance of the sensor, and reducing the

detector noise. Next, we combined all the conclusions from the previous investigations to

design, fabricate, and test two FBGs compatible with our improved setup, one with a high

enough sensitivity and a large enough phase noise to measure the thermal phase noise in an

FBG for first time, and the other with a low enough overall noise to observe the smallest strain

ever measured in an FBG-based sensor. The first, 5 mm in length, enabled us to confirm the

magnitude of the phase noise, and its frequency dependence in the few-kHz range, predicted

by the model presented in Chapter 3. The second one, 20 mm in length, exhibited the smallest

MDS ever reported in a passive FBG, namely 110 fε/√Hz at 2 kHz and 30 fε/√Hz at 30 kHz,

which is about two [11] and about ten [8] times smaller than the previous records,

respectively. The output of this sensor was so stable (no sign of drift in four days) that a

record-low absolute MDS of 250 attostrains at 30 kHz was observed by integrating a 4-day

output trace with an 8-hour integration time. This sensor was so sensitive that it enabled us to

determine that the noise level in our laboratory was higher during the night than during the

Page 23: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

11  

day, an observation that could not be made with the previous generation of sensors. This result

shows that this new generation of ultra-sensitive sensors can be used to study phenomena that

were not detectable before.

Finally, Chapter 7 summarizes the salient results of this work and proposes possible future

research directions in slow-light FBGs.

References [1] R. Di Sante, “Fibre Optic sensors for structural health monitoring of aircraft cmposite

structures: recent advances and applications,” Sensors, 15, 18666 (2015).

[2] C. I. Merzbacher, A. D. Kersey, and E. J. Friebele, “Fiber optic sensors in concrete

structures: a review,” Smart Materials and Structure. 5, 196 (1996).

[3] N. Beverini, Calamai, D. Carbone, G. Carelli, N. Fotino, F. Francesconi, S. Gambino,

R. Grassi, E. Maccioni, A. Messina, M. Morganti, and F. Sorrentino, “Strain sensors

based on Fiber Bragg Gratings for volcano monitoring,” in Fotonica AEIT Italian

Conference on Photonics Technologies, 2015, 1 (2015).

[4] B. Abbott et al., “LIGO: the laser interferometer gravitational-wave observatory,”

Reports on Progress in Physics, 72, 076901 (2009).

[5] A. D. Kersey, T. A. Berkoff, and W. W. Morey, "High resolution fibre-grating based

strain sensor with interferometric wavelength-shift detection," Electronic Letters, 28,

136 (1992).

[6] B. Lissak, A. Arie, and M. Tur, " Highly sensitive dynamic strain measurement by

locking lasers to fiber Bragg gratings," Optics Letters, 23, 1930 (1998).

[7] D. Gatti, G. Galzerano, D. Janner, S. Longhi, and P. Laporta, "Fiber strain sensor

based on a π-phase shifted Bragg grating and the Pound-Drever-Hall technique,"

Optics Express 16, 1945 (2008).

[8] H. Wen, G. Skolianos, S. Fan, M. Bernier, R. Vallée, and M. J. F. Digonnet, “Slow-

light fiber-Bragg-grating strain sensor with a 280-femtostrain/√ Hz resolution,”

Journal of Lightwave Technology, 31, 1804 (2013).

[9] T. Erdogan, “Fiber grating spectra,” Journal of Lightwave Technology, 15, 1277

(1997).

[10] J. E. Sipe, L. Poladian, and C. M. De Sterke “Propagation through nonuniform grating

structures,” Journal of Optical Society of America A 11, 1307 (1994).

Page 24: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

12  

[11] G. Gagliardi, M. Salza, S. Avino, P. Ferraro, and P. De Natale, “Probing the ultimate

limit of fiber-optic strain sensing,” Science, 330, 1081 (2010).

[12] K. H. Wanser, “Fundamental phase noise limit in optical fibres due to temperature

fluctuations,” Electronic Letters, 28, 53 (1992).

[13] K. H. Wanser, “Theory of thermal phase noise in Michelson and Sagnac fiber

interferometers,” in Proceedings of Tenth Fibre Sensors Conference, 584 (1994).

[14] G. A. Cranch, and S. Foster, “Comment on “Probing the ultimate limit of fiber-optic

strain sensing,” Science, 335, 6066 (2012).

[15] G. Skolianos, H. Wen, and M. J. F. Digonnet, “Thermal phase noise in Fabry-Pérot

resonators and fiber Bragg gratings,” Physical Review A, 89, 033818, (2014).

[16] S. Foster, and G. A. Cranch “Comment on “Thermal phase noise in Fabry-Pérot

resonators and fiber Bragg gratings”,” Physical Review A, 92, 017801 (2015).

[17] G. Skolianos, and M. J. F. Digonnet, “Reply to “Comment on ‘Thermal phase noise in

Fabry-Pérot resonators and fiber Bragg gratings,” Physical Review A, 92, 017802

(2015).

[18] K. P. Koo, and A. D. Kersey, “Bragg grating-based laser sensors systems with

interferometric interrogation and wavelength division multiplexing,” Journal of

Lightwave Technology, 13, 1243 (1995).

[19] G.A . Cranch, G. M. H. Flockhart, and C. K. Kirkendall, “Distributed feedback fiber

laser strain sensors,” IEEE Sensors Journal, 8, 1161 (2008).

[20]  H. Jiang, P. Lemonde, G. Santarelli, and F. Kefelian, "Ultra low frequency noise

laser stabilized on optical fiber spool," Frequency Control Symposium, 2009 Joint

with the 22nd European frequency and time forum. IEEE International, 815, Besançon

(2009).

[21] G. Skolianos, A. Arora, M. Bernier, and M. J. Digonnet, “Slowing down light to 300

km/s in a deuterium-loaded fiber Bragg grating,” Optics Letters, 40, 1524 (2015).

Page 25: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

13  

Chapter 2: Basic theory of strain sensing

using a slow-light FBG

In this chapter we discuss how a slow-light FBG can be used for strain sensing. Specifically,

the fundamentals of FBG theory are presented: what is an FBG, and how its transmission and

reflection spectra are simulated from its postulated index-modulation and loss profiles. We

then discuss the basic principles of conventional FBG strain sensors and  of the more recent

FBG strain sensors that utilize slow-light resonances involved in this work. For the latter, we

establish the dependence of their strain sensitivity on the group index and transmission of the

resonance, and discuss their main sources of noise. The final section reviews the various

methods that have been developed to induce slow-light resonances in an FBG prior to this

work, including the introduction of a π phase shift in the approximate middle of the grating

and strong uniform FBGs. The use of strong apodized FBGs to induce slow-light resonances,

the approach that is central to this thesis work, will be discussed in detail in Chapter 4. For

these schemes we also present the scaling laws of the group index dependence on the grating

parameters, which will be useful later to maximize the strain sensitivity.

2.1 Fiber Bragg Grating

A fiber Bragg grating (FBG) is a fiber in which the refractive index of the core region is

periodically modulated over a short length, usually a few mm up to 10 cm, with a period Λ.

This modulation is illustrated in Fig. 2.1 for the particular case of a grating with a uniform

index modulation. The effective refractive index of the unperturbed fiber (before the grating

was written) is n0, and Δnac is the half peak-to-peak amplitude of the index modulation. Because

this modulation does not generally have zero average value, the modulation in the mean

effective refractive index of the fiber also has some finite value Δndc. When these profiles are

constant, the grating is said to be uniform. In all other cases it is generally referred to as

apodized.

For a uniform grating with a sinusoidal modulation the general function that describes the

effective refractive index n(z) as a function of the position z along the grating, is [1]:

Page 26: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

14  

n(z) = n0 +Δndc +Δnac cos2π zΛ

#

$%

&

'(         (2.1)

Figure 2.1. Periodically varying refractive index in a uniform FBG.

As light passes through the FBG, the refractive index is constantly changing. As a

consequence, light inside the FBG gets partially reflected due to Fresnel reflection wherever

and index change takes place, and the reflected light is coupled to the same fundamental mode

but traveling in the opposite direction. The reflections are strongest where the change in the

refractive index is steepest. Thus two strong reflections take place per period with opposite

phases, one on each side of the middle of each modulation period. The phase difference

between two reflections that occur at two different points one period apart is 2π<n>2Λ/λ,

where <n> is the mean effective refractive index, equal to n0+Δndc, and λ is the wavelength of

the incident light in vacuum. This is true because the light that is reflected at the second point

in the FBG has traveled an extra round-trip distance equal to twice the period. Thus when the

wavelength is equal to 2(neff + Δndc)Λ all the reflections from different periods add in phase.

This wavelength is called the Bragg wavelength, and it is given by [1]:

λB (z) = 2(n0 +Δndc )Λ           (2.2)

Using the coupled-mode theory for the two counter-propagating modes, a transfer matrix that

connects the forward and the backward electric fields at the output ( Eout+ and Eout

− , respectively)

and the input ( Ein+ and Ein

− , respectively) of an FBG with length L can be calculated. This

transfer matrix is [1]:

Page 27: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

15  

Ein+

Ein−

"

#$$

%

&''=

cosh κ 2 −σ 2 ⋅L( )-i σ

κ 2 −σ 2sinh κ 2 −σ 2 ⋅L( ) -i κ

κ 2 −σ 2sinh κ 2 −σ 2L( )

i κ

κ 2 −σ 2sinh κ 2 −σ 2L( ) cosh κ 2 −σ 2 ⋅L( )+ i σ

κ 2 −σ 2sinh κ 2 −σ 2 ⋅L( )

"

#

$$$$$

%

&

'''''

Eout+

Eout−

"

#$$

%

&''

(2.3)

where κ and σ are the ac and dc coupling coefficients, respectively, given by:

σ = 2π n0 +Δndc( ) 1λ−1λB

#

$%

&

'(− i

γ2

(2.4a)

κ = πΔnacλ

(2.4b)

where γ is the power loss coefficient that characterizes propagation in the grating. From Eq.

2.3 the field reflection coefficient r and the field transmission coefficient t can be calculated

as:

r =−κsinh κ 2 −σ 2L( )

σsinh κ 2 −σ 2L( )+ i κ 2 −σ 2 cosh κ 2 −σ 2L( ) (2.5a)

t = i κ 2 −σ 2

σsinh κ 2 −σ 2L( )+ i κ 2 −σ 2 cosh κ 2 −σ 2L( ) (2.5b)

In the case of a lossless propagation through the FBG, it can be easily calculated that the

power that is transmitted plus the power that is reflected equals the total power, namely

r 2 + t 2 =1 (energy conservation).

Using Eqs. 2.5a and 2.5b, we plotted in Fig. 2.2 the power reflection and transmission spectra

of a uniform FBG with a period Λ = 534.3 nm, a length L = 2 mm, Δnac = Δndc = 0.5x10-3, and

a uniform loss of γ = 0.1 m-1, as an example. As expected, there is a maximum reflectivity at

the Bragg wavelength defined in Eq. 2.2. In the immediate vicinity of this wavelength, a

region known as the bandgap, most of the partial reflections are in phase, and only a small

fraction of the light is transmitted. Far away from the Bragg wavelength all the light is

transmitted and the grating does not affect the propagating light. Close to the bandgap but

outside the bandgap some ripples occur; the reason for their existence is discussed in detail in

Page 28: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

16  

section 2.3.2b. If a stress is applied to the grating, the two spectra retain their respective shape

but they shift to a new Bragg wavelength by a usually relatively small amount (a few nm or

less). This property has been exploited extensively in FBG sensors, as discussed later.

The bandwidth of the bandgap, defined as the bandwidth between the first zero reflectivities

on either side of the Bragg wavelength, can be calculated using [1]:

Δλ = λBΔnacn0

1+ 2n0ΛΔnacL#

$%

&

'( (2.6)

In the limit of strong and/or long gratings (ΔnacL>>λB) of most interest in this work, the

bandwidth Δλ can be reduced from Eq. 2.6 to:

Δλ ≈ λBΔnacn0

(2.7)

Thus, in this limit, as Δnac increases the bandwidth of the bandgap also increases.

 Figure 2.2. Simulated power (a) reflection and (b) transmission spectrum for a uniform FBG of length

L = 2 mm, index modulation Δnac = Δndc = 5 x 10-4, period Λ= 534.3 nm, and loss γ = 0.1 m-1.

Δnac also determines the maximum reflectivity of the FBG. From Eqs. 2.4a and 2.5a it can be

seen that in the lossless case, at λ = λΒ, we have σ = 0 and r = itanh(κL). Thus the reflectivity

Page 29: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

17  

increases as κ increases. Consequently, a high κ, which is proportional to Δnac, leads to a high

reflectivity. This is an important property that we use extensively later to design ultra-sensitive

sensors. Also, this last equation states that the reflected light is phase shifted by π/2 relative to

the incident light. When light is slightly detuned from the Bragg wavelength, the total phase

shift of the reflected light is different from π/2, because each reflection has a slightly different

phase. This is why in equation 2.5 the reflection and the transmission coefficients are complex

numbers and their phase is wavelength dependent. Using this phase, the group delay τg and the

group index ng in transmission can be calculated from the following equations

τ g = −dφtdω

=ngLc           (2.8a)

ng =cvg           (2.8b)

 

where φt is the phase of the transmitted electric field relative to the input field, ω is the angular

frequency of the light, c the speed of light in free space, and vg the group velocity. Using an

equation similar to Eq. 2.8a, the group delay in reflection can also be calculated [1].

2.2 Strain sensing using FBGs

2.2.1 General principle of strain sensing using an FBG

If a longitudinal strain ε is applied to an FBG, three effects take place [2]:

1) The period of the FBG is changed by a quantity ΔΛ which, assuming that the applied

strain is uniform, is given by:

ΔΛΛ

=ΔLL= ε (2.9)

2) The refractive index n of the material is changed due to the applied strain through the

elasto-optic effect

3) The diameter of the fiber changes because of the change in fiber length, causing the

propagating mode to slightly change and hence the effective refractive index to

change. This effect is negligible.

Page 30: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

18  

The second and third effects can be lumped together as a change in the effective refractive

index of the propagating mode n0 expressed as:

Δn0n0

=1n0dn0dε

ε (2.10)

By taking the partial derivative of λB (Eq. 2.2), it is clear that these three effects result in a

shift in the Bragg wavelength given by the sum of Eq. 2.9 and 2.10:

ΔλBλB

=ΔΛΛ

+Δn0n0

⇒ΔλB = λB 1+1n0dn0dε

$

%&

'

()ε (2.11)

The wavelength shift, which applies to the Bragg wavelength and the entire spectrum, is

proportional to the applied strain ε, the Bragg wavelength λΒ, and a material constant,

(1+1/n0•dn0/dε). For silica fibers, which are most commonly used in telecommunications and

fiber sensors, this material constant has been measured to be 0.78 [2]: the change in grating

period accounts for a positive contribution equal to 1, and the strain-induced change for a

negative contribution equal to -0.22. Thus, for a silica fiber, the reflection-spectrum shift due

to an applied strain is given by:

ΔλB = 0.78λBε         (2.12)

Figure 2.3. Shift of the Bragg wavelength and the whole spectrum due to an applied strain. The initial

transmission is indicated with the solid line while the spectrum after a strain has been applied to the

FBG is shown with the dashed line.

 

Page 31: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

19  

The strain applied to an FBG is then almost always sensed by measuring the strain-induced

shift in the reflection or transmission spectrum (see Fig. 2.3). An obvious way to do this is by

launching broadband light into the FBG and using an optical spectrum analyzer (OSA) to

measure the spectral shift induced by a strain. However, because of the relatively low

resolution of OSAs, the strain resolution is not very good. A typical OSA has a resolution of

~10 pm. For an FBG with a Bragg wavelength at 1.55 µm, Eq. 2.12 gives ΔλB=1.2 pm/µε,

leading to a minimum detectable strain (MDS) of only 8 µε.

To overcome the limited resolution of OSAs, the FBG can interrogated with a laser tuned to

the steepest slope of the FBG reflection or transmission spectrum, as discussed in the

introduction. In this scheme the change in the output power is measured, and the applied strain

can be recovered from this measurement and the knowledge of the spectrum’s slope at the

probe wavelength, which is also obtained easily with a measurement of the spectrum.

For this general type of sensing scheme, the normalized strain sensitivity SN is defined as:

SN (λprobe ) =1Pin

dPoutdε

!

"#

$

%&λprobe

(2.13)

where dPout is the change in output power at the probe wavelength λprobe in response to a small

applied strain dε, and Pin is the power input into the FBG. Since the transmission is by

definition T = Pout/Pin, this expression can be rewritten as:

SN (λprobe ) =1Pin

dPoutdε

!

"#

$

%&λprobe

=dTdε

!

"#

$

%&λprobe

=dTdλ

dλdε

!

"#

$

%&λprobe

=dλBdε

dTdλ!

"#

$

%&λprobe

  (2.14)

From Eq. 2.14 it can be seen that the sensitivity is proportional to the slope of the transmission

spectrum dT/dλ evaluated at the probe wavelength. Essentially, this slope converts the

wavelength shift into a change in transmission, and therefore a change in output power. In this

case the MDS is limited by how well a small change in the output power can be resolved,

namely by the output power noise Pnoise, since this is the smallest change in the output power

that can be measured. Thus the MDS, or equivalently the strain resolution of the sensor, is

defined as:

Page 32: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

20  

MDS= PnoisePin

1SN         (2.15)

Stated differently, the MDS is the noise of the sensor expressed in units of strain. To achieve a

low MDS, the output noise power (relative to the input power) must be low and/or the

sensitivity must be high. Lissak et al. were first to report the use of the steep slope in the

FBG’s band edge to achieve a very low MDS (~50 pε/√Hz) [3], as discussed in the

introduction.

2.2.2 Using a sharp resonance to increase the strain sensitivity

The slope and the sensitivity can be significantly increased by using a sharp resonance.

Assume that we can fabricate an FBG with a very narrow transmission peak, as illustrated in

Fig. 2.4. As discussed previously, if the peak is interrogated on the steepest slope a

wavelength shift is converted to a transmission change. This change increases as the slope of

the peak is increased. From Fig. 2.4 one can easily see that if the peak linewidth is decreased,

and/or if the peak transmission is increased, the slope is increased, the sensitivity increases, as

stated by Eq. 2.14.

Figure 2.4. Using a sharp peak for strain sensing.

If the peak is narrow, by definition the quality factor is high too, and so are the group delay

and the group index of the resonance. So we expect intuitively that the sensitivity scales like

the group index, or the group delay. The relationship between sensitivity and group delay can

Page 33: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

21  

be readily expressed in a closed form in the particular case of a resonance with a Lorentzian

lineshape, which is applicable to many structural slow-light resonances (when the Q factor of

the cavity is spoiled, the energy decays exponentially, hence the lineshape is Lorentzian). In

this case it is easy to show, by taking the derivative relative to the wavelength of the general

expression of a Lorentzian transmission function and looking for the wavelength where this

derivative is maximum, that the maximum strain sensitivity is [4]:

SNmax =

3.22cT0τ g,maxλ

          (2.16)

where T0 is the maximum transmission of the resonance (i.e., at the linecenter of the

resonance) and τg,max is the maximum group delay of the resonance (i.e., at the linecenter of the

resonance). Equation 2.16 shows that the strain sensitivity is proportional to the peak

transmission and the peak group delay of the resonance, as was expected based on the

previous discussion. It is worth mentioning that even if the peak is not exactly a Lorentzian,

the scaling laws of Eq. 2.16 are still valid; only the numerical factor (3.22) differs a little. Thus

for a highly sensitive sensor a slow-light peak with a large peak transmission times group

delay product is needed. From now on, we will concentrate solely on strain sensors that utilize

a sharp resonance to achieve a high sensitivity.

2.2.3 Noise sources in an ultra-high sensitive strain sensor

According to Eq. 2.15, to reduce the MDS we not only have to increase the sensitivity, but

also to reduce the noise. Noise sources can be grouped in two general categories: (1) noise

sources that are independent of the sensor’s strain sensitivity and (2) noise sources that depend

on the sensitivity.

The first category, sensitivity-independent noise sources, includes the intensity noise of the

laser, the photodetector noise, and the shot noise (electrical and optical). The intensity noise of

a laser depends solely on the laser type and its output power. The photodetector noise is a

characteristic of the photodetector alone. Finally, the shot noise depends on the power of the

detected optical signal and exists because of the quantum nature of the light. Because these

Page 34: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

22  

noise sources are independent of the sensitivity, their individual contribution to the MDS

decreases as the sensitivity is increased (see Eq. 2.15).

The second category, sensitivity-dependent noise sources, includes the intrinsic thermal phase

noise of the fiber, the laser frequency noise, and environmental noise. The thermal phase noise

is caused by random thermodynamic fluctuations of the fiber temperature, which induce

variations in the optical path length experienced by light that propagates through the fiber.

This noise can be interpreted as a random strain applied to the FBG, since the strain has the

same effect—a change in optical path length— and also causes the slow-light resonance

frequency to fluctuate randomly around a mean value. The laser frequency noise, as discussed

in the introduction, is random fluctuations in the laser mean wavelength. Thus the position of

the laser wavelength relative to the resonant wavelength of the resonance fluctuates randomly.

This has the same effect as if the resonance used for sensing was randomly fluctuating around

a mean value. Finally, the environmental noise, such as vibrations, acoustic noise, etc., is an

actual signal that is not useful but constitutes a real strain signal that is also detected and

constitutes a source of noise. Thus, these three noise sources introduce an uncertainty in the

relative spectral position of the slow-light resonance and the laser signal, as a true strain signal

would do. The FBG being interrogated on the slope of this resonance, it converts each of these

three sources of noise into a power noise at the detector. Consequently, each of these three

noise sources produces power noise that is proportional to the slope of the transmission at the

probe wavelength, and therefore to the sensitivity.

According to this analysis the output power noise, and consequently the MDS, can be

decomposed into two main components, one due to the sensitivity-independent noise sources

Pi, and one due to the sensitivity-dependent noise sources Pd=ASN, where A is a factor that

doesn’t depend on the sensitivity. The MDS can therefore be written formally as:

MDS = PnoiseSNPin

=Pi2 +Pd

2

SNPin=1Pin

Pi2

SN2 + A

2 (2.17)

From Eq . 2.17 it can be clearly seen that if the sensitivity is high enough (SΝ >> Pi/A), the

sensitivity-independent noise sources become negligible. On the other hand at low sensitivities

the noise sources that are sensitivity dependent become dominant. In this limit, the MDS is

Page 35: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

23  

given by A/Pin and is independent of the sensitivity. Thus, increasing the sensitivity further

does not improve (reduce) the MDS.

2.3 Temperature sensitivity of slow-light FBGs

The analysis in the previous section can be applied also for other quantities that change the

Bragg wavelength of the FBG, and hence shift the spectrum, like humidity, acoustic waves,

temperature, etc. As long as a change in the environment shifts the Bragg wavelength, this

change is converted to a change in the output power of the FBG—which is proportional to the

slope of the operating point and therefore the group delay—via the shift in the FBG’s

spectrum. In this section we focus on thermal sensitivity and temperature sensing because it is

important for the thermal stability of our strain sensor.

As in the previous case of an applied strain, when the temperature of an FBG changes three

effects take place [5]:

1) The FBG expands longitudinally, which changes its period;

2) The refractive index of the fiber material changes, and hence so does the effective

refractive index of the propagating mode;

3) The diameter of the fiber changes, which changes the mode’s effective index too.

All three effects on the Bragg wavelength can be modeled as [5]:

ΔλB = λB ξ +1n0dn0dK

"

#$

%

&'ΔK (2.18)

where ξ is the thermal expansion coefficient of the fiber and ΔK is the temperature change.

For a silica fiber, ξ=5x10-7 K-1and dn0/dK=1.1x10-5 K-1. Thus, at a wavelength of 1.55 µm

ΔλB = 12.5 pm/˚K.

The normalized temperature sensitivity ST of the FBG is defined similarly to its normalized

strain sensitivity (Eq. 2.14):

Page 36: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

24  

ST (λprobe ) =1Pin

dPoutdK

!

"#

$

%&λprobe

=dTdK!

"#

$

%&λprobe

=dTdλ

dλdK

!

"#

$

%&λprobe

=dλBdK

dTdλ!

"#

$

%&λprobe

(2.19)

Hence, the temperature sensitivity is also proportional to the slope dT/dλ of the resonance at

the operating wavelength, as expected. Equation 2.19 and 2.14 differ only by the factor in

front of the slope. Therefore the analysis of sections 2.2.2 and 2.2.3 is directly applicable to

the temperature sensitivity and temperature sensing. In fact it can be used to model sensing of

any parameter, as long as the factor for this specific parameter is used to scale the slope. This

shows that as the strain sensitivity is increased by increasing the slope of the resonance, the

temperature sensitivity is increased too, and the thermal stability of the sensor is degraded. On

the other hand, this high thermal sensitivity allowed us to measure random temperature

fluctuations in the atomic level of the fiber (phase noise), as we discuss in Chapter 5.

In practice other changes in the environment of the fiber besides temperature drift have the

same effect: they change the output power of the FBG when the latter is probed on the edge of

a resonance. In general it is hard to distinguish the different sources that cause the power

change. It can be done, however, by using multiple FBG sensors co-located on the same or

nearby fibers and comparing their outputs [6-8]. The thermal drift can also be greatly reduced

by mounting the FBG on a support that compensates for the direct effect of the temperature

change on the FBG (athermal FBGs) [9]. In our sensor, because we were interested in

measuring a dynamic strain, we were able to easily isolate the slowly-varying thermal drift

from the fast strain using a lock-in amplifier, which isolates the specific frequency of the

strain and effectively filters out the signals at other frequencies, in particular the slow thermal

drift. We were also able to cancel out the effect of the drift on the sensitivity by using a

feedback loop with a proportional-integral-derivative (PID) controller, as discussed in detail in

Chapter 5.

2.4 Slow-light resonances in an FBG

Until now we have explained how a highly sensitive sensor can be achieved in principle by

making use of a slow-light resonance with a high transmission. A well-known optical device

that exhibits such resonances is a Fabry-Perot (FP) interferometer. In this section the group

delay and group index of the resonances of an FP are derived algebraically. This derivation

Page 37: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

25  

provides general scaling laws for the group delay and consequently the strain sensitivity of an

FP. We then discuss how an FP can be formed in a single FBG, and how the scaling laws

derived for an FP apply to the FP-like resonances of slow-light FBGs.

2.4.1 Slow-light resonances in a Fabry-Perot interferometer

We consider a conventional Fabry-Perot interferometer with two parallel reflectors M1 and M2

with field reflection coefficients r1 and r2, and transmission coefficients t1 and t2, respectively.

The mirrors M1 and M2 can be two FBGs, or any number of conventional optical reflectors. In

the particular case of two FBGs of interest here, the field reflection and transmission

coefficients are the quantities defined in Eqs. 2.5a and 2.5b, respectively. The distance

between the two mirrors is L and the medium between the mirrors has a uniform refractive

index n (Fig. 2.5).

Figure 2.5. Schematic of a generic Fabry-Perot interferometer

These four coefficients (r1, r2, t1 and t2,) implicitly incorporate the medium’s single-pass

transmission exp(-αlossL), where αloss is the medium’s loss coefficient. In such an FP the

transmission after the second mirror is given by [10]:

Et

E0

=t1t2

1− r1r2ej2Φ1

ej Φ1−π−

ωnLc

#

$%

&

'(

Φ1 = 2πnL / λ +ϕr1 +ϕr2

2

(2.20)

where φr1 and φr2 are the phase shift associated with reflection on the first and the second

mirror, respectively. The factor π in the phase term arises from the phase associated with

Page 38: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

26  

transmission through each of the two mirrors, φt1 = φr1 - π/2 and φt2 = φr2 - π/2. The group delay

of the transmitted signal is by definition (Eq. 2.8a) proportional to the derivative of the

argument of the right-hand side of Eq. 2.20, or

ng =ddω

2Φ1 −π −ωnLc

− tan−1 r 1r2 sin2Φ1

r 1r2 cos2Φ1 −1#

$%

&

'(

#

$%%

&

'((cL     (2.21)

After some straightforward algebraic manipulations, it is easy to prove from Eq. 2.21 that

ng = n 2 1− r1r2 cos(2Φ1 )1+ r1r2( )

2− 2r1r2 cos(2Φ1 )

−1#

$%%

&

'(( (2.22)

On a resonance, Φ1 is a multiple of π, and this expression reduces to:

ng = n2

1− r1r2−1

"#$

%&' = n

2Fπ r1r2

−1"

#$

%

&' (2.23)

where F is the finesse of the FP.

In the limit of high finesse:

ng ≈ n2Fπ

(2.24)

A similar derivation applied to the signal reflected by the FP gives the relationship for the

group index in reflection ng ' :

ng' =

2r2t12

1− r1r2( ) r2 r12 + t1

2( ) − r1( ) (2.25)

In the limit of high finesse, Eq. 2.25 reduces to Eq. 2.24.

 

Page 39: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

27  

Equation 2.23 states that in order to achieve a large group delay, the round-trip loss of the

cavity (1-r1r2) must be small, namely the mirrors must have high reflectivity, and the

absorption loss in the mirrors and in the medium between the mirrors must be low. Also, from

Eq. 2.20 it can be seen that for the transmission to be high the absorption loss must be low,

namely the round-trip loss of the cavity must be as close as possible to the light that is

transmitted outside the cavity from the mirrors. Thus, if the two mirrors were FBGs, a high

sensitivity requires that the two FBGs have a high Δnac and a low loss.

2.4.2 Forming an FP and highly sensitive slow-light resonances

In this section we discuss how FP-like resonances can be created in a single FBG. To create

slow-light resonances an effective FP needs to be formed. Thus the analysis of section 2.3.1 is

directly applicable to FBGs as well. An effective FP can be formed in a single FBG using one

of the following approaches:

1) A π-shifted FBG

2) A strong uniform FBG

3) A strong apodized FBG

The first two approaches, which have been investigated before to demonstrate ultra-sensitive

strain sensors [4, 11], are described in this chapter. The third approach, using an apodized

FBG, is described in Chapter 4.

2.4.2a Slow-light in π-shifted FBGs

A π-shifted FBG is an FBG with a π phase shift in the middle of the grating’s refractive index

profile.  This type of grating can be conceptualized as two nominally identical FBGs separated

by a gap of length λ/4 (a phase shift of π/2 rad), which means that light traveling in the gap

accumulates a phase of π per round trip. Each grating acts as a reflector, and the two gratings

together form an FP separated in the phase space by π. This FP is so short that it supports a

single transmission resonance, located around the middle of the grating’s bandgap (Fig. 2.6).

This resonance can have a high sensitivity to strain because this type of FBG can exhibit a

high Δnac.

Page 40: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

28  

Figure 2.6. Calculated reflection spectrum of a representative π -shifted grating (the grating’s index

profile is shown schematically in the inset).

Very often in a π-shifted FBG, because of fabrication imperfections, for example if the phase

shift is not exactly π or not exactly in the middle of the FBG, the slow-light resonance is not

exactly in the middle of the bandgap. This can lower the peak transmission of the resonance

because the FP is not symmetric (the two mirrors have different reflectivities), thus the

maximum transmission is lower than in the symmetric case. The merit of this application was

demonstrated with the report of an MDS of 5 pε/√Hz by Gatti et al. in 2008 [11].

2.4.2b Slow-light in uniform FBGs

As we have already discussed earlier in this chapter, in a uniform grating in the middle of the

bandgap the FBG is essentially fully reflective, while far away from the bandgap it is

essentially fully transmissive. Between these two extremes, over a short span of wavelengths

on the sharp edges of the bandgap, the FBG is partially transmissive. Thus the FBG acts as a

distributed partial mirror. The light component that is not transmitted at each point of the FBG

is reflected, travels through the FBG back towards the input, where the portion of it that is not

transmitted is partially reflected again, and so on and so forth. Light therefore makes multiple

round-trip passes through the FBG, as illustrated in Fig. 2.7.

Page 41: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

29  

Figure 2.7. a) For wavelengths near the band edge of a strong FBG, multiple reflections from both ends

of the FBG cause recirculation of light inside the FBG. The situation is analogous to a Fabry-Perot (FP)

interferometer. b) Similar to a FP, sharp resonances are formed in the transmission spectrum.

At certain specific wavelengths, the multiple small output fields that are transmitted at each

pass are in phase and add coherently to each other, which leads to a strong output. The FBG

behaves like a distributed FP [12]. The two mirrors are distributed and spatially overlap. The

optical distance between them is defined through the phase of the light, as in the π-shifted

FBGs, where the physical space between the two FBGs is essentially zero. Like in an FP, the

phase-matching condition for constructive interference occurs only at discrete wavelengths.

Additionally, like in an FP, to produce substantial group delays and therefore high

sensitivities, the FBG must reflect sufficiently over a fairly short distance, which requires in

practice a strong ac index modulation (~10-3 or greater) and a low loss.

A uniform FBG is not the most effective way to generate slow light and high sensitivities.

When the index modulation is uniform, at all slow-light wavelengths the grating reflects

strongly throughout its length, and light experiences a large reflective “loss” instead of being

transmitted by the nominally lossless medium between the two reflectors as it is in typical

free-space or fiber FPs. Although a uniform FBG is not the optimal way to increase the

sensitivity, He Wen used an almost uniform FBG to achieve a record low MDS of 280 fε/√Hz

[4]. In Chapter 4, we will discuss, among other things, how apodizing the FBG’s index-

Page 42: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

30  

modulation profile helps to solve this problem and increases the group delay and consequently

the sensitivity even further.

References [1] T. Erdogan, “Fiber grating spectra,” Journal of Lightwave Technology, 15, 8, 1277

(1997).

[2] A. Kersey, M. A. Davis, H. J. Patrick, M. Leblanc, K. P. Koo, C. G. Askins, M. A.

Putnam, and E. J. Friebele, “Fiber grating sensors,” Journal of Lightwave Technology

15, 1442 (1997).

[3] B. Lissak, A. Arie, and M. Tur, “Highly sensitive dynamic strain measurement by

locking lasers to fiber Bragg gratings,” Optics Letters, 23, 1930 (1998).

[4] H. Wen, G. Skolianos, S. Fan, M. Bernier, R. Vallée, and M. J. Digonnet, “Slow-light

fiber-Bragg-grating strain sensor with a 280-femtostrain/√ Hz resolution,” Journal of

Lightwave Technology, 31, 1804 (2013).

[5] E. J. Friebele, M. A. Putnam, H. J. Patrick, A. D. Kersey, A. S. Greenblatt, G. P.

Ruthven, M. H. Krim, and K. S. Gottschalck, “Ultrahigh-sensitivity fiber-optic strain

and temperature sensor,” Optics Letters, 23, 222 (1998).

[6] J. D. C. Jones, “Review of fibre sensor techniques for temperature-strain

discrimination,” Optical Society of America Technical Digest Series in 12th

International Conference on Optical Fiber Sensors, 16, 36 (1997).

[7] M. Song, S. B. Lee, S. S. Choi, and B. Lee, “Simultaneous measurement of

temperature and strain using two fiber Bragg gratings embedded in a glass tube,”

Optical Fiber Technology, 3, 194 (1997).

[8] W. C. Du, X. M. Tao, and H. Y. Tam, “Fiber Bragg grating cavity sensor for

simultaneous measurement of strain and temperature,” IEEE Photonics Technology

Letters, 11, 105 (1999).

[9] Y. L. Lo, and C. P. Kuo, “Packaging a fiber Bragg grating with metal coating for an

athermal design,”  Journal of Lightwave Technology, 21, 1377 (2003).

[10] G. Skolianos, H. Wen, and M. J. F. Digonnet, “Thermal phase noise in Fabry-Pérot

resonators and fiber Bragg gratings,” Physical Review A, 89, 033818, (2014).

[11] D. Gatti, G. Galzerano, D. Janner, S. Longhi, and P. Laporta, “Fiber strain sensor

based on a π-phase shifted Bragg grating and the Pound-Drever-Hall technique,”

Optics Express, 16, 1945-1950 (2008).

Page 43: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

31  

[12] J. E. Sipe, L. Poladian, and C. M. De Sterke “Propagation through nonuniform grating

structures,” Journal of Optical Society of America A ,11, 1307 (1994)

Page 44: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

32  

Chapter 3: Modeling the thermal phase noise

in a passive Fabry-Perot resonator

As discussed in the two previous chapters, the intrinsic thermal phase noise of the fiber is

expected to be the next lowest source of noise after laser frequency noise in the slow-light

fiber Bragg gratings we have been developing for ultra-sensitive strain sensing. The phase

noise being generally exceedingly small, in most cases smaller than 50 fε/√Hz in terms of

strain, and given the wide disagreement in the predictions of the magnitude of this phase noise

reported in the literature, it was important to perform our own theoretical predictions of its

magnitude and spectral dependence.

This chapter presents an original derivation of the thermal phase noise in a fiber Fabry-Perot

resonator operated in transmission or in reflection. It reveals a number of interesting properties

that remained unpublished until these results were reported in Physics Review A in 2014 [1].

It shows that at low frequencies the power spectral density (PSD) of the phase noise of an FP

interferometer is proportional to the group index of the resonance [1]. As mentioned in the

introduction, this was anticipated. When light propagates through a resonance, it bounces back

and forth in the resonator several times and accumulates phase noise at every pass. Thus the

phase noise in the output light signal is expected to be the phase noise picked up in a single

pass, which is proportional to the square root of the length, times the number of passes, which

is proportional to the group index. Thus, in a resonator the phase noise is expected to be

“amplified” by the group index. The MDS of a sensor is defined as the noise over the

sensitivity (see Chapter 2). Therefore if the MDS is limited by phase noise, it will be

proportional to the square root of the length of the resonator, since the sensitivity is

proportional to the group index times the length of the resonator. These results were later

verified in [2,3] and enabled us to design the sensor presented in Chapter 5, in which the

dominant noise source is the phase noise of the fiber.

Specifically in this chapter, it is shown that although in the most general case the PSD of the

phase noise in an FP can only be expressed as an infinite series, this series can be written in a

simple and convenient closed-form expression under two broad general regimes, namely at

low noise frequencies for any optical frequency (on or off an FP resonance), and at any

Page 45: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

33  

frequency for a signal on resonance. This distinction is important, because for maximum

sensitivity most FP sensors operate slightly off resonance, as discussed in Chapter 2, so this

first closed-form expression is critical to this work. Numerical evaluations of this infinite

series in the case where it cannot be expressed in a closed form (off resonance at high

frequencies) demonstrates that the PSD has a sinc dependence on frequency [1]. This

dependence is valid for an FP of arbitrary finesse. The PSD of the phase noise picked up by a

light signal that has traveled through a fiber twice therefore does not have a cosine dependence

on frequency, unlike reported in [4], but a sinc dependence [1]. Although the two dependences

are asymptotically identical in the low-frequency limit, this correction is significant when

dealing with high-frequency sensors; we predict, unlike previous models, that the phase noise

does not have zeros at certain high frequencies. While all previous publications have been

concerned with FPs operated in reflection, in this chapter a derivation of the phase-noise PSD

for an FP operated in transmission is provided as well. This is of direct relevance in particular

to the ultra-sensitive slow-light FBG sensors reported in this work, which operate in

transmission. This analysis is also instrumental in that it shows how to design FBGs in order

to (1) achieve the lowest possible MDS using our sensing principle, (2) measure the thermal

phase noise in an FBG, and (3) use this measurement to verify the magnitude of the phase

noise predicted by our theory.

3.1 Modeling phase noise in a Fabry-Perot interferometer 3.1.1 Background

As we briefly discussed in the introduction, when light propagates through a fiber,

thermodynamic temperature fluctuations change randomly both the refractive index and length

of the fiber, which in turn translates into noise in the phase of light propagating through the

fiber. Wanser studied this phenomenon and published in 1992 a simple formula for the PSD of

the thermal phase noise of a signal that has traveled once through an optical fiber [5], which

we refer to as the single-pass phase noise. He showed that the single-pass phase noise is

proportional to the square root of the length of the fiber. This dependence is based on the

assumption that the phase noise accumulated in two different points of the fiber is completely

uncorrelated to each other, which is correct at all except possibly exceedingly low (sub-mHz)

noise frequencies. This expression was used to predict the phase-noise-limited minimum

detectable phase shift in a number of fiber sensors in which light travels through a fiber only

Page 46: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

34  

once, as in a Mach-Zehnder interferometer [4]. Wanser later extended this study to the phase

noise of a signal that has traveled twice through the same fiber [4], as occurs, for example, in a

Michelson interferometer. Because slow-light FBG sensors utilize FP-like resonances, it is of

interest to further extend this work to these multi-pass interferometers.

Expressions for the thermal phase noise in multi-pass interferometers used in reflection have

been published [6–9]. However, careful comparison shows that these expressions do not agree

with each other, and often disagree markedly. A recent manifestation of this lack of consensus

is the report by Gagliardi et al. of an FP sensor that was allegedly limited by phase noise [6].

This claim was challenged in [7], which provided mathematical arguments that the phase

noise in the FP resonator of [6] was significantly lower than claimed. The error in [6] came

from the fact that the authors misused Wanser’s formula. They replaced the length of the fiber

in Wanser’s formula with an effective length equal to the group index times the length of the

fiber. This led the authors to the wrongful conclusion that the phase noise is proportional to

the square root of the group index, since they inadvertently made the assumption embedded in

Wanser’s formula that the phase noise accumulated between different passes is completely

uncorrelated. This assumption is correct for two adjacent points seen by light propagating in

one direction, the case studied by Wanser, but as we shall see below it is not for light that sees

the same point when traveling several times through the same fiber, as in an FP. The authors

of [7] also presented an expression for the phase noise in an FP, but they used a different

definition of phase noise without explicitly mentioning it, which led to further confusion.

They essentially calculated the noise in the resonant frequency of the cavity, and called it

phase noise, which resulted in a “phase noise” that is off by a factor of √2 and independent of

group index.

3.1.2 Intuitive picture

 In a fiber Fabry-Perot interferometer probed near resonance, photons travel multiple times

back and forth between the two reflectors. During each pass through the medium contained in

the cavity, they pick up a phase noise that adds to the phase noise already accumulated during

previous passes. Two types of correlation play a role in this system, namely (1) correlation of

the temperature at any two different points along the fiber at a specific time (spatial

correlation), and (2) correlation of the temperature at two different times at a specific point

Page 47: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

35  

(temporal correlation). Regarding the spatial correlation, the temperature variations at two

different points are generally considered to be fully uncorrelated [10]. A consequence of this

property is that for light propagating once through a fiber the phase noise PSDs at each point

add up, and the total PSD is proportional to the fiber length L. Regarding the temporal

correlation, for low perturbation (phase noise) frequencies f, the round-trip transit time through

the cavity is short compared to the perturbation period 1/f. During the time it takes light to

travel through the very short cavity (typically a few cm or less), there is not enough time for

the temperature to change, and therefore the noise components picked up in successive passes

are fully correlated. A corollary of this property is that the total phase noise is expected to be

the phase noise of a fiber of length L multiplied by the number of passes N through the cavity,

and therefore for a high-finesse cavity it is expected to be significantly larger than the single-

pass phase noise. The PSD should therefore be proportional to the square of the number of

passes through the cavity. The number of passes through an FP cavity is approximately equal

ng/n, where ng is the group index of the light in the cavity and n the refractive index of the

medium between the mirrors. The PSD at low frequencies should therefore be proportional to

(ng/n)2. Since ng depends on the frequency of the light, and is maximum on a resonance, so

should the phase noise. The PSD at low frequencies, proportional to ng2, should be a very

strong function of optical frequency near a resonance. Detailed algebraic manipulations of the

expressions of the phase-noise power and/or PSD at low frequencies published by several

other groups prior to this work show that they do not support these physical predictions [6–8].

However, the predictions published in [1] have been confirmed by others [2,3].

3.1.3 Phase noise in the transmitted signal

 We consider a generic fiber Fabry-Perot interferometer consisting of a medium of length L

and uniform refractive index n placed between two parallel reflectors M1 and M2 with field

reflection coefficients r1 and r2, and transmission coefficients t1 and t2, respectively (Fig. 3.1).

The medium’s single-pass transmission exp(-alossL), where aloss is the medium’s loss

coefficient, is implicitly incorporated in these four coefficients. Without loss of generality, we

assume that the medium is an optical fiber; the results are, however, applicable to any linear

intra-cavity medium. The phase noise in the reflectors is assumed to be negligible, as

supported by [9], compared to the phase noise of any reasonable length of fiber. The general

process used to compute the phase noise of this FP is essentially the same as the text-book

method to calculate the electric field at the two outputs of an FP, namely by adding the fields

Page 48: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

36  

produced after each successive round trip through the cavity [11]. The difference is that the

phase of the fields is now not just the propagation term, but this term plus the phase noise

arising from thermodynamic fluctuations inside the medium, the latter being given by the

aforementioned analytical expression of the single-pass phase noise derived by Wanser [5].

Figure 3.1. Schematic of the computation of the phase noise in the signal transmitted by a Fabry-Perot

interferometer

The phase variation per unit length at a point z along the fiber is φ(z, t) . If this variation is

caused by temperature fluctuations ΔT(z,t), then it is given by [4]:

φ(z, t) = 2πλ

dndT

+ nα!

"#

$

%&ΔT z, t − L − z

v!

"#

$

%&         (3.1)  

 

where dn/dT is the dependence on temperature T of the fiber mode effective index n, α is the

thermal expansion coefficient of the fiber, and v = c/n is the phase velocity of light in the fiber.

Integrating this incremental phase noise along the fiber length yields the phase fluctuations per

pass

Φ(t) = φ(z, t)dz0

L

∫ =2πλ

dndT

+ nα#$%

&'( ΔT (z, t − L − z

v)dz

0

L

∫      

(3.2)  

 

The PSD of the phase noise of the signal that has traveled through the fiber only once

(“single-pass PSD” for short) is the Fourier transform (denoted by F) of the autocorrelation of

Page 49: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

37  

Φ(t) [5]

 

Sf (ω ) =F < Φ(t)Φ(t + τ ) >( )         (3.3)  

 

where τ is the variable of integration in the autocorrelation, and ω is the thermodynamic

fluctuations frequency. Equation 3.3 can be rewritten in terms of the autocorrelation of dφ(z, t)

<Φ(t)Φ(t +τ )>= < φ(z, t − L − zv)dzφ(z ', t − L − z '

v+τ )dz ' >

0

L

∫0

L

∫     (3.4)  

 

As discussed in section 3.1.2, thermodynamic fluctuations along a fiber are typically assumed

to be uncorrelated in space [4], which means that the autocorrelation of φ(z, t) is a delta

function of z - z’ and only a function of the time delay τ of light through the fiber

< φ(z, t)φ(z ', t + τ ) >= Rφ(τ )δ(z − z ')         (3.5)  

In [5], Wanser described the noise statistics with a particular function Rφ(τ), or Rw(τ), which he

omitted to specify but which looks roughly like an exponential symmetric about τ = 0.

Entering this function in Eq. 3.5, then entering this autocorrelation in Eqs. 3.4 and 3.3, yields

the single-pass PSD presented in [5]

Sϕϕ (ω) =F Rw (τ )δ(z− z ')dzdz '0

L

∫0

L

∫#

$%

&

'(=F Rw (τ )( )L = 4πL

λ 2kBT

2

k(dndt+ nα)2 log

kmax2 + (ω

v)2

#

$%

&

'(2

+ωD#

$%

&

'(2

kmin2 + (ω

v)2

#

$%

&

'(2

+ωD#

$%

&

'(2

#

$

%%%%

&

'

((((

(3.6)

where kB is the Boltzmann constant, k is the thermal conductivity, and D the thermal

diffusivity of the fiber material, kmax = 2/W0, where W0 is the 1/e mode field radius, and

kmin = 2.405/af, where af is the fiber cladding radius. The fiber jacket is assumed to have

negligible effect on the thermal phase noise, since little to no energy of the mode propagates in

the jacket.

This expression has been widely used in the literature. It has been shown, however, to be

Page 50: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

38  

incorrect at very low frequencies (typically under ~1 kHz) because it ignores 1/f noise,

dominant at these frequencies [12]. The physical origin of the 1/f noise and consequently its

expression are still a subject of debate [12-16]. However, the physical mechanism that best

matches experimental observations of phase noise in optical fibers at low frequencies is

thermomechanical noise, which is spontaneous length fluctuations caused by mechanical

dissipation, or equivalently random extensions and contractions of the fiber due to internal

friction [12-15]. The PSD of this contribution to phase noise is described analytically by

[12,15]:

STM (ω) =nλ 22kΒTLφ03E0Af

2πω

(3.7)

where E0 is the bulk modulus of the material, Af the cross-sectional area of the fiber including

the polymeric jacket, and φ0 the loss angle that characterizes mechanical dissipation [12,15].

Unlike the main component of phase noise, the 1/f thermomechanical noise does depend on

the jacket thickness because in a first order approximation the fiber is modeled as a glass rod,

and the larger the area of the rod the smaller the mechanical energy dissipation (less internal

friction), which causes the length change and hence the phase change. The total phase noise

PSD of a fiber, Sf(ω), is then the sum of the thermodynamic fluctuations contribution (Eq. 3.6)

and this 1/f thermomechanical contribution (Eq. 3.7):

Sf (ω) = Sϕϕ (ω)+ STM (ω) (3.8)

This sum has been shown to predict correctly the phase noise PSD measured in an optical

fiber down to 20 Hz, the resolution of the instruments [12]. Above a few kHz, and often above

1 kHz [12], the thermomechanical contribution is negligible and the total single-pass PSD

measured experimentally is very well represented by Wanser’s formula (thermodynamic

fluctuations) alone [12]. An example of the spectral shape of this total single-pass PSD will be

given in Section 3.1.4. In any case, the results presented in this chapter do not depend on the

specific shape of the total single-pass PSD.

We are interested in calculating the phase noise in both the signals reflected and transmitted

by the FP. For the transmitted signal, we must calculate the phase noise in the circulating field

Page 51: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

39  

Ec2 at the output reflector M2 by summing all the fields traveling in the cavity (shown in Fig.

3.1), then transmit this signal through this reflector to obtain the transmitted signal and its

phase noise. The transmitted field being simply Et = t2Ec2 (see Fig. 3.1), the transmitted and

intra-cavity fields have exactly the same phase (except for a constant, deterministic phase shift

on transmission). Consequently, the transmitted phase noise PSD SFPt is equal to the intra-

cavity phase noise PSD SFP. The derivation of the phase noise PSD of the circulating field at

reflector M2, and hence of the transmitted field, is presented in the rest of this section and in

sections 3.1.4, and 3.1.5. The more complex case of the reflected field is dealt with in section

3.1.6.

In the Fabry-Perot interferometer of Fig. 3.1, the intra-cavity field Ec2(t) just before mirror M2

can be calculated classically by adding all the fields that have traveled back and forth an odd

number of times between the mirrors. For this derivation, we invert the coordinates in Fig. 3.1,

so that z = 0 is at mirror M2 and z = L is at mirror M1, which produces simpler equations. If E0

is the field incident on the input mirror M1, the first term in this series is

t1 exp −iΦ0( )exp −i φ(z, t − zv)dz

0

L

∫$%&

'() , where Φ0 = 2πnL / λ +ϕ t1 is the single-pass phase through

the fiber plus the phase shift ϕ t1 associated with the transmission through the first mirror, and

the integrated phase term is the total phase noise picked up by the signal as it traveled between

z = 0 and z = L. The time variable t - z/v accounts for the fact that the field picked up the phase

noise at different points and therefore at different times, since the field propagates.

Iterating this process over an infinite number of passes n through the FP yields the intra-cavity

field

Ec2

E0

= t1e−iΦ0 exp[−i φ(z, t − z

v0

L

∫ )dz]{1+ (r1r2e−i2Φ1 )n

n=1

∑ ⋅ exp[−i φ(z, t − 2mL − zv

)+φ(z, t − 2mL + zv

)'()

*+,dz

0

L

∫m=1

n

∑ ]}

Φ1 = 2πnL / λ +ϕr1 +ϕr 2

2

(3.9)

where φr1 and φr2 are the phase shift associated with reflection on the first and the second

mirror, respectively.

Page 52: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

40  

Since the phase noise terms in the last exponential terms are small, the latter can be expanded

to first order in φ(z, t) to get

Ec2

E0

= A{1− i φ(z, t − zv)dz

0

L

∫ − i (r1r2e−i2Φ1 )m

m=1

∑ ⋅ φ(z, t − 2mL − zv

)+φ(z, t − 2mL + zv

)'()

*+,dz

0

L

∫ }      (3.10)  

where A is the usual field enhancement factor provided by an FP

A = t1e−iΦ0

1− r1r2e−2iΦ1

          (3.11)  

The phase variations due to random temperature fluctuations are described by the argument

θ(t) on the right side of Eq. 3.10 without the inclusion of A (whose phase term does not

include any phase noise). The PSD of the phase noise of the intra-cavity field is the Fourier

transform of the autocorrelation of θ(t). The general form of the phase noise PSD is therefore

SFPt =F [< θ (t)θ (t + τ ) >]=F [< Arg{1− i φ(z, t − zv)dz

0

L

∫ − i (r1r2e−i2Φ1 )m

m=1

∑ .

φ(z, t − 2mL − zv

)+φ(z, t − 2mL + zv

)&'(

)*+dz}

0

L

∫ Arg{1− i φ(z ', t − z 'v+ τ )dz '

0

L

∫ − i (r1r2e−i2Φ1 )m

m=1

∑ .

φ(z ', t − 2mL − z 'v

+ τ )0

L

∫ +φ(z ', t − 2mL + z 'v

+ τ )dz '} >]

(3.12)

As a check, since this expression is valid for any values of the reflectivities, in the limit of

r1 = r2 = 0 it should give the single-pass PSD. In this limit, the two infinite series vanish and

Eq. 3.12 reduces to

SFPt =F [Arg{1− i φ(z, t − zv)dz

0

L

∫ } ⋅ Arg{1− i φ(z ', t − z 'v+ τ )dz '

0

L

∫ }]     (3.13)  

Since the phase noise ε is small, in Eq. 3.13 the argument of the two terms of the form (1 – iε)

is -ε, and Eq. 3.13 is simply

Page 53: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

41  

SFPt =F [ φ(z, t − zv)dz

0

L

∫ φ(z ', t − z 'v+ τ )dz '

0

L

∫ ]       (3.14)  

which is by definition the single-pass PSD (see Eqs. 3.3 and 3.4), as expected.

Note that since Eq. 3.12 is the phase noise inside the cavity at the output reflector, it cannot be

used to calculate the PSD of the phase noise of a signal that has traveled twice through the

fiber (as one might be tempted to do by setting r1 = 0 and r2 = 1). To do so, one must calculate

the PSD of the intra-cavity at the input reflector, which we do in Appendix.

Under the most general conditions, Eq. 3.12 does not have a simple closed form, and it must

be evaluated numerically. As described in Chapter 2, we are interested in FBGs strain sensors

interrogated at one of the two optical frequencies on either side of a resonance where the slope

of the transmission (or reflection) spectrum is steepest. Calculating the phase noise at these

interrogation frequencies requires evaluating Eq. 3.12 (or equivalently 3.10) away from

resonance. In other devices, operated on a dark fringe for example, the phase noise must be

evaluated on resonance. These two cases are investigated separately in the following sub-

sections.

3.1.4 Phase noise on resonance

On a resonance, Φ1 = pπ, where p is an integer, and Eq. 3.10 can be greatly simplified. It can

be shown, after some lengthy but straightforward calculation (see Appendix), that in this case

the intra-cavity phase noise PSD can be written as

SFPt (ω)= Sf (ω)[1+ cos 2ωLv

+sinc 2ωLv

!

"#

$

%&. 2r1r21+ (r1r2 )

2 −2r1r2 cos2ωLv

]     (3.15)  

As expected, on resonance the PSD of an FP is equal to the single-pass PSD Sf(ω) times a

scale factor that depends on both the frequency and the mirror reflectivities, and hence on the

finesse. This expression is valid at all frequencies, but again only on resonance.

Page 54: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

42  

To get a physical sense for this scale factor, it is instructive to look at Eq. 3.15 in the limit of

low frequency, when 2ωL/v << π (a condition roughly equivalent to a frequency f = ω/(2π)

much smaller than the reciprocal of the single-pass transit time through the FP). For a fiber FP

sensor made of a silica fiber (n ≈ 1.45), the FP length is typically less than 1 m, and this

condition is met for all frequencies much smaller than ~50 MHz, which covers the frequency

range of the overwhelming majority of sensors. In this frequency range, the three

trigonometric terms in Eq. 3.15 are close to unity, and the PSD becomes independent of

frequency. The reason for this independence is that at low frequency the temperature

essentially does not fluctuate during the very short transit time of the light through the fiber,

and all frequencies produce the same phase noise. Equation 3.15 then becomes

SFPt (ω ) = Sf (ω ) 1− 21− r1r2

"#$

%&'

2

when ω <<πv2L

      (3.16)  

The factor in parentheses is the ratio of the group velocity ng (at resonance) to the mode

effective index n, which is the slowing-down factor η of the FP on resonance (see Eq. 2.23).

Hence

SFPt (ω ) = Sf (ω )ngn

!

"#

$

%&

2

= Sf (ω )η2 when ω <<πv2L    

  (3.17)  

As predicted earlier from basic physical principles (see Section 3.1.2), the phase noise of an

FP, which is proportional to the square root of the PSD, is proportional to the group index,

hence to the finesse (in an FP the group index on resonance is related to the finesse F by

ng(ω0) = 2nF/π for a large finesse, see Εq. 2.24).

To illustrate the properties of phase noise, we simulated an FP made with a silica fiber

(n = 1.45 at 1.55 µm) with a length L = 1 cm, a cladding diameter 2af = 125 µm, reflectivities

|r1|2 = |r2|2 = 0.99 with φr1 = φr2 = 0, a 1/e mode field radius W0 = 5.2 µm, k = 1.37 W/m/K, and

D = 0.86x10-6 m2/s for silica. The thermodynamic component was calculated using Eq. 3.15.

To calculate the 1/f noise component of Sf(ω) (Eq. 3.7), we used a fiber cross-sectional

diameter of 250 µm (Af = 4.91 10-4 cm2) to account for the presence of the 62.5-µm-thick

jacket, and an effective bulk modulus E0 =1.9x1010 Pa for the composite silica-jacket fiber, as

Page 55: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

43  

done in [12,15]. The fiber temperature was taken to be T = 300K. In this wavelength region

the fiber loss is negligible compared to the reflectors’ transmissions, hence the group index is

imposed by the latter. From the values of r1 and r2, the slowing-down factor is η= ng/n = 199

(see Eq. 1.19), corresponding to a group index ng = 289. Figure 3.2 plots the PSD of this fiber

FP used in transmission as a function of the fluctuation frequency f in Hz. The lower red curve

is the total single-pass PSD Sf(ω) given by Eq. 3.8. It decays as 1/f up to ~80 Hz, and above

this frequency exhibits the same dependence predicted by Wanser (dashed red curve) [5],

namely it is constant up to ~10 kHz and it decays as 1/f2 above this value. Thus, even though

Wanser’s formula ignores the thermomechanical contribution, in this particular example it

predicts the phase noise well for any frequencies above ~80 Hz, which covers most sensor

applications. The dashed blue curve is the low-frequency limit of the FP PSD given by Eq.

3.17. It is the same as the single-pass PSD (red lower curve) but multiplied (translated in the

logarithmic space) by η2 ≈ 104.

Figure 3.2. Phase-noise PSD of the signal on resonance transmitted by a 1-cm fiber Fabry-Perot

interferometer (see text for details).

The exact PSD (solid black middle curve), calculated with Eq. 3.15, exhibits three distinct

regions. Up to ~10 MHz, it overlaps perfectly with the low-frequency PSD. Above ~10 MHz,

it decreases rapidly toward the single-pass PSD. At frequencies above ~10 GHz it exhibits

sharp resonances confined in amplitude roughly between 3-dB below the low-frequency limit

and 3-dB below the single-pass PSD.

Page 56: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

44  

The physical reason for the rapid decrease in the second region, followed by the oscillations in

the third region, is successively destructive and constructive interference. The total phase

noise is the sum Zf of the noise accumulated by all the forward waves, which are correlated to

each other, plus the sum Zb of the noise accumulated by the backward propagating signals,

which are also correlated to each other. Because the noise at high frequency accumulated in

the forward and in the backward directions are uncorrelated, the PSD is the sum of the square

of these two terms

 

SFPt (ω p ') = Sf (ω p ') Z f2 + Zb

2( )       (3.18)  

The magnitudes of Zf and Zb depend strongly on how the noisy waves interfere in the FP,

which in turn depends strongly on the noise frequency ω. On resonance and at dc noise

(ω = 0), and assuming the common condition φr1 = φr2 = 0, the optical signal resonates in the

FP at frequencies given by the usual condition 2ω0L/v = ±2mπ, where m is an integer. At noise

frequency ω, the resonance condition becomes 2(ω0+ω)L/v = 2mπ. For certain high enough

noise frequencies ωp, the noisy signal resonates on a higher order of interference m + p, where

p is an integer, i.e., 2(ω0+ωp)L/v = 2(m+p)π. The phase noise is therefore expected to exhibit

resonant frequencies ωp = pπv/L. At these frequencies the correlated noise components

(forward together, and backward together) add up in phase at each consecutive round trip,

leading to a large total noise. These resonant frequencies agree well with the sharp peaks

above ~1010 Hz in Fig. 3.2. For example, for the FP modeled in Fig. 3.2 (L = 1 cm, n = 1.45)

the first resonance (p = 1) is expected at ω1 = πv/L, or ω1/(2π) = 10.3 GHz, in agreement with

Fig. 3.2. Similarly, for certain noise frequencies ωp’ the noisy signal is anti-resonant, i.e.,

when 2(ω0+ωp’)L/v = ±2(m+p)π + π. These anti-resonant frequencies are then given by

ωp’ = (p+1/2)πv/L. At these frequencies the correlated noise components picked up in

successive round trips add up out of phase, leading to a weak noise. This sequence of anti-

resonant frequencies also agrees well with the sequence of minima in Fig. 3.2. For example,

the first anti-resonant frequency (p = 0) is expected at ω1’ = πv/(2L), or ω1’/(2π)=5.17 GHz.

This value is in broad agreement with the first minimum in Fig. 3.2 (6.85 GHz). The

discrepancy arises from the sinc dependence of SFPt and the Sf term, which were ignored in this

simplified argument.

Page 57: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

45  

Each of Zf and Zb is an infinite series of terms diminishing exponentially in amplitude. At an

anti-resonant noise frequency, these terms have opposite phase, i.e.,

 

Z f = Zb = (1− t + t2 − t3 + ...) = 1

1+ t         (3.19)  

where t is the round-trip transmission of the FP cavity (i.e., t = 1 - δ where δ is the round-trip

loss). Since the loss δ assumed in the simulations of Fig. 3.2 is small, the right hand side of

Eq. 3.19 equals 1/2. Inserting this value in Eq. 3.18 yields SFPt (ω p ') = Sf (ω p ') / 2 , which

explains why the minima in the PSD are Sf/2 (3 dB below the red curve Sf(f), see Fig. 3.2). It is

worth mentioning that although the phase noise is clearly very low at these anti-resonant

frequencies, the sensitivity of a sensor should also be minimum, for the same physical reason.

So we expect no net benefit in terms of minimum detectable signal (or signal-to-noise ratio)

when operating at one of these frequencies. A similar argument can be made to explain why

the noise maxima at the resonant frequencies occur at half the low-frequency limit.

In summary, the PSD is never larger than the low-frequency limit. As a consequence, in an FP

with a known finesse F (related to the slowing-down factor by η = 2F/π , see Eq. 2.24), an

upper bound value of the PSD can be quickly evaluated by multiplying the single-pass PSD by

(2F/π)2. At certain (anti-resonant) frequencies the PSD drops below the single-pass PSD. The

phase noise can therefore be drastically reduced compared to its low-frequency limit by

operating a device at one of these frequencies.

3.1.5 Phase noise off resonance

 In the more general case where the FP is probed off resonance, the terms containing Φ0 and

Φ1, in the general expression of the PSD (Eq. 3.12), are no longer equal to unity and Eq. 3.12

cannot be simplified. However, in the limit of low frequencies (2ωL/v << π), it can be shown

that the intra-cavity field (Eq. 3.9) takes the simple form

 

Ec2

E0

=t1 exp −iΦ0

'( )1− r1r2 exp −i2Φ1

'( )         (3.20)  

Page 58: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

46  

where Φ'0 = Φ0 +Φ(L, t) , Φ

'1 = Φ1 +Φ(L, t) and Φ(L, t) is the single-pass phase noise integrated

along the length of the fiber, defined in Eq. 3.2. This equation, valid off and on resonance but

at low frequencies only, states that the expression of the noisy intra-cavity field has exactly the

same form as the expression of the intra-cavity field in the absence of phase noise, provided

the usual phase term Φ0 is replaced by Φ0 plus the integrated phase noise, which makes

physical sense.

If we make the further assumption that the single-pass phase noise is small (i.e., Φ(L, t) << π),

which is very well satisfied for many practical situations, we can approximate the argument of

Eq. 3.20 by taking its Taylor-series expansion around Φ(L, t) at low frequency. Taking the

Fourier transform of the autocorrelation of the noise component of this argument gives the

PSD

 

SFPt (ω ) = Sf (ω )ng (λ)n

!

"#

$

%&

2

when ω <<πv2L

        (3.21)  

This result generalizes the result of Eq. 3.17 (which was derived only on resonance using a

different method) in that it applies at any arbitrary wavelength (on or off resonance), at low

frequency, and when the phase noise is small (it makes no other assumption). As long as we

are only concerned with low frequencies, which again covers almost all practical applications

of FP, the phase noise PSD is simply proportional to ng2(λ), and the phase noise to ng(λ), i.e.,

to the number of passes through the Fabry-Perot. Since this derivation makes no assumption

about the phase noise itself, it is true for each of the two contributions of the single-pass phase

noise individually, namely the thermodynamic noise (Eq. 3.6) and the 1/f noise (Eq. 3.7).

The salient and most important finding is that in a resonator such as an FP cavity, the phase

noise PSD is enhanced by a factor of (ng/n)2. The thermal phase noise, which is proportional to

Φ(L, t) , is enhanced by a factor of ng/n in the presence of slow light (in contrast to √ng claimed

in [17]).

In summary, simple closed-form expressions have been derived for the phase-noise PSD of

the transmitted signal in two different regimes, namely on resonance at any frequency, and off

Page 59: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

47  

resonance at low frequencies. Off resonance, the PSD is very simply the single-pass PSD Sf

multiplied by η2, the square of the slowing-down factor η, which is proportional to the group

index at the optical frequency. The thermal phase noise is highest on resonance, and it

decreases away from resonance in proportion to the group index. On resonance, the PSD is

also proportional to Sf. The proportionality factor is also η2 but only at low frequencies; at high

frequencies it depends on η in a more complex fashion. In this last case there is no simple

closed-form expression for the PSD, and one must resort to numerical simulations.

Figure 3.3. Phase-noise PSD spectrum at the transmission port of a silica-fiber FP with reflectivities

r1 = r2 = √0.99, normalized to the single-pass PSD spectrum. Solid curves are the numerical solutions

for different detuning δΦ0 from resonance.

Figure 3.3 plots the phase-noise PSD spectrum for the same symmetric FP as in Fig. 3.2 but

now calculated for non-zero detuning from a resonance up to high frequencies, in which case

no closed-form solution exists. These plots were calculated numerically using Eq. 3.9 and the

following procedure. The FP cavity length L was divided into p = 101 segments of equal

length. For each segment, a temporal sequence of random phase fluctuations δφ(ti) was

generated at q = 100001 equidistant times ti ranging from t = 0 to t = 100001T, where T is the

reciprocal of the sampling frequency. The sampling frequency was chosen to be a multiple of

v/(2L). The random fluctuations at a given point had the same PSD spectrum as shown in Fig.

3.2 (single-pass curve). These p distributions were independent, since thermal fluctuations are

uncorrelated in space. In order to generate these random sequences of noisy phase with a

specific PSD spectrum, we followed the process described in [18]. The ratio Ec2/E0 was

Page 60: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

48  

calculated for 1001 passes. This is equal to ~5 times the number of passes through the FP,

which is more than enough to reach convergence (increasing the number of passes changed

the outcome by less than 1%). We then calculated the autocorrelation of the phase and its

Fourier transform to obtain the PSD spectrum. This process was repeated many times (~1000),

each time using a different set of p temporal distributions. These PSD spectra were finally

averaged to obtain the final PSD spectrum.

This process was carried out for two different fiber PSDs, namely first assuming that the

single-pass PSD is given by Eq. 3.6, i.e., that it does not include the thermomechanical

contribution (an assumption certainly valid in most cases at frequencies above ~1 kHz, as

illustrated in the example of Figure 3.2 where an ~80-Hz cutoff was predicted), then assuming

that it does include the thermomechanical noise (and is given by Eq. 3.8). As suggested by Eq.

3.15, and as verified with numerical simulations, in both cases when normalized to the single-

pass PSD the transmitted PSD depends on frequency as ωL/v, which explains the

normalizations of each of the two axes in Fig. 3.3. All curves are the numerical solutions

plotted for increasing detuning from resonance δω0, characterized by the round-trip phase

detuning δΦ0 = 2δω0L/v. The top black solid curve is plotted on resonance (δΦ0 = 0); it is the

same as the oscillatory solid curve in Fig. 3.2, except normalized to Sφφ(ω) and plotted versus

ωL/v instead of ω. At low frequency (ωL/v ≤ ~10-3 rad) the numerical simulations confirm the

prediction of Eq. 3.17 that the PSD is proportional to the ratio (ng(λ)/n)2, i.e, it decreases

rapidly with increasing detuning. Above this frequency, in the off-resonance case an

additional resonance occurs at low frequency (e.g., at ωN ≈ 1.65x10-2π rad for δΦ0 = 0.05 rad),

and the high-frequency resonances each split into two.

The origin of this splitting can be better understood by re-plotting Fig. 3.3 with a linear

coordinate (Fig. 3.4). This figure clearly shows that off resonance all the noise resonances

split, by the same amount, and that this amount increases with increasing detuning, up to the

maximum possible detuning of δΦ0 = π. The extra resonance that appears at low frequency

simply arises from the splitting of the resonance at dc. The reason for this splitting is again

interferometric. As stated earlier, at dc noise the optical signal resonates at frequencies given

by 2ω0L/v = ±2mπ. When the signal is detuned from resonance by δω0, it is the sum of the

signal and the noise that resonates, at frequencies given by 2(ω + ω0+ δω0)L/v = ±2mπ.

Combining these last two equations shows that the noise is expected to resonate at frequencies

Page 61: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

49  

ω = ±δω0, i.e., the splitting in the noise peaks is constant and equal to 2δω0. This prediction is

consistent with the splitting in Fig. 3.4. For example, for δΦ0 = 0.05 rad the splitting should be

2δω0 = δΦ0v/L =1.03x109 rad/s, which is precisely the peak splitting predicted numerically in

Fig. 3.4. Note that the split peaks are decreased by a factor of 4 relative to the original peaks

(energy conservation dictates a factor of two reduction, which is squared in a PSD). The split

peaks are not quite symmetric only because of numerical errors.

Figure 3.4. PSD spectrum on resonance (dashed curve) and off resonance for δΦ0 = 0.05 (solid curve)

illustrating the splitting that occurs in the phase-noise resonances when the optical signal is detuned

from a Fabry-Perot resonance.

The normalized FP PSD spectra plotted in Figs. 3.3-3.4 depend on η, but they are independent

of the fiber length as long as the fiber loss has a negligible contribution to the total round-trip

cavity loss (so that η remains the same). In this case, when the length is increased from the

value used to generate the figures, first, the single-pass PSD increases in proportion to L, and

so does the FP PSD, but their ratio remains the same. Second, the resonant frequencies also

increase proportionally to L, so the resonances occur at different frequencies but the

normalized frequencies also remain unchanged. If the length becomes so long that the fiber

loss increases the total round-trip cavity loss noticeably, then one other contribution comes in:

the group index decreases, and as a result the normalized FP PSD spectrum also decreases in

proportion to the ratio η2 (see Fig. 3.2). The normalized FP PSD spectra in Figs. 3.3-3.4 then

need to be recalculated for the new η.

Page 62: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

50  

3.1.6 Phase noise in the reflected signal

 The case of the reflected field (Er) is more complex, because this field is the sum of two terms,

the fraction t1 of the intra-cavity field Ec1 at the input mirror that is transmitted by the input

mirror, and the fraction r1 of the incident field reflected by the input mirror. Compared to the

transmission case, calculating Er therefore requires passing the signal one more time through

the fiber, from M2 to M1. When the FP has a high finesse, this additional pass is expected to

have negligible effect on Ec1, and therefore on the t1Ec1 contribution. When the finesse is low,

it is expected to have a significant impact. On the other hand, the reflected contribution is

noiseless (since it has not traveled through the cavity), so it is expected to reduce the relative

phase noise (relative to the total reflected power) compared to the phase noise in the

transmitted signal.

The derivation of Ec1 mirrors the derivation of the intra-cavity field Ec2 (see Appendix). The

end result is that the phase-noise PSD of the signal reflected by an FP at low frequency and at

any wavelength, on or off resonance, is again given by Eq. 3.21, i.e., it is equal to the single-

pass PSD times the square of the slowing-down factor at the wavelength of interest. For an

arbitrary frequency (including low frequency) and for an asymmetric or lossy FP, on

resonance the FP field reflection RFP is non-zero, and it can be shown, similarly to the

derivation in Appendix for the phase noise in transmission, that the PSD on resonance and at

any frequency is

SFPr (ω ) = 2Sf (ω )r2t1

2

1− r1r2( )RFP

"

#$

%

&'

2 1+ sinc 2ωLv

1+ r1r2( )2− 2r1r2 cos

2ωLv

    (3.22)  

 

The right-hand side of Eq. 3.22 can be recast in terms of the group index of the reflected

signal ng ' (see Eq. 2.24):

 

Page 63: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

51  

SFPr (ω ) = Sf (ω )ng 'n

!

"#

$

%&2 1− r1r2( )

21+ sinc 2ωL

v!

"#

$

%&

2 1+ r1r2( )2− 2r1r2 cos

2ωLv

!

"#

$

%&

        (3.23)  

Note that for a lossless and symmetric FP these last two expressions do not apply on

resonance because the reflected power is zero and therefore the phase noise cannot be defined.

Off resonance the expression at low frequencies is the same as Eq. 3.21 but using the group

index in reflection ( ng ' ) instead of in transmission (ng).

Setting r1 = 0, t1 = 1, and r2 = 1 in Eq. 3.23 gives the PSD of a signal that has traveled twice

through a fiber. Equation 3.23 shows that this PSD depends on frequency as sinc(2ωL/v). This

result contradicts [4], which predicts a cos(2ωL/v) dependence (in our opinion arising from an

error in the Fourier transform calculation in [4]). It is consistent, however, with the sinc

dependence of the PSD of a Sagnac interferometer [10] (also a two-pass device).

Figure 3.5. PSD spectra in reflection (black curves) for the same FP as in Fig. 3.3, calculated for two

values of the detuning from resonance. The PSD spectra in transmission for the same two detunings

(solid red curves, reproduced from Fig. 3.3) are also shown for comparison.

The PSD spectrum in reflection, calculated numerically using the same technique as in the

previous section, is shown in Fig. 3.5 for the same symmetric and lossless FP as in the

previous figures. The phase noise is only plotted off resonance, where there is a non-zero

Page 64: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

52  

reflected signal from the FP. Because of the first reflection off the input mirror M1 is noiseless,

at low frequencies the PSD in reflection is exactly the same as in transmission, since it is

proportional to the group index (and the latter is the same in reflection and transmission for

the present case of a lossless symmetric FP). At high frequencies, however, it is lower than the

phase noise in transmission. We hypothesize that this reduction is a result of the partial

correlation of the various contributions to the phase noise at high frequencies. Halfway

between resonances (i.e., for δΦ0 = π/2, a case not shown in Fig. 3.5), the phase noise is zero,

since the intracavity field is zero and there is therefore no reflection from the cavity.

3.1.7 Applicability to FBGs and FBG-based FPs

 The PSD expressions derived so far have assumed that the two mirrors are so thin that their

contributions to phase noise are negligible, as is the case for dielectric coatings for example.

When one or both of the reflectors are FBGs, there are two differences in the derivation of the

FP phase noise. First, the phases on reflection φr1 and φr2 may be different from what they are

for dielectric (or metallic) coatings. This may shift the resonance frequencies at high

frequencies, but it will not affect the phase noise at a particular optical frequency with respect

to a resonance. The second difference is that each mirror now has a finite length and therefore

adds its own phase noise to the signal. This effect is discussed in the next paragraph.

Calculating the phase noise of an FP made of two FBGs separated by a length of fiber

therefore requires calculating first the phase noise picked up by the signal reflected by an FBG.

This calculation can be done by summing the incremental fields reflected by each of the

periods in the FBG, in much the same way as the transmitted fields were summed in Section

3.1.3. The same remarks apply to the phase noise picked up by a signal with a wavelength in

the vicinity of the slow-light resonances of an FBG [19].

In this last device, the light is concentrated in the middle of the FBG, as discussed later on in

this thesis. At each pass through the FBG, the light therefore travels an effective length that is

shorter than the physical length LFBG of the FBG. As a consequence, an upper bound value of

the phase noise of this device (in reflection or in transmission) can be obtained by equating the

FBG to an FP made with thin reflectors spaced by a length LFBG. For the same reasons, an

upper bound value of the phase noise of the signal reflected or transmitted by an FP made of

two FBGs spaced by a length of fiber can be obtained by equating the device to an FP made

Page 65: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

53  

with thin reflectors spaced by a length equal to the total length of the device, including the

length of the two FBGs.

3.1.8 Phase noise in experimental fiber FP sensors

 To compare the predictions of this study to experimental noise data, we plot in Fig. 3.6 the

predicted dependence on group index of the phase noise in transmission for three highly

sensitive fiber strain sensors reported in three references [6,20,21], and compare these

predictions to the measured noise reported in the same references. In order to calculate the

predicted phase noise we used Wanser’s Formula (Eq. 3.6), with and without the

thermomechanical noise (Eq. 3.7). The difference between the two results was negligible,

showing in particular that the 1/f noise was negligible in these sensors. It is worth mentioning

that when expressed in units of strain, the phase noise scales as the reciprocal of the square

root of the FBG length, since the sensitivity is proportional to thelength, and the phase noise is

proportional to the square root of the length.

The first sensor (the work of our research group) consists of a 2-cm FBG with a large index

modulation, which created sharp slow-light peaks on the edges of its bandgap as explained in

section 2.3.2b [20]. It was probed at the wavelength of steepest slope of one of these

resonances (to maximize its strain sensitivity), and tested for strain sensitivity at 23 kHz. The

measured group index at the peak of this resonance was 58.2. The predicted dependence of the

phase noise on group index for this sensor is shown in Fig. 3.6 as the solid black curve.

Because this curve was generated assuming that the effective length of the FBG was equal to

its physical length, as discussed in the previous section, this curve (slightly) overestimates the

phase noise. As expected from the previous sections, the phase noise increases linearly with

increasing group index. The measured total noise of this sensor (solid circle in Fig. 3.6) is 13.5

times larger from the predicted phase noise (solid curve). This is consistent with the noise

measurements reported in [20], which concluded that the noise was dominated by laser

frequency noise.

The second sensor is an FP-based sensor made with two FBGs spaced by 13 cm, with a

finesse of 110 [6], or a group index of 101.5. This sensor was probed at a wavelength tuned to

the steepest slope of a resonance. The open circle in Fig. 3.6 is the measured noise of this

sensor at 1.5 kHz, calculated by converting the measured minimum detectable strain reported

Page 66: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

54  

in [6] into a phase error. The predicted phase noise for this sensor (dashed red curve) is much

higher than for the previous sensor because its length is longer and the frequency lower. The

measured total noise of this sensor is 9 times higher than the predicted phase noise. Unlike

claimed in [6], the noise of this sensor was not limited by phase noise, as demonstrated in [7].

Figure 3.6. Theoretical predictions of the phase noise dependence on group index for three highly

sensitive strain sensors utilizing FBGs, and experimental noise measured for each of them.

The third highly sensitive experimental sensor is the fiber Fabry-Perot sensor reported in [21].

This sensor is similar to the one reported in [6], with a length of 2 cm and a finesse of 48, and

it was operated at 216 Hz. Once again the experimental noise is larger than the predicted phase

noise, here by a factor of 4.5.

This study shows that in the three most sensitive passive fiber strain sensors reported prior to

this work, the measured noise is a factor of 4.5–13.5 higher than the predicted phase noise.

Thus this gives significant room for further decreasing the minimum detectable signal of these

three sensors before the phase noise limit is reached.

3.2 Conclusions

 In this chapter a rigorous derivation of the PSD spectrum of the phase noise in the signals

Page 67: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

55  

transmitted and reflected by a Fabry-Perot interferometer was reported. For practical reasons,

we endeavored to express this PSD in terms of the group index ng of the light and of the

detuning of the optical signal from a Fabry-Perot resonance. We found, as supported by

physical arguments, that in transmission the PSD at low frequencies (much smaller than the

reciprocal of the single-pass transit time through the FP, or typically under 10 MHz for a

typical cm-long FBG) is equal to the PSD of the phase noise acquired by a signal that has

traveled once through the FP (single-pass PSD Sf(ω)) times the square of the FP’s slowing

down factor, (ng/n)2, where n is the phase index of the light inside the FP. Equivalently, the

phase noise (expressed in radian), which is the square root of the PSD, is equal to the single-

pass phase noise times ng/n. The higher the finesse F (or equivalently the group index) of the

FP, the larger the number of passes light makes through the FP, and the larger the phase noise

it acquires, in proportion to this number of passes, i.e., in proportion to ng or F. At low

frequency, when the signal is on an optical resonance, ng is maximum and so is the PSD.

When the signal is detuned from a resonance, the PSD is described by the same expression,

i.e., it is still proportional to (ng/n)2, except that ng is lower because of the detuning.

Application of this model to three highly sensitive FP strain sensors reported in the literature

shows that in these sensors, the total noise was 4.5–13.5 times higher than their respective

predicted phase noise. The conclusion is that the phase-noise limit had not been reached by

anyone in a passive FBG-based sensor, unlike claimed in the literature [6], before the work

reported in this thesis.

At high frequency (above ~10 MHz for a typical 1-cm FBG), we showed that the PSD

spectrum can also be expressed in a simple closed-form expression in transmission when the

signal is on resonance, and that off resonance the PSD must be calculated numerically.

Numerical simulations showed that in this frequency range, the PSD oscillates between 3-dB

below the low-frequency PSD and 3-dB below the single-pass PSD. This behavior is

attributed to constructive and destructive interference between the signals co- and counter-

propagating in the FP. Equivalent expressions are derived for the phase noise in the signal

reflected by a Fabry-Perot, which is found to have almost the same magnitude at low

frequencies, but to be lower at high frequencies, because of the presence of the noise-free

reflection off the input FP mirror. Other than this detailed analysis of the phase noise spectrum

dependence on detuning, the main value of this study is that it provides simple expressions to

calculate the PSD spectrum of the phase noise in any linear Fabry-Perot interferometer

Page 68: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

56  

(including slow-light fiber Bragg gratings) at low frequencies (which turns out to be the range

of frequency of interest to the overwhelming majority of FP applications), in reflection or in

transmission, from the knowledge of the group index or finesse alone.

This study was the first prong of a four-pronged effort to observe phase noise in a short

passive FBG and measure the smallest MDS in an FBG. Specifically, it enabled us to properly

design the FBGs reported in Chapter 5 to (1) measure thermal phase noise in an FBG and

verify this theory, and (2) achieve the lowest possible MDS using our slow-light FBG sensors.

The second prong was to increase the group index. Since we need to use a short FBG to

measure phase noise, we needed to maintain the same or achieve a higher sensitivity in a very

short fiber. Thus there was an immediate need to increase the group index in an FBG to

increase the sensitivity and/or to “amplify” the phase noise relative to other sources of noise.

The way we achieved this is discussed in the next Chapter 4. The third prong was to improve

our experimental sensor to (1) reduce the laser frequency noise, (2) reduce the environmental

noise, and (3) improve the stability of the experimental setup. The fourth and final prong was

to combine these results and design two FBGs that are compatible with our improved

experimental setup, with high enough sensitivities, and with a high and low phase noise,

respectively, and measure the thermal phase noise in an FBG, as well as the smallest strain

measured ever in an FBG sensor. These last two efforts are discussed in Chapter 5.

References  

[1]     G.   Skolianos,  H.  Wen,   and  M.   J.   F.  Digonnet,   “Thermal   phase  noise   in   Fabry-­‐Pérot  

resonators  and  fiber  Bragg  gratings,”  Physical  Review  A,  89,  033818  (2014).  

[2]   S.   Foster,   and   G.   A.   Cranch   “Comment   on   “Thermal   phase   noise   in   Fabry-­‐Pérot  

resonators  and  fiber  Bragg  gratings”,”  Physical  Review  A,  92,  017801  (2015).  

[3]   G.  Skolianos,  and  M.  J.  F.  Digonnet,  “Reply  to  “Comment  on  ‘Thermal  phase  noise  in  

Fabry-­‐Pérot   resonators   and   fiber  Bragg  gratings,”  Physical  Review  A,  92,  017802  

(2015).  

[4]   K.   H.   Wanser,   “Theory   of   thermal   phase   noise   in   Michelson   and   Sagnac   fiber  

interferometers,”  Proc.  SPIE  2360,  Tenth  International  Conference  on  Optical  Fibre  

Sensors,  584  (1994).  

[5]   K.  H.  Wanser,  “Fundamental  phase  noise  limit  in  optical  fibres  due  to  temperature  

Page 69: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

57  

fluctuations,”  Electronic  Letters,  28,  53  (1992).  

[6]   G.  Gagliardi,  M.  Salza,  S.  Avino,  P.  Ferraro,  and  P.  De  Natale,  “Probing  the  ultimate  

limit  of  fiber-­‐optic  strain  sensing,”  Science,  330,  1081  (2010).  

[7]   G.  A.  Cranch,  and  S.  Foster,  “Comment  on  “Probing  the  ultimate  limit  of  fiber-­‐optic  

strain  sensing,”  Science,  335,  286  (2012).  

[8]   V.   Annovazzi-­‐Lodi,   S.   Donati,   and   S.  Merlo,   “Thermodynamic   phase   noise   in   fiber  

interferometers,”  Optical  and  Quantum  Electronics,  28,  43  (1996).  

[9]   N.  Nakagawa,  E.  K.  Gustafson,  P.  T.  Beyersdorf,  and  M.  M.  Fejer,  “Thermal  noise  in  

half-­‐infinite  mirrors  with   nonuniform   loss:   A   slab   of   excess   loss   in   a   half-­‐infinite  

mirror,”  Physical  Review  D,  65,  102001  (2002).  

[10]   K.   Kråkenes,   and   K.   Blotekjaer,   “Comparison   of   fiber-­‐optic   Sagnac   and   Mach-­‐

Zehnder  interferometers  with  respect  to  thermal  processes  in  the  fiber,”  Journal  of  

Lightwave  Technology,  13,  682  (1995).  

[11]   A.  Siegman,  “Lasers”,  Chapter  11.3,  University  Science  Books  Sausalito,  CA  (1986).  

[12]   R.  E.  Bartolo,  A.  B.  Tveten,  and  A.  Dandridge,  “Thermal  phase  noise  measurements  

in   optical   fiber   interferometers,”   IEEE   Journal   of   Quantum   Electronics,   48,   720  

(2012).  

[13]   S.  Foster,  “Low-­‐frequency  thermal  noise  in  optical  fiber  cavities,”  Physical  Review  A  

86,  043801  (2012).  

[14]   L.   Duan,   “General   treatment   of   the   thermal   noises   in   optical   fibers,”   Physical  

Review  A,  86,  023817  (2012).  

[15]   L.  Duan,   “Intrinsic   thermal   noise   of   optical   fibres   due   to  mechanical   dissipation,”  

Electronic  Letters,  46,  1515  (2010).  

[16]   S.  Foster,  A.  Tikhomirov,  and  M.  Milnes,  “Fundamental  thermal  noise  in  distributed  

feedback  fiber  lasers,”  IEEE  Journal  of  Quantum  Electronics,  43,  378  (2007).  

[17]   G.  Gagliardi,  M.  Salza,  S.  Avino,  P.  Ferraro,  and  P.  De  Natale,  “Response  to  Comment  

on   “Probing   the   ultimate   limit   of   fiber-­‐optic   strain   sensing,”   Science,   335,   286  

(2012).  

[18]   W.  H.   Tranter,   K.   S.   Shanmugan,   T.   S.   Rappaport,   and  K.   L.   Kosbar,   “Principles   of  

communication   systems   simulation   with   wireless   applications,”   Prentice   Hall,  

Upper  Saddle  River,  NJ,  ,  Chap.  7,  278-­‐282  (2003).  

[19]   H.   Wen,   G.   Skolianos,   S.   Fan,   and   M.   J.   F.   Digonnet,   “Slow   light   in   fiber   Bragg  

Page 70: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

58  

gratings,”  Proc.  SPIE  7949,  Advances  in  Slow  and  Fast  Light  IV,  79490E  (2011).  

[20]   H.  Wen,  G.  Skolianos,  S.  Fan,  M.  Bernier,  R.  Vallée,  and  M.  J.  F.  Digonnet,  “Slow-­‐light  

fiber-­‐Bragg-­‐grating  strain  sensor  with  a  280-­‐femtostrain/√  Hz  resolution,”  Journal  

of  Lightwave  Technology,  31,  1804  (2013).  

[21]   J.   H.   Chow,   D.   E.   McClelland,   and   M.   B.   Gray,   “Demonstration   of   a   passive  

subpicostrain  fiber  strain  sensor,”  Optics  Letters,  30,1923  (2005).  

 

Page 71: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

59  

Chapter 4: Improving the performance of slow-light FBGs

In this chapter we discuss the practical improvements that were made in the slow-light FBGs

used as sensors in order to achieve even lower MDSs than our group’s previous world record,

and observe and measure quantitatively the thermal phase noise in an FBG sensor for the first

time. To this end, we needed FBGs with stronger resonances and achieve higher sensitivities

and lower noise, mainly lower laser frequency noise and detector noise. As we discussed in

Chapter 2, to obtain strong resonances the gratings needed to have (1) a weak internal loss (as

in all resonators), (2) as large a Δnαc as possible, and (3) a suitably apodized profile. In this

chapter we discuss how proper apodization leads to an increase in group delay. Next, we show

that to satisfy these three conditions it is preferable to use femtosecond gratings, which

produce high index modulations and a much lower loss (for a given index modulation) than

traditional UV-written gratings [1]. It is also beneficial to write the gratings in a deuterium-

loaded fiber [2], a technique that yields much higher photosensitivity and index modulations

[3]. Furthermore, the grating should be annealed to reduce the loss introduced by the writing

process. Finally, we describe in detail the process used to fabricate the gratings tested in this

thesis and how we modeled the index-modulation and loss profiles of the FBGs written with

this technique.

4.1 Apodized FBGs

As discussed in Chapter 2, high group delays can be achieved in a uniform grating. The main

two disadvantages of a uniform grating are first that the slow-light peaks are outside the

bandgap, which leads to a lower reflectivity, and second that the whole FBG is reflective at

the same wavelengths, resulting in a higher cavity loss.

A solution to both of these problems is to shift the Bragg wavelength in the middle of the

FBG, namely to use an FBG with an apodized index profile (Fig. 4.1a). In such a grating the

average index modulation varies along the grating. The Bragg condition is then z dependent:

λB (z) = 2(n0 +Δndc (z))Λ (4.1)

Page 72: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

60  

where n0 is the effective index of the fiber mode before the grating was written, Λ is again the

grating period, z is the position along the grating, and Δndc(z) is the average index modulation

at z. Since Δndc depends on position, so does the Bragg wavelength, according to the solid

green curve in Fig. 4.1b, which is linearly related to the Δndc(z) curve in Fig. 4.1a. For a given

Bragg wavelength, for example λB1, there are two positions, z1 and z’1, around which the FBG

reflects at λB1. The FBG therefore supports an infinite number of FPs, with mirror spacings

varying from zero to some maximum value equal to the physical length of the FBG, and

centered on a Bragg wavelength that ranges from the value at the top of the green curve to the

value at the bottom of the curve in Fig. 4.1b.

Figure 4.1. Creating strong slow-light resonances in an apodized FBG. a) ac and dc index-modulation

profiles of an apodized fiber Bragg grating. b) Dependence of the Bragg wavelength on position along

the grating. c) Effective mirrors in the apodized FBG forming equivalent FPs. d) The transmission

slow-light resonances formed as a result of these multiple equivalent FPs.

In this continuum of Bragg wavelength pairs, only for some pairs (shown as i = 1 to 4 in Fig.

4.1b) is the optical path length at λBi between zi and z’i and back to zi a multiple of 2π. For

these pairs, the transmitted fields add in phase to produce a resonance. The FBG then behaves

like the superposition of these few resonant FPs (four are represented in the example of Fig.

Page 73: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

61  

4.1c). The mirror pair closest to the center of the FBG (z1 and z1’ in Fig. 4.1c) has the largest

Δnac of all mirrors, and therefore the highest reflectivity of all phase-matched pairs, and hence

this FP has the resonance with the narrowest linewidth (see Fig. 4.1d). This mirror pair also

has the shortest spacing; therefore, it supports the fundamental mode (the electric field

distribution along the FBG exhibits a single lobe). The other phase-matched pairs have a

reflectivity that decreases with increasing distance from the center of the FBG, since Δnac is

decreasing, and they support a broader resonance (see Fig. 4.1d). Since they have a larger

mirror spacing, they support a higher order mode (i.e., their electric field distributions have

multiple lobes). A strong apodized FBG therefore exhibits a series of resonances narrower and

generally stronger toward the center of the grating. One advantage compared to uniform FBGs

is that between zi and z’i the Bragg wavelength is not equal to λBi (see green curve in Fig. 4.1b)

and the grating is essentially transparent at λBi, which leads to much stronger slow-light

resonances than if the index-modulation profile were uniform. For this to be true, the rate at

which the Bragg wavelength changes with position should be larger than the rate at which the

size of the bandgap increases with position. Thus, the rate of change of Δndc—proportional to

Bragg wavelength—with position must be larger than the rate of change of Δnac—proportional

to the bandwidth of the bandgap—with position. If this condition is not satisfied, the fiber

between zi and z’i will still act as a reflector, and therefore an effective FP cannot be formed

and the FBG does not slow down light as effectively as it can.

For this kind of FBGs, and in general for an FBG with arbitrary index and loss profiles, there

is no analytical solution to calculate the field reflection and transmission coefficients, unlike in

the uniform case discussed in Chapter 2. In order to model these FBGs and to predict their

transmission and group index/delay spectra we developed a Matlab program. This application

numerically calculates these spectra based on the piecewise method described in [4]. As an

input it uses the Δnac, the Δndc and the loss, all of which are functions of the position in the

fiber. We derived these functions for the particular FBG that was modeled based on their

fabrication technique, described below in section 4.2. This derivation is explained thoroughly

in section 4.3. After the three desired profiles were obtained, we divided the FBG along its

length into a small number N of uniform FBGs of equal length δL, in our case approximately

100. Each uniform FBG has the Δnac, Δndc and loss of the original FBG in its position. The

number of the uniform FBGs must be chosen so as to fulfill two requirements: (1) each

uniform FBG must include several periods of the grating (L>>Λ) in order for the analytical

Page 74: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

62  

solution for the uniform FBG (Eq. 2.3) to be valid, and (2) the Δnac, Δndc and loss of the

original FBG must not change significantly between the beginning and the end of any of the N

uniform FBGs, otherwise this piecewise approximation breaks down. We can then calculate

the transfer matrix of each uniform FBG (Eq. 2.3), which connects the forward (denoted with

“+” superscript) and backward (denoted with “-” superscript) electric fields at its output and

input at one particular wavelength λ. The output of the ith uniform FBG is the input of the

(i+1)th uniform FBG. Using the matrix in Eq. 2.3 we can then calculate the input electric field

of the ith uniform FBG from the input of the (i+1)th uniform FBG:

Ei+

Ei−

"

#$$

%

&''=

cosh κ i2 −σ i

2 ⋅δL( )-i σ i2

κ i2 −σ i

2sinh κ i

2 −σ i2 ⋅δL( ) -i κ i

2

κ i2 −σ i

2sinh κ i

2 −σ i2δL( )

i κ i2

κ i2 −σ i

2sinh κ i

2 −σ i2δL( ) cosh κ i

2 −σ i2 ⋅δL( )+ i σ i

2

κ i2 −σ i

2sinh κ i

2 −σ i2 ⋅δL( )

"

#

$$$$$$

%

&

''''''

Ei+1+

Ei+1−

"

#$$

%

&''

= AiEi+1+

Ei+1−

"

#$$

%

&''

(4.2)

where κι and σι are the ac and dc coupling coefficients, respectively, for the ith uniform FBG

given by:

σ i = 2π n0 +Δndci( ) 1

λ−1λB

#

$%

&

'(− i

γ i

2 (4.3a)

κ i = πΔnac

i

λ (4.3b)

where γ is the power loss coefficient that characterizes propagation in the grating.

Afterwards, by multiplying the transfer matrices for all the N segments, a single matrix C is

obtained that relates the input and output electric fields of the whole FBG:

Ein+

Ein−

"

#$$

%

&''=E1+

E1−

"

#$$

%

&''= A1A2!AN

Eout+

Eout−

"

#$$

%

&''=C

Eout+

Eout−

"

#$$

%

&'' (4.4)

From Eq. 4.4 assuming no backward propagating ( Eout− =0) field is injected into the FBG (as is

the case in our sensing and other applications) the field transmission and reflection coefficient

Page 75: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

63  

can be easily calculated by evaluating C numerically. Then it is trivial to calculate the group

delay and index spectra, as described in Chapter 2 (see Eqs. 2.8).

As an example of this process, and to illustrate the benefits of apodization by the same token,

we show in Figs. 4.2 and 4.3 the simulated group-index and transmission spectra of two

FBGs, one with a uniform index-modulation profile and the other with an apodized profile.

The gratings have the same length L = 2 cm, a half peak-to-peak ac index modulation

Δnac ≈ 10-3, a period Λ = 532.98 nm, and a uniform loss with a power loss coefficient

γ = 0.4 m-1, close to the loss measured in FBGs of this strength [5].

Figure 4.2a shows the index profile of the uniform FBG, and Fig. 4.2b its simulated spectra.

The slow-light resonances occur just outside the bandgap, as described in Chapter 2. The

strongest resonance is the one closest to the band edge (see insets in Fig. 4.2b). Its maximum

group index is 146.9. In a uniform FBG, Δnac = Δndc (see Fig. 4.2a), and as discussed in

Chapter 2, the effective FP has a relatively low finesse, which explains this relatively low

group index.

Figure 4.2. a) Index profile of a uniform FBG (period Λ not to scale) and b) simulated transmission and

group index spectra of this FBG.

Figure 4.3a shows the profile of the second FBG, a Gaussian-apodized grating with

Δnac = Δndc = 10-3 and a Gaussian’s full width at half maximum (FWHM) equal to twice the

Page 76: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

64  

FBG length. The loss profile was assumed to be uniform, with the same power loss coefficient

of 0.4 m-1. Both the transmission and the group-index spectra (Fig. 4.3b) exhibit sharp

resonances on the short-wavelength side of the bandgap. The resonances in the group index

and in the transmission spectra occur at the same wavelengths, as expected (this is true for any

index-modulation profile). The strongest slow-light resonance is the one furthest inside the

bandgap; its peak group index is as large as 361.5, i.e., ~2.5 times slower than in the uniform

grating of Fig. 4.2. When the power loss coefficient is reduced from 0.4 m-1 to 0.12 m-1, the

peak group index increases to 1204. These simulations illustrate the importance of apodization

to obtain slow-light resonances with large group indices, and the critical importance of

reducing the loss (a requirement that applies to all gratings, apodized or not).

Figure 4.3. a) Index profile of a Gaussian-apodized FBG (period Λ not to scale), and b) simulated

transmission and group index spectra of this FBG.

4.2 Realization of improved FBGs

It has been discussed and explained that producing higher group indices and larger

transmission leading to high sensitivity in an FBG generally requires (1) high index

modulation, (2) strong apodization (e.g., Gaussian rather than uniform), and (3) reducing the

internal loss. In this section we discuss how we met these goals, namely by using femtosecond

FBGs, fabricating them in deuterium-loaded fibers, and annealing the gratings.

Page 77: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

65  

4.2.1 Fabrication of slow-light FBGs with femtosecond lasers in deuterium-loaded fibers

The FBGs used in this work were fabricated at Université Laval using a femtosecond laser and

a conventional phase-mask technique. He Wen, a former graduate student in our research

group, started the study of slow light in FBGs. After characterizing several FBGs written by

different methods and studying the literature she concluded that femtosecond-written FBGs

have a much lower loss for the same index of modulation than gratings written with other

techniques, in particular using UV lasers [6]. This property is very useful to achieve strong

resonances. In Fig. 4.4 the summary of this study is plotted in the form of the loss versus Δnac.

The data combine values found in the literature as well as He Wen’s measured values (filled

red shapes) and a new representative data point for a femtosecond FBG written in a

deuterium-loaded fiber that we characterized in this thesis. It can be seen first that FBGs

written with a femtosecond laser clearly produce lower loss for the same Δnac, by an

approximate factor of 2–5. However, a larger sample is needed to establish a more precise

quantitative comparison, and further studies to understand the origin of this lower loss.

Second, there is a trend, especially with the conventional writing methods, that as Δnac

increases the loss increases too. This is expected because the same defects that induce the

increase in the fiber refractive index also introduce loss. These defects are introduced when

light with a high intensity is concentrated on a specific point of the fiber.

Figure 4.4. Power loss coefficients versus ac index modulation for different writing techniques.

Page 78: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

66  

In general it is believed that during fabrication of an FBG using light two main

photosensitivity  mechanisms are responsible for the change in refractive index of the fiber [7],

namely color centers and glass densification. In the first mechanism light breaks specific

bonds (also known as precursors) in some of the glass-matrix molecules, which creates color

centers, namely species that exhibit one or more strong absorption peaks, usually in the ultra-

violet region [8]. This induced absorption forces a change in the refractive index of the

material around 1.55 µm through the Kramers-Kronig relationship [9]. This mechanism, which

is the only one present in conventional non-femtosecond FBGs, is directly connected with the

internal loss of the FBG, thus the higher the index change the higher the loss, which is in

agreement with the trend in Fig. 4.4. Furthermore, this mechanism needs a dopant—it is not

present in pure silica fibers—to create bonds that can be broken. In conventional

telecommunication fibers the core is usually doped with germanium to increase its refractive

index relative to the cladding. This dopant provides the necessary bonds (for example, Ge-Si

bonds) [9] that change the refractive index of the fiber when they are broken. Because of this

need for dopants, this mechanism can be greatly enhanced by loading the fiber with hydrogen

or deuterium [10–12]. The downside of loading the fiber with hydrogen is that hydrogen

bonds with oxygen in the glass matrix to form hydroxyl radicals (OH-), which absorb strongly

around 1.4 µm and introduce undesirable loss at 1.55 µm where the FBGs operate [13]. In

contrast, when using deuterium instead of hydrogen, OD- radicals are formed, but they absorb

at a longer wavelength (~1.9 µm [12]) and have little effect on the FBG loss around 1.5 µm.

Figure 4.4 illustrates, as has been known for a long time, that doping the fiber with H2 or D2

produces much larger index modulations. It also shows the newer result that for the same

index modulation, deuterium-loaded FBGs have a significantly lower internal loss when

fabricated with a femtosecond laser than with a UV laser, by a factor of 4 to 5, which as we

shall see has a major benefit for slow-light generation.

The second mechanism responsible for the formation of a grating, glass densification, is

present only when the FBGs are written with high-intensity light, namely with a femtosecond

laser. In this mechanism the high incident intensity causes compaction to the glass, and hence

changes its density. This density change causes the refractive index to change. In femtosecond

FBGs, the index modulation is believed to arise in general from a combination of these two

mechanisms, although further studies are needed to determine their exact contributions and the

dependence of these contributions on other fabrication parameters such as the pulse width,

Page 79: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

67  

pulse energy, exposure time, and glass composition. At this time the physical reasons why

compaction introduces a lower loss than color-center formation have not been fully elucidated.

Figure 4.5. Setup and exposure conditions used to write the deuterium-loaded fiber Bragg gratings

using a femtosecond laser. (Courtesy Martin Bernier).

The FBGs reported in this thesis were realized thanks to Professor Martin Bernier at

Université Laval in Québec using the method as follows. They were written through the jacket

in a single-mode silica fiber from OFS (SMF-28 compatible germanosilicate fiber BF04446)

with a polyimide jacket. The fiber has a core/cladding/jacket diameter of 8.4/125/155 microns,

and a numerical aperture of 0.11. Polyimide was used because during subsequent annealing of

the grating, it can withstand higher temperatures than conventional jacket materials, up to

~400°C for short periods of time, and the jacket did not have to be removed for post-

fabrication annealing. By preserving the jacket during the entire grating fabrication process,

the mechanical strength of the FBG was not compromised [14].

To enhance the germanium-doped core photosensitivity, as discussed above, the fiber was first

loaded with deuterium in a pressurized chamber at 2000 psi and room temperature for 14 days.

The FBGs were written using the method described in [14] and illustrated in Fig. 4.5. It uses a

femtosecond near-IR laser for efficiently writing through the jacket, combined with the

scanning phase-mask technique [15] to extend and control the FBG length. The laser was an

806-nm Ti:sapphire laser (Coherent Legend-HE) producing 34-fs pulses at a repetition rate of

1 kHz. The laser beam was focused through an acylindrical lens with an 8-mm focal length,

and sent through a phase mask placed about 150 µm in front of the fiber. The pulse energy

measured before the lens was adjusted to a typical value of 75 µJ. This value was selected to

be (1) sufficiently higher than the type-I writing threshold, estimated to ~40 µJ, to benefit

Page 80: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

68  

from a significant writing speed; and (2) sufficiently lower than the Type-II damage threshold,

estimated to ~150 µJ, to ensure that the FBG is free of void-like defects, which would have

significantly increased photo-induced losses [16]. Under such exposure conditions in

deuterium-loaded fibers, the FBG formation is a result of a combination of color-centers and

glass densification, as discussed above [17]. The laser induces a narrow channel of refractive

index change about 1-µm wide that propagates along the focusing axis. Since this channel is

significantly narrower than the ~10-µm field diameter of the fiber core mode, in order to

maximize the effective index modulation the beam was scanned over the core’s cross section

during the writing process. This was achieved by mounting the lens on a piezoelectric actuator

(not shown in Fig. 4.5) to sweep the focused beam around the fiber core over ±10 µm at a

frequency of 1 Hz. The change in refractive index remains non-uniform across the cross-

section of the fiber, and therefore birefringence is induced [5,18,19]. The phase mask was

fabricated in-house at Université Laval using holographic lithography to obtain a uniform

pitch of 1070 nm and produce a first-order Bragg resonance around 1550 nm. The Gaussian

beam incident on the lens had an FWHM of 5 mm, which defines the exposure length along

the non-focusing axis of the lens. To produce gratings of various lengths, both the incident

beam and the lens were mounted on an air-bearing linear motion stage (not shown in Fig. 4.5)

and scanned along the fiber’s main axis while both the mask and the fiber were kept stationary

(to avoid vibrations, which would have reduced the fringe visibility). The scanning length was

adjusted up to 20 mm. The scanning speed was adjusted to control the total exposure fluence

and consequently the index modulation, typically from 4 mm/min down to a minimum of

1 mm/min to reach saturation and maximize the index modulation to ~3.5x10-3. The FBG

samples were left at room temperature for about two weeks following their fabrication to

make sure that most of the deuterium had out-gassed and the material properties stabilized

before measurements and annealing.

It has been observed that in femtosecond FBGs, for both mechanisms the index change

evolves with laser fluence, namely the incident power times the exposure time, in the same

fashion. Specifically, below a certain threshold fluence the glass suffers no change in

refractive index [17]. Above this threshold the change in the refractive index grows linearly

with incident fluence. Finally for high fluences the index change saturates, namely any further

increase in the fluence has no effect on the index (presumably because all the color-centers

Page 81: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

69  

precursors have been converted into color centers and because the density of the fiber cannot

change anymore). In section 4.3 we explain how we used a sigmoid function to model this

behavior in an abstract way and lump both mechanisms together. A more detailed model,

based on basic physical principles, is being developed currently in our research laboratory by

Arushi Arora.

4.2.2 FBG annealing

To further reduce the loss and improve the sensitivity of our FBGs they were annealed by

heating them up to 400°C. Annealing is believed to mitigate the loss via at least two

mechanisms, namely by reducing the concentration of shallow defects and the local stresses

introduced by exposure to intense light during fabrication [20]. Annealing reduces the internal

loss of the grating, but it also reduces its index modulation. Our experience shows that the rate

at which the index modulation is reduced is lower than the rate of loss reduction, so that

annealing generally produces stronger and more transmissive slow-light resonances, if it is

done optimally, meaning up to the right temperature for the right amount of time.

The annealing process gave us a post-fabrication knob to empirically optimize our FBGs for

each particular application (i.e., sensitivity, group delay, etc.), by controlling the loss and the

reflectivity of our effective FPs. We were able to find an optimal point, where the pair of loss

and index modulation maximizes a particular figure of merit, by annealing the FBGs in small

temperature steps (down to 10 °C) for 30 minutes each. At the point where our figure of merit

reached its maximum, as determined by measurements, we stopped the annealing process.

This method enabled us to optimize our FBGs with a controllable method, something that was

not possible to do during the fabrication since it is not accurately modeled yet. Thus we cannot

predict the exact impact of the writing conditions on the loss and the index change, i.e.,

because of alignment issues in the fabrication setup, etc. In the future with the new model that

is being developed from Arushi Arora we expect to overcome this problem and have better

control over the loss and the index change during the fabrication process itself.

Figure 4.6 shows an example of the evolution with annealing temperature of the peak

transmission T0 and peak group delay τg,max of the first seven slow-light resonances of a

particular FBG, and their product. The grating was 12.5 mm long and was written with the

Page 82: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

70  

method mentioned in section 4.2.1. It was annealed at a particular temperature in a fiber oven

for 30 minutes. After the oven was switched off the FBG was taken out of it quickly (~1 s)

and rapidly cooled down to room temperature, at which point the transmission and group

delay spectra were re-measured. As the FBG was annealed to higher and higher temperatures,

the transmission (see Fig 4.6a) and the T0τg,max (see Fig 4.6c) product increased monotonically,

whereas the group delay reached a maximum then decreased (see Fig 4.6b). These trends are

representative of the behavior of a large number of femtosecond FBGs that were annealed in

the course of this work. The transmission increases monotonically because both effects of the

annealing process (reducing both the loss and the index modulation) lead to an increase in the

transmission. In contrast, these two effects have opposite impacts on the group delay. When

the loss is reduced the group delay increases, but when the index modulation (and hence the

reflectivity of the effective mirrors) is reduced, the group delay decreases. At one specific pair

of loss and index modulation, the group delay reaches a maximum similar to a normal FP in

which the group delay reaches its maximum when it is matched, namely the transmission of

the mirrors is equal to the round-trip loss of the FP [21]. It is worth mentioning that the closer

the peak is to the middle of the bandgap the higher the annealing temperature where the group

delay reaches its maximum. Also, as we annealed this FBG new resonances arose inside the

bandgap that we were not able to observe before because of their very low transmitted power.

Annealing reduced the loss sufficiently that these peaks acquired a sufficiently high

transmission to be detected. Figure 4.6c is informative when designing a strain sensor, whose

sensitivity is proportional to the T0τg,max product. It shows in particular up to what maximum

temperature the grating must be annealed in order for the best resonance (the one with the

highest product) will reach its highest sensitivity. In this particular grating it was resonance #7

after annealing it up to (~230 °C).

To illustrate the significance of these three improvements, in an apodized femtosecond-written

FBG 2 cm in length we were able to achieve a record group delay of 20 ns [18], 4 times larger

than our previous record of 5.1 ns reported in [5]. Then by introducing deuterium in the

fabrication process and annealing the FBGs with a goal to achieve the highest possible group

delay, we were able to double the group delay up to 42 ns in a 12.5-cm long FBG [22]. This

shows that with these combined improvements we were able to have an eight-fold

improvement in group delay. This value is the current world record for a group delay in an

optical fiber.

Page 83: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

71  

Figure 4.6. Measured evolution with annealing temperature of (a) the transmission, (b) the group

delay, and (c) their product, for the slow-light resonances of a particular FBG.

4.3 Modeling the index profile of FBGs written with a femtosecond laser

This section discusses how we modeled the index modulation of the FBGs written with the

technique described in the previous section.

As discussed in section 4.2 the index change is related to the laser fluence. If the index change

induced by the femtosecond laser was proportional to the laser fluence, the ac index

modulation of the FBG would have the same profile as the convolution f(z) of the focused

laser intensity profile (a Gaussian with a FWHM W ) and a rectangular function (to account

for the scanning of the laser beam over a length L ):

f (z) = exp(−4 ln(2) z2

W 2 )∗Rect(z / L) (4.5)

where z is the longitudinal axis of the fiber, and Rect(z/L) = 1 for 0 ≤ z ≤ L and 0 elsewhere.

This convolution f(z) is plotted (not to scale) as the dashed curve in Fig. 4.7a for L = 10 mm

and W = 5 mm.

However, as discussed in section 4.2.1 the index change is not proportional to the fluence, and

two effects reshape this profile. First, there is a threshold intensity below which the fiber

refractive index is virtually unmodified [17]; this effect shortens the profile and gives it

sharper edges. Second, the index change saturates at high fluence, which flattens the index-

modulation profile in its central region. Since the exact dependence of Δnac on the fluence is

not known, these two physical effects were incorporated in the model of the index-modulation

profile Δnac(z) by processing the convolution f(z) through a sigmoid function [23], which

yielded the following trial index-modulation profile:

Page 84: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

72  

Δnac (z) = Δn01

1+ exp(−af (z)+ b)−

11+ exp(b)

#

$%

&

'( (4.6)

Figure 4.7. Index modulation profile example of a fiber Bragg grating written with a femtosecond laser;

(a) convoluted profile, half-peak-to-peak ac index profile, and dc index profile; (b) upper and lower

envelopes of the index modulation.

The parameters a and b control the threshold and the saturation applied to the original profile.

The second term in the bracket ensures that the ac index modulation is zero outside of the

FBG. The factor Δn0 is a constant scaling parameter. The dc index profile was assumed to

have the form:

Δndc (z) =αΔnac (z) (4.7)

where α is a constant.

By definition of the ac and dc index-modulation profiles, the index modulation of the grating

is confined between an upper envelope Δn+(z) = Δndc(z) + Δnac(z) and a lower envelope

Δn–(z) = Δndc(z) – Δnac(z), as illustrated in Fig. 4.7b. Note that, as expected from the symmetry

of the laser scanning during fabrication, these profiles are symmetric about the middle of the

grating (Δnac(z) = Δnac(L - z)). Also, the lower envelope Δn–(z) is negative along some of the

grating, as expected for a grating induced by a nonlinear process [24]. The index modulation

in our gratings is at least partly induced by a multi-photon process [25], and as such it is a

Page 85: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

73  

nonlinear function of the laser intensity. The index modulation therefore has frequency

components not only at the fundamental frequency, but also at harmonics of this frequency.

The fitted ac index profile contains only the fundamental frequency (since harmonics do not

affect the FBG properties around 1550 nm), which explains the negative lower envelope. In all

simulations the FBG loss was assumed to have the same profile shape as Δnac(z), on the basis

that the loss increases with index modulation, as discussed in relation to Fig. 4.4.

Figure 4.8. a) Measured and simulated transmission spectra of the FBG. b) Magnified portion of slow-

light peak region of the transmission spectra shown in a). c) Measured and simulated group-delay

spectra in the same wavelength range as in b).

The general profiles of Eqs. 4.6 and 4.7 were used as an input to our code (described in detail

in section 4.1) to predict the grating’s transmission and group-index spectra. Specifically, to

infer the index-modulation profile of a particular fabricated FBG, its group delay and

transmission spectra were first measured. The profiles calculated using the index-modulation

profiles of Eqs. 4.6 and 4.7 were then visually best fitted to these measured spectra by

Page 86: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

74  

adjusting the values of the five free parameters in the model, namely Δn0, a, and b in Eq. 4.6,

α in Eq. 4.7, and the maximum loss coefficient γmax (at the center of the grating) in Eq. 4.3a.

As an example of the performance of our modeling, in Fig. 4.8 we plot the measured spectra

(solid line) and the predicted spectra (dashed line) using our model for the FBG that gave us

the record-high group delay of 42 ns. The measured spectra were measured using the

experimental shown in Fig. 4.9. Light from a fiber-coupled tunable 1550-nm laser was

amplitude modulated at a frequency fm of 1 MHz with the help of a function generator. The

modulated light was passed through a fiber polarization controller, then coupled into the FBG.

The output signal from the FBG was sent through a 3-dB fiber splitter, which was connected

at one output end to an optical power meter to measure the output power, and at the other

output end to a detector followed by a lock-in amplifier to measure the phase of the output

signal. The laser was first tuned to a wavelength far away from the FBG bandgap (1555 nm),

which provided a point of reference where the output transmission (unaffected from the FBG)

was high and wavelength independent, and set to unity, and the output phase (unaffected by

the FBG dispersion) was set to zero. Both outputs were acquired using a data acquisition

system. The laser wavelength was then scanned and the measurements repeated at individual

wavelengths across the slow-light region. The measured spectrum of the phase delay φ

provided the spectrum of the group delay τg, related to the phase delay by τg = φ/(2πfm).

Figure 4.9. Experimental setup used to measure the group delay and transmission of the FBG (see text

for details).

The transmission spectrum exhibits a steep-walled reflection bandgap, with several slow-light

resonances on its short-wavelength edge, and none on the long-wavelength side (solid green

curve in Fig. 4.8a). Figure 4.8b zooms in on the slow-light transmission resonances on the

Page 87: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

75  

short-wavelength edge. Toward the center of bandgap (longer wavelengths), the sharpness of

the resonances increases (solid green curve). Their peak transmissions of the resonances

decrease toward the Bragg wavelength (longer wavelengths). The slow-light resonances occur

in pairs. Each peak in a pair corresponds to a resonance of one of the FBG’s two

eigenpolarizations. This resonance splitting occurs as a result of birefringence that develops in

the FBG during writing, as described in section 4.2.1. The polarization controller (see Fig. 4.9)

was adjusted prior to each spectrum measurement to launch as much of the light as possible in

one of the eigenpolarizations of the FBGs, so as to minimize these spurious peaks, but the

latter could not all be eliminated. Figure 4.8c shows the measured group delay spectrum of the

same grating. The group delay generally increases for resonances with longer wavelengths,

then decreases. The slowest resonance exhibited the aforementioned record group delay of

42 ns.

The index profiles Δnac(z) and Δndc(z) used for the fitting are shown in Fig. 4.7. They gave the

best fits to the measured spectra. These profiles were calculated using the following values:

Δn0 = 2.4x10-3, a = 0.135, b = 1.6, and α = 0.667 in Eq. 4.6, and a very low peak power loss

coefficient of 0.19 m-1 (a single-pass loss as low as 0.010 dB) in Eq. 4.3a. The Gaussian width

and scan length in Eq. 4.2 were kept at their measured values of W = 5 mm and L = 10 mm,

respectively. Because of the high peak intensity (~5.2x1012 W/cm2) of the laser beam and

fairly slow scan rate (~1 mm/min) used during fabrication for this particular FBG, the profiles

exhibit a strong saturation over a sizeable fraction of the length of the grating (see Fig. 4.7).

The FWHM length of the apodized profiles is 12.5 mm, which is the FWHM of the index

profile that yielded the best fit (Fig. 4.8). The maximum half-peak-to-peak ac index

modulation is 1.98x10-3. The grating period, known from the fabrication conditions

(Λ ≈ 538 nm) was also finely adjusted in the simulations (to 537.99 nm) to fine-tune the Bragg

wavelength (this parameter produces only a shift in the spectra). The noise floor of the

measured transmission spectra (imposed by the instrument) was simulated by adding a

constant offset to the simulated spectrum (-32.2 dB in Fig. 4.8a and -34.8 dB in Fig. 4.8 b)

(the noise floors differ because of different settings in the measurement apparatus).

There is a reasonable agreement between the two sets of measured and predicted spectra,

especially the FBG’s bandgap (Fig. 4.8a) and the locations and amplitudes of both the

transmission peaks (Fig. 4.8b) and the group-delay peaks (Fig. 4.8c). The model predicts that

Page 88: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

76  

this grating should have a few more slow-light peaks (deeper in the bandgap). However, these

peaks have too low a transmission to be observed. A similar agreement for the spurious peaks

due to the second eigenpolarization can be obtained by suitably adjusting the Δndc profile for

this second polarization, as described in [18]. It is important to point out that the predicted

spectra are fairly sensitive to the fitted values and to the shape of the index profile. Other

profile shapes close to the one we used gave fits of noticeably different quality. Thus our

model can predict experimental spectra very accurately.

4.4 Conclusions

In this chapter we discussed the improvements we made in the design of the FBGs and their

fabrication process. Specifically, we showed that apodized FBGs can significantly increase the

group delay and consequently the sensitivity of FBGs compared to uniform FBGs. We also

showed that deuterium-loaded FBGs written with a femtosecond laser have some of the largest

Δnac reported in an FBG and at the same time very low internal losses, two properties that are

very important to achieve high group delays and high sensitivities to strain. Finally, we

described in detail how the FBGs used in this thesis were fabricated and how we modeled

their index-modulation profiles, their loss profiles, and their transmission and group-delay

spectra.

At this point we have all the needed parts in place. In Chapter 3 we studied the thermal phase

noise in an FP-like resonance. In this chapter we showed how we can get very strong

resonances by fulfilling the requirements discussed in Chapter 2. In the next chapter we

combine the results from these chapters to demonstrate how we designed and fabricated an

FBG that verifies our phase-noise theory, and observed with it the thermal phase-noise of a

grating for the first time. We also show how, based on the results of Chapter 3, we reduced

this thermal phase noise in a second FBG to achieve a record low MDS of 30 fε/√Hz and an

absolute measurement of strain of 250 attostrains, both at 30 kHz.

         

Page 89: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

77  

References

[1] H. Wen, G. Skolianos, F. Fan, M. Bernier, R. Vallée, and M. J. F. Digonnet, “Slow-

light fiber-Bragg-grating strain sensor with a 280-femtostrain/√Hz resolution,” Journal

of Lightwave Technology, 31, 1804 (2013).

[2] L. A. Wang, C. W. Hsu, and H. L. Chen, “Characteritics of deuterium-loaded fiber

Bragg gratings,” Japanese Journal of Applied Physics, 37, 6001 (1998).

[3] M. Bernier, Y. Sheng, and R. Vallée, “Ultrabroadband fiber Bragg gratings written

with a highly chirped phase-mask and infrared femtosecond pulses,” Optics Express,

17, 3285 (2009).

[4] T. Erdogan, “Fiber grating spectra,” Journal of Lightwave Technology, 15, 1277

(1997).

[5] H. Wen, G. Skolianos, M. J. F. Digonnet, and S. Fan, “Slow light in fiber Bragg

gratings,” Photonics West, San Francisco, California, Proc. of SPIE Vol. 7949, 79490-

E1–E11 (2011).

[6] H. Wen “Ultra-high sensitivity strain sensor using slow light in fiber Bragg gratings”

Ph.D. dissertation (Stanford University, 2012) Sect. 2.3.

[7] R. Kashyap “Fiber Bragg Gratings” 2nd edition, Academic press, 35-7 (2010).

[8] A. Othonos, and K. Kalli, “Fiber Bragg Gratings Fundamentals and Applications in

Telecommunications and Sensing” ,Norwood, MA: Artech House, 25 (1999).

[9] D. P. Hand, and P. S. Russell, “Photoinduced refractive-index changes in

germanosilicate fibers,” Optics Letters, 15, 102 (1990).

[10] T. E. Tsai, G. M. Williams, and E. J Friebele, “Index structure of fiber Bragg gratings

in Ge–SiO2 fibers,” Optics Letters, 22, 224 (1997).

[11] P. J. Lemaire, A. M. Vengsarkar, W. A. Reed, and V. Mizrahi, “Refractive index

changes in optical fibers sensitized with molecular hydrogen,” Conference on Optical

Fiber Communication, vol 4 of 1994 OSA Technical Digest Series, Optical Society of

America, 47 (1994).

[12] J. Stone, “Interactions of hydrogen and deuterium with silica optical fibers: A

review,” Journal Lightwave Technology 5 712 (1987).

[13] R. M. Atkins, and P. J. Lemaire, “Effects of elevated temperature hydrogen exposure

on short-wavelength optical losses and defect concentrations in germanosilicate

fibers,” Journal of Applied Physics, 72, 344 (1992).

Page 90: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

78  

[14] M. Bernier, F. Trépanier, J. Carrier, and R. Vallée, “High mechanical strength fiber

Bragg gratings made with infrared femtosecond pulses and a phase mask,” Optics

Letters 39, 3646 (2014).

[15] W. H. Loh, M. J. Cole, M. N. Zervas, S. Barcelos, and R. I. Laming, “Complex

grating structures with uniform phase masks based on the moving fiber–scanning

beam technique,” Optics Letters 20, 2051 (1995).

[16] C. Smelser, S. Mihailov, and D. Grobnic, “Formation of type I-IR and type II-IR

gratings with an ultrafast IR laser and a phase mask,” Optics Express 13, 5377 (2005).

[17] M. Bernier, S. Gagnon, and R. Vallée, “Role of 1D optical filamentation process in

the writing of first order fiber Bragg gratings with femtosecond pulses at 800 nm,”

Optical Materials Express, 1, 832 (2011).

[18] G. Skolianos, M. Bernier, R. Vallée, and M. J. F. Digonnet, “Observation of ~20-ns

group delay in a low-loss apodized fiber Bragg grating,” Optics Letters, 39, 3978

(2014).

[19] D. Grobnic, S. J. Mihailov, and C. W. Smelser, “Localized high birefringence induced

in SMF-28 fiber by femtosecond IR laser exposure of the cladding,” Journal of

Lightwave Technology, 25, 1996 (2007).

[20] M. Bernier, R. Vallée, B. Morasse, C. Desrosiers, A. Saliminia, and Y. Sheng,

“Ytterbium fiber laser based on first-order fiber Bragg gratings written with 400nm

femtosecond pulses and a phase-mask,” Optics Express, 17, 18887 (2009).

[21] A. Siegman, “Lasers”, Chapter 11.3, University Science Books Sausalito, CA,

(1986).

[22] G. Skolianos, A. Arora, M. Bernier, and M. J. F. Digonnet, “Slowing down light to

300 km/s in a deuterium-loaded fiber Bragg grating,” Optics Letters 40, 1524 (2015).

[23] L. M. McDowall, and R. A. L. Dampney, “Calculation of threshold and saturation

points of sigmoidal baroreflex function curves,” American Journal of Physiology —

Heart and Circulatory Physiology, 291, (2006).

[24] C. W. Smelser, S. J. Mihailov, and D. Grobnic, “Rouard’s method modeling of type I-

IR fiber Bragg gratings made using an ultrafast IR laser and a phase mask,” Journal of

Optical Society of America B, 23, 2011 (2006).

[25] A. Dragomir, D. N. Nikogosyan, K. A. Zagorulko, P. G. Kryukov, and E. M. Dianov,

“Inscription of fiber Bragg gratings by ultraviolet femtosecond radiation,” Optics

Letters, 28, 2171 (2003).

Page 91: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

79  

Chapter 5: Measuring the intrinsic thermal phase noise and

250 attostrains using slow-light FBGs

In the previous chapters we have discussed the properties that an FBG needs in order to

achieve a high strain sensitivity (Chapter 2), and how we can realize such FBGs in practice

(Chapter 4). Also, in Chapter 3, we studied theoretically the thermal phase noise in a Fabry-

Perot cavity. This study showed that the phase noise in a FP-like resonance is proportional to

the group index of the resonance and to the square root of the length of the resonator L. As we

have shown in Chapter 2, the sensitivity in a slow-light resonance is proportional to the group

delay, which is proportional to the group index and the length of the resonator, and the

minimum detectable strain (MDS) is by definition the noise over the sensitivity. Thus in a

strain sensor with an MDS limited by phase noise, the MDS increases as the sensor becomes

shorter as 1/√L. Consequently, to make the thermal-phase noise the dominant noise in our

sensor and therefore be able to measure it, we had to fulfill two requirements: (1) our sensor

must have high sensitivity to be limited by sensitivity-dependent noise sources, which include

thermal phase noise; (2) our sensor must be short enough so that the thermal phase noise is

dominant. To fulfill both requirements simultaneously was challenging, since as the length of

the sensor becomes shorter and shorter, the sensitivity becomes smaller. We were able to

achieve both—high sensitivity in a short sensor— by increasing the group index using the

techniques discussed in Chapter 4. In a second device, to measure the lowest possible MDS

we decreased the thermal phase by increasing the length of the sensor, keeping all the other

noise sources at low levels. This second sensor had to be compatible with the rest of the

experimental setup, namely the probe laser, which raised another challenge because this laser

had a narrow tuning range, hence the FBGs’ slow-light resonances and the laser wavelength

had to be carefully matched.

In this chapter we report the observation and quantitative measurement of the thermal phase

noise in a 5-mm slow-light FBG over a range of frequencies between 1 and 10 kHz. The

measured phase-noise spectrum is in excellent agreement with the theoretical model presented

in Chapter 3, which provides experimental evidence of its validity. To further verify this

result, we replaced the 5-mm FBG with a longer FBG (2 cm) in which the thermal phase noise

in terms of MDS was predicted to be significantly lower, and the total noise also lower. As a

Page 92: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

80  

result of this lower total noise, we measured in this FBG sensor a record low MDS of

110 fε/√Hz at 2 kHz, and 30 fε/√Hz at 30 kHz. When integrating the output of this sensor at

30 kHz for about 8 hours, a record low absolute strain of 250 attostrains was measured. These

last two figures at 30 kHz are the lowest ever reported in a fiber strain sensor.

   Figure 5.1. Experimental setup used to characterize the noise spectra and the sensitivity of FBG strain

sensors. The PZT plate excited by the function generator induces a known strain on the FBG to

calibrate its sensitivity. The lock-in amplifier measures the sensor’s response

5.1 Experimental Setup

Figure 5.1 shows the experimental setup that we used to characterize our sensors. Light from

the narrow-linewidth, low-noise laser first traveled through a variable optical attenuator. This

allowed us first to operate the laser at the maximum output power, which gave the lowest

noise performance, and second to control the input power to our sensor for reasons explained

below. Then the light passed through an isolator, to protect the laser from back-reflections,

and a polarization controller (PC), which was used to tune the light polarization to excite only

one of the two eigenpolarizations of the FBG, as discussed in Chapter 4. At this point the light

was in its proper state (proper power and polarization) and was launched into the FBG sensor.

All components were fiber-pigtailed and connectorized and/or spiced together. The sensor was

mounted on a linear PZT plate, needed to induce a known strain to the FBG, and was placed

inside an anechoic enclosure to isolate it from the environment. A function generator applied a

voltage to the PZT. Finally, the optical output from the sensor was converted into an electrical

signal using two photodetectors. The first photodetector is an optical power meter (OPM),

which monitored the operating point of our sensor. For long-term measurements (several days)

the output of the OPM was used as an input to the PID controller in order to stabilize the

sensor. The second photodetector was connected to the lock-in amplifier, which measured the

response of the sensor at one specific frequency of interest. By switching the frequency we

Page 93: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

81  

were able to reconstruct the spectrum of the noise or of the MDS. The outputs from the lock-in

amplifier and the OPM were captured using a data acquisition system (DAQ).

The probe laser was a new laser from Orbits Lightwave with a linewidth below 200 Hz. The

previous laser used in earlier work had a linewidth of 8 kHz [1]. This 40-fold reduction in the

linewidth was expected to reduce the laser frequency noise, and therefore the MDS of the

FBG strain sensor, by at least 6.5 times, since the frequency noise is approximately

proportional to the square root of the linewidth (this approximation is valid for a

semiconductor laser [2]). This laser had a limited tunability of ±0.2 nm. Its actual laser

frequency noise spectrum, provided by Orbits Lightwave and converted in units of strain, is

showed as the red curves in Figs. 5.2 and 5.3.

To match the operating wavelength of the laser and of the FBG, we actually order a custom-

made laser to match the wavelength of our FBGs. The main issue is that the slow-light

resonances of a typical strong FBG span a fraction of a nm, and the control over the

fabrication parameters is not accurate enough to predict from these parameters alone the exact

wavelength of the steepest slope of the best resonance of a grating. We therefore could not

predict that this wavelength would fall within the 400-pm tuning range of the laser. The

approach we took to solve this problem is to fabricate a large number of strong apodized

FBGs in D2-loaded fibers, and annealed them one by one to maximize the T0τg,max product.

We then purchased thr custom-made laser at the wavelength that maximizes the number of

FBGs whose best resonance (in terms of the highest T0τg,max product) fall within the tuning

range of the laser.

As discussed briefly in Chapter 4, the method used to inscribe the FBGs induces strong

birefringence in the fiber, causing the slow-light peaks to split [3,4]. If both eigenpolarizations

were excited, only a portion of the input light would be transmitted and would be useful for

the sensor. The rest of it would be reflected. Thus in this case the sensitivity would be greatly

reduced. In the extreme, unrealistic case that the probing wavelength is at the steepest slope of

one eigenpolarization but the input light is polarized orthogonal to this eigenpolarization the

sensitivity would be zero, since no light is transmitted. Thus, in order to achieve the

maximum possible output power and hence the maximum sensitivity, we used the PC to excite

only one eigenpolarization.

Page 94: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

82  

The second reason for the use of an optical attenuator is as follows. If the input power of the

light prior to the FBG is too high, because of the expected intensity enhancement in a

resonator with a high Q factor, as explained in Chapter 6 the intensity inside the FBG is very

high. This intensity is converted into heat via absorption loss, causing the fiber to heat up. The

temperature change shifts the Bragg wavelength, causing the output power to vary. Then

because the laser is off resonance, the intensity inside the FBG decreases to a very low value,

the FBG cools down, and the Bragg wavelength shifts back to its initial, normal-temperature

value. This intensity-induced power variation results in an instability in the FBG output, as

observed experimentally for example in [5]. Thus in order to reduce this unwanted source of

noise, we had to reduce the input power and hence reduce the power that is converted into

heat. Consequently, the presence of the attenuator was crucial. It is worth mentioning that the

attenuator was needed mainly for the 5-mm FBG, since, as described in Chapter 6, the

intensity enhancement of the FBGs written with our method reaches its maximum around this

length.

In order to stabilize our sensor against laboratory temperature fluctuations, we used a PID

feedback loop. We did not use the derivative component in the PID controller because it

drastically increased the noise in the kHz range. The PID controller monitored the output

power of the FBG using the OPM (see Fig. 5.1). We first measured the output power of the

FBG at the point where the sensitivity was maximum, and we set this value as the target value

in the PID controller. If the relative position between the resonance and the laser drifted, either

because the laser wavelength or the slow-light resonance drifted, the output power of the FBG

changed and the PID controller applied a correction voltage to the dedicated input of the laser

in order to counter-act the power change. This applied voltage finely tuned the laser

wavelength by changing the length of the laser cavity. Thus the PID controller forces the laser

wavelength to be always the same as the wavelength where the resonance has its steepest

slope, and hence cancel out any drift in the sensitivity of the FBG sensor. This feedback loop

was very effective: as discussed further on, it enabled us operate the sensor continuously up to

four days.

Page 95: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

83  

5.2 Measuring the strain sensitivity and MDS of FBG sensors

As described in Chapter 2 the sensitivity is defined as:

SN (λprobe ) =1Pin

dPoutdε

!

"#

$

%&λprobe

=dλBdε

dTdλ!

"#

$

%&λprobe

(5.1)

Thus, in order to measure the sensitivity of a very sharp resonance, we need to measure the

change in the input power when a known strain is applied to the FBG. To apply a strain to the

FBG, the fiber was attached to a linear PZT plate. In our earlier work [1], a circular PZT was

used instead, which induced a lateral strain on the FBG that shifted the transmission spectrum.

This constant shift was easily corrected by adjusting the wavelength of the tunable laser.

However, in this work the limited tunability of the new laser made it impossible to correct for

this bend-induced shift. This problem was corrected by using a linear PZT instead of a PZT

ring. A specific peak-to-peak voltage of 0.01 V at 27 kHz was applied to the PZT, using the

function generator shown in Fig. 5.1. This frequency is close to the resonant frequency of the

PZT, thus a relatively high strain was applied (~5 nε), as explained later in detail. The

calibration of the strain applied to the sensor was achieved by using both of the following two

methods:

1. We used a thermally stable fiber Mach-Zehnder interferometer (MZI). The FBG

mounted on the linear PZT was placed in the first arm of the MZI. The length of the

second arm was carefully matched to the arm that included the FBG to minimize the

impact of differential thermal fluctuations between the two arms and have a relatively

thermally stable MZI. The MZI was further stabilized by controlling the length of the

second arm using a PZT and a feedback loop to stabilize the length mismatch between

the two arms. This measurement was only a few minutes long so we were not

interested in achieving long-term stability. Then we used a tunable laser tuned far

away from the FBG’s bandgap, where the FBG acts like a normal fiber, to interrogate

the MZI. Afterwards, we applied a known voltage to the PZT, which applied an

unknown strain (the strain we needed to calibrate) to the FBG. This unknown strain

changed the path mismatch between the two arms, causing the output power of the

Page 96: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

84  

MZI to oscillate at the same frequency as the applied strain. From the measured

amplitude of the modulated output power we calculated the phase modulation, and

hence the length modulation of the first arm. Finally, by dividing this length

modulation by the total length of the fiber that was attached to the PZT (which

includes the FBG), we calculated the applied strain to the FBG.

2. In this second method, we tuned a tunable laser to the maximum sensitivity of a

relatively broad slow-light resonance (low strain-sensitivity) of the FBG of interest,

which was attached to the PZT. We had previously measured the transmission

spectrum of this resonance, as described in section 4.3, which is possible to do

accurately because this resonance was by choice broad. Then, we were able to

calculate the sensitivity numerically with high accuracy. This was done by using the

rightmost equation in Eq. 5.1, namely by taking the derivative of this measured

transmission spectrum and multiplying it by the constant term dλB/dε (equal to 1.2

pm/µε, see Chapter 2). Then, as in the previous method we applied a known voltage to

the PZT, which induced an unknown strain to the FBG. This applied strain caused a

power modulation at the FBG output, as described in Chapter 2. The applied strain

was then easily inferred by dividing the measured output modulation by the measured

sensitivity and solving the middle equation Eq. 5.1 for the only unknown, the applied

strain dε=dPout/(SNPin).

Both methods gave similar results, namely an rms applied strain of ~5 nε for a peak-to-peak

voltage of 0.01 V applied to the PZT, or a response of 500 nε/V.

Knowing the strain that the PZT applied to our FBG for a given applied voltage, we could

proceed to measure the highest sensitivity of the FBG. To do that we tuned our laser at the

point where the response of our FBGs was maximum when we applied an rms strain of

Δε = 5 nε. At this point, we measured the response of our sensor ΔPout. Afterwards, using the

middle equation of Eq. 5.1, we inferred the sensitivity SN=ΔPout/(PinΔε). The noise spectrum

was measured by repeating this measurement after turning off the voltage applied to the PZT,

and tuning the operating frequency of the lock-in amplifier. The MDS spectrum was obtained

by dividing the measured noise spectrum by the measured sensitivity.

Page 97: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

85  

5.3 Measuring thermal phase-noise in a 5-mm FBG

As discussed in Chapter 3, the phase noise in an FP-like resonance in terms of strain is

inversely proportional to the square root of the resonator’s length. Thus for the thermal phase

noise to be the dominant noise source, we had to increase it above the frequency noise of our

laser. Simulations have shown that FBGs shorter than 2-3 mm do not support slow light—with

the current technology they are too short to exhibit strong enough reflections. Thus we chose

to study a 5-mm long FBG since it was relative easy to fabricate, as discussed in Chapter 4.

Figure 5.2 shows the different noise components in units of strain per square root of Hz versus

noise frequency. The green dashed curve represents the predicted phase noise in a FBG 5-mm

long, as described in Chapter 3, whereas the red solid curve represents the frequency noise.

The frequency noise of our laser was provided to us by the manufacturer, so it was

independently measured. It can be seen that between 1 kHz and 10 kHz the FBG’s thermal

phase noise is higher than the laser frequency noise. Thus, as long as the sensor is limited by

the sensitivity-dependent noise sources—which occurs when the sensor has high sensitivity—

and as long as the environmental noise is low, the dominant noise source would be the

intrinsic thermal phase noise of the FBG. The predicted total noise for such a grating is

represented by the black dashed curve in Fig. 5.2.

Figure 5.2. Noise contributions in a 5-mm FBG slow-light sensor.

Based on the foregoing, we fabricated a saturated 5-mm FBG using a femtosecond laser in a

deuterium-loaded fiber, as described in section 4.2.1. Our goal was to have an FBG with the

Page 98: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

86  

highest possible index modulation. The length of the FBG (5 mm) was achieved by writing the

FBG with a stationary laser beam with a 5-mm FWHM. Figure 5.3 shows the measured

transmission spectrum of this FBG. This spectrum was measured as described in section 4.3,

before annealing the FBG. At short wavelengths the transmission is much lower than the

expected 0 dB (100%) transmission. This is due to the fact light is coupling to cladding modes

of the fiber, which are lossy, and it does not couple back to the propagating mode. The

bandwidth of the bandgap was 3.9 nm. From this number and using Eq. 2.7 the half-peak-to-

peak index modulation prior to annealing was inferred to be 3.7x10-3, close to the highest ever

reported (~5x10-3) in [6], which met our goal of a very high index modulation.

Figure 5.3. Transmission spectrum of the FBG used for measuring thermal phase noise, measured after

fabrication (before any annealing took place).

The FBG was then annealed at 100°C, 200°C, 300°C, and 400°C for 30 minutes at each

temperature. As annealing proceeded the product Τ0τg,max was observed to constantly increase,

and therefore the sensitivity did too (see Fig 5.4). We stopped the annealing process at 400°C

since further annealing would have deteriorated the fiber and the sensitivity was already high

enough. The resonance used for sensing had a measured group delay of ~8 ns and a measured

transmission of ~0.65. The group delay was measured using our previous laser as described in

Chapter 4, therefore it may have been underestimated because of the limited resolution of that

laser (0.1 pm). Using Eq. 2.16 the sensitivity can be calculated from these two values to be

~3.25x106. Using the method described in section 5.2 the measured sensitivity was slightly

higher, namely ~4x106. The small discrepancy is mainly due to the fact that Eq. 2.16 is an

approximation, and also that the group delay is underestimated.

Page 99: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

87  

With the experimental setup described earlier the noise in this FBG (black open circles line in

Fig. 5.2) was measured. The blue solid curve represents the measured noise due to all

sensitivity-independent sources, namely the intensity noise of our laser, the photodetector

noise, and the shot noise (both electrical and optical). These noise sources were measured

using exactly the same setup, shown in Fig. 5.1, with the same output power but by replacing

the FBG with a normal fiber with zero sensitivity, and therefore in which the sensitivity-

dependent noise sources were negligible. The combined sensitivity-independent noise sources

are clearly negligible. The total noise (dashed black curve) was calculated by geometrically

summing up all noise components, namely the measured laser frequency noise (solid red

curve), the predicted FBG phase noise (dashed green curve), and the sensitivity-independent

noise sources. There is only a small discrepancy around 8 kHz due to a resonance observed in

the measured data, which is due to an environmental noise in our laboratory. This resonance

was due to environmental noise in our laboratory: we have observed the same resonance in

other FBG sensors (see Fig 5.5 for example). Since the noise is dominated by the FBG phase

noise over most of the measurement frequency range, the conclusion is that this sensor was

dominated over this range by phase noise.

Figure 5.4. Evolution of Τ0τg/c versus annealing temperature for the FBG that was used to measure

phase noise.

This observation constitutes the first measurement of the thermal phase noise in a short

passive FBG sensor (it has been observed in active FBG sensors [7]). The single-pass phase

Page 100: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

88  

noise in this 5-mm fiber in this frequency range is of the order of 1.6 nrad/√Hz, which is

extremely small and hence was difficult to measure; the use of the “amplification” by a factor

of ng (about 480 in this FBG) is what made the measurement possible. The good agreement

between the measured and predicted noise spectra gives credence to the validity of the model

on which this noise spectrum was predicted, namely the theory presented in Chapter 3. In the

next section, we provide further validation in a grating with a longer length in which the phase

noise is indeed found to be negligible, as predicted by the model.

5.4 Measuring an absolute strain of 250 attostrains

In the previous section we discussed a sensor having thermal phase noise as a dominant noise

source in the frequency range of 1-10 kHz. According to the analysis in Chapter 3, if we

increase the length L of our FBG by a factor of N while maintaining the same high sensitivity,

the thermal phase noise should be reduced by 1/√N. Thus in a ~20 mm sensor (4 times longer

than the previous sensor) the thermal phase noise should be reduced by a factor of ~2. In Fig.

5.5 we can see the noise components in this case. The color code is the same as in Fig. 5.2,

with the green dashed curve representing the thermal phase noise, the frequency noise

represented with the red solid curve, and black dashed curve representing the geometrical sum

of all noise sources. In this figure the sensitivity-independent noise sources have been omitted

for clarity purposes since they are negligible. With this longer FBG the dominant noise source

is clearly the frequency noise of the laser for the whole range of frequencies between 1 kHz

and 30 kHz.

To verify these predictions experimentally, we fabricated a new, longer FBG sensor in

deuterium-loaded fiber by scanning the femtosecond laser for 2 cm, as described in section

4.2.1. Because this FBG was longer than the previous FBG we did not aim for the highest

possible index modulation, since the index modulations was not so important to achieve a high

enough sensitivity. The FBG round-trip loss increases with length, and so does the FBG

reflectivity. Thus by reducing the index modulation in a longer FBG we can reduce the loss

while maintaining the same reflectivity as in a shorter FBG. As in the previous FBG, from the

measured bandwidth of the bandgap prior to any annealing, we inferred the half peak-to-peak

index modulation to be ~1.7 x10-3, approximately half that of the previous FBG. We then

annealed the grating at 50°C, 80°C, 120°C, and 140°C for 30 minutes at each temperature. We

Page 101: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

89  

stopped at this lower temperature because the sensitivity was already high enough for the

noise to be limited by sensitivity-dependent noise sources. Further annealing would have had

two possible outcomes: (1) an increase in sensitivity, leading to a narrower peak and a smaller

dynamic range of the measured strain with no improvement in the MDS); (2) a decrease in

sensitivity because the group delay deteriorates, possibly leading to a smaller MDS. The

resonance that was used for strain sensing had a peak transmission of ~0.55 and a peak group

delay of 11.2 ns. These values lead to a predicted sensitivity of ~3.8x106. The measured

sensitivity was about the same as in the previous FBG (~4x106). The solid black curve with

the circles in Fig. 5.5 represents the noise in units of strain measured in this sensor. It can be

seen that there is a very good agreement between the measured and the predicted total noise.

The spurious resonance at ~8 kHz is observable in this case too, as well as one additional

resonance at ~ 25 kHz, again due to environmental noise. Figure 5.5 shows a measured MDS

of 110 fε/√Hz at 2 kHz, which is a factor of 2 lower than what has been reported in a Fabry-

Perot made with two FBGs [8], and 30 fε/√Hz at 30 kHz, about one order of magnitude lower

than what has been reported in our previous work [1]. This measurement also shows for the

first time that in a passive FBG sensor the thermal phase noise expressed in units of strain is

reduced as the length increases, a counter-intuitive result of the analysis in Chapter 3.

Figure 5.5. Measured and calculated noise spectrum contributions in a 20-mm slow-light FBG strain

sensor in units of strain.

Using the stabilization setup described in section 5.1 we monitored the output of our sensor

for four days. Figure 5.6 shows the time trace of this measurement. An interesting observation

Page 102: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

90  

of Fig. 5.6 is that the noise level during the night was higher than during the day. This

behavior was quite periodic and repeatable, showing that it was not just a glitch in our setup

but an actual repeatable signal from the environment. This noise variation may have been

caused by machinery that was operating during the night; we have no explanation for its

origin. This is an observation that we were not able to make before, because our previous

sensor had a poorer resolution (280 fε/√Hz at 23 kHz), which is much higher than the noise

level during both night and day. This observation shows the significance of developing

simple, low-cost ultra-sensitive strain sensors: they enable us to make observations that were

impossible before. The sharp peaks recorded by our sensor during the day (see Fig. 5.6) could

have been caused by a construction project that was taking place in an adjoining laboratory. It

is worth mentioning that in spite of these spikes the sensor continued to operate at its optimal

point, further demonstrating the efficacy of the feedback control system.

Figure 5.6. Time trace of the 20-mm sensor’s response at 30 kHz.

Using the time trace in Fig. 5.6 the Allan deviation can be calculated. The Allan deviation is a

plot of the noise in a system as a function of integration time [9]. It is obtained by integrating

the output signal of an instrument recorded over a long time with and increasingly longer

integration time, and plotting the resulting integrated noise as a function of integration time.

Using the Allan deviation we can derive some interesting conclusions about the limiting

factors in our sensor, for example the optimal integration time to achieve the lowest minimum

detectable absolute strain. In Fig. 5.7 we plot a generic Allan deviation curve to illustrate its

main characteristics. If the noise in a system is white, as the integration time increases the

Page 103: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

91  

Allan deviation decreases with a slope of -1/2: as expected for white noise, the integrated

Figure 5.7. Generic Allan deviation curve.

noise decreases as the square root of the measurement bandwidth. If there is a drift in the

instrument under test (because of some temperature change for example) around the time

constant characteristic of this drift the Allan deviation starts to increase with a slope that

depends on the physical origin of the drift, but which is typically 0, 1/2 (as in the hypothetical

example of Fig. 5.7), or 1. The absolute minimum value that can be measured occurs for the

integration time where the Allan deviation is minimum: for larger integration times the noise

decreases because of the reduced bandwidth, but integration also incorporates the drift into the

integrated noise, leading to an increased integrated noise.

Figure 5.8. Allan deviation in units of strain calculated from the data of Fig. 5.6.

Using the data from Fig 5.6 the Allan deviation in units of strain was calculated for our FBG

sensor (see Fig. 5.8). Figure 5.8 shows that even after four days there was no drift in our

Page 104: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

92  

sensor output whatsoever, in spite of the extraordinary temperature sensitivity of our sensor

(which also increases proportional to the slope of the resonance, see section 2.3). When we

integrate its output signal for ~8 hours, the absolute strain that we were able to resolve is

250 attostrains (2.5x10-16). This value is only slightly larger than the absolute strain calculated

by dividing the measured MDS of 30 fε/√Hz by the square root of the integration time

(8 hours) expressed in second (176 attostrains). This discrepancy is due to the fact that during

the night the noise was significantly increased (almost doubled), whereas the MDS of

30 fε/√Hz was measured during the day. Nevertheless, to best of our knowledge, this value of

250 attostrains is the lowest experimental absolute strain ever reported in an FBG sensor.

5.5 Conclusions

In this chapter we experimentally characterized two FBG strain-sensors. One was specifically

designed to be limited by thermal phase noise, and the second one to reduce the thermal phase

noise and achieve a record-breaking strain resolution. These measurements enabled us to

verify our theory of phase noise developed in Chapter 3, and to show in particular that longer

sensors have a lower phase noise in units of strain. Also, we experimentally verified that the

next noise source—below the laser frequency noise—that limits our ultra-sensitive FBG strain

sensors is the thermal phase noise. In an FBG with a longer length, in which by design the

phase noise was suppressed, we measured a record low MDS of 110 fε/√Hz at 2 kHz and

30 fε/√Hz at 30 kHz, two times and one order of magnitude lower than the previous records at

those frequencies, respectively [1,8]. Finally we monitored the output of our sensor for four

days at 30 kHz, and observed no sign of drift, thus demonstrating the efficacy of our electronic

control system at correcting for temperature-induced drift in the FBG spectrum. This

measurement led us to a measurement of a record absolute strain resolution of 250 attostrains.

This measurement also allowed us to make an interesting and novel observation: the noise

level during the night was much higher than during the daytime.

This chapter completes the report of the main goal of this thesis, which was to observe thermal

phase noise in an FBG sensor and to reduce the MDS significantly below the previous records.

All the previous chapters and the topics we discussed were building up to the results reported

in this chapter. In the next chapter we discuss other very interesting facts and potential

applications of these slow-light FBGs.

Page 105: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

93  

References

[1] H. Wen, G. Skolianos, S. Fan, M. Bernier, R. Vallée, and M. Digonnet, “Slow-light

fiber-Bragg-grating strain sensor with a 280-femtostrain/√ Hz resolution,” Journal of

Lightwave Technology, 31, 1804 (2013).

[2] G. Skolianos, M. Bernier, R. Vallée, and M. J. F. Digonnet, “Observation of ~20-ns

group delay in a low-loss apodized fiber Bragg grating,” Optics Letters, 39, 3978

(2014).

[3] H. Wen, G. Skolianos, M. J. F. Digonnet, and S. Fan, “Slow Light in Fiber Bragg

Gratings,” Procedings SPIE 7949, 79490E (2011).

[4] G. Skolianos, M. Bernier, R. Vallée, and M. J. F. Digonnet, “Observation of ~20-ns

group delay in a low-loss apodized fiber Bragg grating,” Optics Letters, 39, 3978

(2014).

[5] J. Upham, I. D. Leon, D. Grobnic, E. Ma, M. N. Dicaire, S. A. Schulz, S. Murugkar,

and R. W. Boyd, “Enhancing optical field intensities in Gaussian-profile fiber Bragg

gratings,” Optics Letters, 39, 849 (2014).

[6] P. J Lemaire, A. M Vengsarkar, W. A. Reed, and V. Mizrahi, “Refractive index

changes in optical fibers sensitized with molecular hydrogen,” Conference on Optical

Fiber Communication of 1994, Optical Society of America Technical Digest Series, 4,

47 (1994).

[7] G. A. Miller, G. A. Cranch, and C. K. Kirkendall, “High-Performance Sensing Using

Fiber Lasers,” Optics & Photonics News, 23, 30 (2012).

[8] G. Gagliardi, M. Salza, S. Avino, P. Ferraro and P. De Natale, “Probing the ultimate

limit of fiber-optic strain sensing,” Science 330, 1081 (2010).

[9] D. W. Allan, “Statistics of atomic frequency standards,” Proceedings of the IEEE, 54,

221 (1966).

Page 106: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

94  

Chapter 6: Other applications of slow-light FBGs

So far we have discussed how we can achieve very high group delays in an apodized FBG and

how we can use this kind of FBGs as strain sensors with unprecedented strain resolutions that

can resolve the fiber’s extremely small thermodynamic phase noise. In this chapter we discuss

other applications for this kind of slow-light FBGs [1]. Since structural slow light is often

associated with an increase in optical energy density, this research can also be directed

towards devices offering strong three-dimensional confinement. This is particularly important

to couple light efficiently to quantum emitters, of interest for low-threshold lasers [2], cavity

quantum electrodynamics [3], and enhanced nonlinear interactions [4]. These applications

generally benefit from a high Purcell factor (or cooperativity), defined as [3,5]:

f = 3λ3

4π 2n3QVm

(6.1)

where n is the medium’s refractive index, Q is the resonance quality factor, Vm is the mode

volume, and λ is again the wavelength in vacuum. Achieving a high Purcell factor therefore

requires a high Q factor, or equivalently a large group delay, in a small mode volume, which

previous chapters have shown is now readily achievable in the strong FBG that have emerged

from this work.

Purcell factor have been measured in several types of nanophotonic structures. A low value

(1.2) was observed in a photonic-crystal microcavity embedded with InAs/InP quantum wires

[6]. In a GaP microdisk a Purcell factor of 13 was reported [7]. Although these devices are

valuable because they involve materials with interesting optical properties, they tend to

produce lower Purcell factor because of higher loss. Larger Purcell factors are possible in

fiber-based devices because of the very low propagation loss of fibers around 1.5 µm. These

devices are also advantageous because light can be coupled into them easily and efficiently. In

a very short free-space Fabry Perot (FP) resonator consisting of two high reflectors deposited

on the curved tip of two single-mode fibers, a Purcell factor of 206 was achieved [8]

(calculated from Eq. 6.1 and experimental values supplied in [8]). This device, however,

requires extreme care in mirror surface preparation, and it is very sensitive to misalignment.

Page 107: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

95  

All-fiber devices are preferable for practical applications because of their immunity to

misalignment. A Purcell factor of 29.8 was measured in an all-fiber FP made of a tapered fiber

placed between FBGs [9]. This result stemmed mostly from the very strong confinement

afforded by the taper’s 500-nm waist, the downside being a fragile and less practical device

for some applications. A more appealing approach is strong FBGs, because when designed

properly they exhibit large Q factors [10,11]. FBGs are also much more robust than fiber

tapers and free-space FPs, and simpler to fabricate. An intensity enhancement of 45 was

reported in a π-shifted FBG [12], and 800 in an apodized FBG designed to exhibit slow-light

resonances [13].

In this chapter, we investigate theoretically and demonstrate experimentally that much

stronger confinements and higher Purcell factors can be obtained in apodized femtosecond

FBGs fabricated in deuterium-loaded fibers, as described in Chapter 2. Because of their large

index modulations (>3x10-3), low internal loss (as low as 0.02 m-1), and strong apodization,

these FBGs act as FP resonators and support slow-light resonances with very high group

indices, as demonstrated in Chapter 2. They should therefore also exhibit large Purcell factors

[1]. They are particularly attractive because they can be written in any fiber material, e.g.,

chalcogenide [14] or bismuth oxide [15], which would lead to interesting low-power

nonlinear-optic applications in mm-scale devices. They can also be fabricated in fibers doped

with PbS quantum dots [16] for low-threshold laser and amplifier applications. These gratings

can be written through the jacket, hence they retain the fiber’s mechanical strength [17], of

interest to many applications. This chapter begins with a numerical study showing that for

their current loss and index modulation, if their length is optimized in the range of 5–6 mm

these FBGs can support intensity enhancements in excess of 2000 and Purcell factors as large

as ~40 [1]. These predictions are verified experimentally with two FBGs with near optimum

length (~4–5 mm). In both gratings we report observing for the first time the fundamental

slow-light mode. This mode has a measured Purcell factor of 38.7 in the first grating, and an

intensity enhancement in the second grating of 1525, to our knowledge both new records for

an all-fiber device.

Page 108: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

96  

6.1 Optimizing slow-light FBGs for high Purcell factor and intensity

enhancement

The FBGs were written through the jacket in a single-mode SMF-28 compatible germano-

silicate fiber loaded with deuterium, as detailed in Chapter 4. In brief, the beam of an 806-nm

mode-locked Ti:sapphire laser was focused with an acylindrical lens through a phase mask

onto the fiber. The measured pulse energy incident on the lens was 75 µJ, high enough to

produce type-I FBGs at a fairly high writing speed while remaining below the threshold of

type-II FBGs (~150 µJ) [18]. The phase mask had a pitch of 1070 nm (Bragg wavelength of

~1550 nm).

In a femtosecond FBG the Q factor and group delay of the resonances have a complex

dependence on the shape, width, and peak magnitude of the ac and dc index-modulation

profiles, and on the loss profile as described in Chapters 2 and 4. As we discussed in Chapter 4

the index-modulation profiles themselves depend on the laser beam width and fluence, and on

the nonlinear processes relating this fluence to the photo-induced index change. To achieve

the lowest possible mode volumes, we investigated only short FBGs, i.e., fabricated with a

stationary beam. As shown in Chapter 4 the measured transmission and group-delay spectra of

these FBGs are well predicted by an ac index-modulation profile of the form

Δnac (z) = Δn01

1+ exp(−af (z)+ b)−

11+ exp(b)

#

$%

&

'( (6.2)

where the constants a and b model the nonlinear response of the index change to the laser

intensity, and Δn0 is a constant scaling factor. The function f(z) is the laser intensity profile,

which is a Gaussian with a FWHM W

f (z) = exp(−4 ln(2) z2

W 2 ) (6.3)

Page 109: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

97  

where z is the location along the FBG measured from the center. The dc index-modulation

profile is Δndc(z) = aΔnac(z), where a is a constant. Finally, the power loss coefficient is

assumed to have the same z dependence, with a maximum value (at the center) γmax.

To make headway through this large parameter space, we first modeled an FBG with the

parameter values of [11], which is representative of the best strong FBGs we currently

produce (Δn0 = 2.4×10-3, a = 0.135, b = 1.6, a = 0.667, γmax = 0.19 m-1). In these simulations,

the beam width W was the only free parameter, and it controlled the grating length. We then

asked the question of what beam widths produce the largest Purcell factor and the largest

enhancement. The width was varied from 1 to 10 mm. For each value the transfer-matrix

model described in Chapter 4 was used to calculate numerically the group delay for the

fundamental mode (the resonance closest to the bandgap), along with its peak transmission.

These simulations required dividing the FBG in segments containing many index-modulation

periods each, and therefore multiplying a modest number of matrices (~100). Calculating the

mode volume and intensity enhancement required computing the electric-field distribution

along the FBG, as done for a stack of thin films [19]. Each period was then divided into 50

segments. These simulations were far more computer intensive, requiring of the order of

100,000 matrices per mm of grating.

The predicted dependence of the Purcell factor on the laser beam width is plotted in Fig. 6.1a

(solid black curve). As the beam width (and hence the FBG width) is increased, the Purcell

factor first increases, because the reflectivity of the FBG increases and hence the Q factor

increases rapidly (for the same reason that the finesse of an FP increases when the mirrors’

reflectivities are increased), while the mode volume increases more slowly. When the width is

increased too much, the grating is longer and lossier, the Q factor decreases, and so does the

Purcell factor. There is therefore a beam width that maximizes the Purcell factor, as shown in

Fig. 6.1a. This maximum Purcell factor occurs for W ≈ 6.4 mm and is equal to ~41.4. The

intensity enhancement shows a similar dependence (solid red curve in Fig. 6.1a), for the same

physical reasons: there is a beam width (4.5 mm) that maximizes the intensity enhancement, to

a value of ~2180, nearly three times larger than previously reported [13]. As the beam width is

increased from zero the transmission at the center of the resonance (solid curve in Fig. 6.1b) is

nearly constant at first, then it decreases monotonically for widths above ~4 mm. For most

applications, a low transmission is undesirable because it leads to deterioration of the noise in

Page 110: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

98  

the output signal. As a compromise between a larger Purcell factor/field enhancement and a

lower transmission, we opted to fabricate FBGs with a beam width of 5 mm.

Figure 6.1. Simulated dependence on the laser beam width of (a) the Purcell factor and intensity

enhancement, and (b) the transmission of the best slow-light resonance of a saturated FBG, evaluated at

the peak of the slowest resonance. Solid curves simulate an FBG with the parameters of [11], and

dashed curves the experimental FBGs.

6.2 Fabrication and characterization of two FBGs optimized for high

Purcell factor and intensity enhancement

Two FBGs were fabricated. The laser delivered ~34-fs pulses at 1 kHz with a peak intensity of

~5x1012 W/cm2 and a beam width W = 5 mm. The writing time was 300 s. Both FBGs were

saturated (an increase in fluence would have led to an insignificant increase in peak index

modulation). For each FBG the transmission and group-index spectra were measured as

described in section 4.3, then the grating was annealed at 210°C, 220°C, 230°C, 240°C, and

250°C for 30 minutes each, and the two spectra re-measured after each step. The spectra

measured after annealing at 250°C are shown in Fig. 6.2. The pattern of resonances is similar

to earlier observations in strong FBGs [10,11,20,21,22]. After each annealing step all

transmission and group-delay resonances were found to have increased. Between 240°C and

250°C the group delay of the slowest resonance (1550.15 nm, see Fig. 6.2b) stopped

increasing, while the transmission resonance slightly increased. Based on previous experience,

this behavior indicated that further annealing would have degraded the group delay of this and

other resonances, and hence the Purcell factor. Annealing was thus discontinued at 250°C. All

results reported for this FBG pertain to this temperature.

Page 111: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

99  

The measured peak group delay of the slowest resonance is 22 ns (see Fig. 6.2b). Its Q factor,

obtained from its measured FWHM linewidth (0.11 pm), is 1.33×107. Experimental evidence

that this resonance is the fundamental mode is that although it has a reasonable transmission (–

15 dB, see Fig. 6.1b), no other resonance appeared on its long-wavelength side as annealing

proceeded. In contrast, in longer FBGs additional resonances continue to appear as the

annealing temperature is gradually increased (and the loss gradually decreases). Fitting these

spectra to a model, presented in the next paragraph, also confirms that this resonance is the

fundamental mode.

Figure 6.2. Measured and fitted (a) transmission and (b) group-delay spectra of the first FBG. Inset

shows its inferred index-modulation profiles.

The dashed red curves in Figs. 6.2a and 6.2b are the simulated spectra obtained by best fitting

Eqs. 6.2 to the measured spectra, as described in Chapters 2 and 3. The beam width W was

taken to have the experimentally measured value (5 mm). The ac and dc index-modulation

profiles that produced these fits are plotted in the inset of Fig. 6.2b. The parameter values of

these profiles, fitted to obtain the best visual match to the measured spectra, are

Δn0 = 3.3×10-3, a = 3.75, b = 2.289, and a = 0.689. The fitted loss is γmax = 0.44 m-1. The

higher loss compared to the FBG of [11] (0.19 m-1) is consistent with the higher index

modulation (2.38×10-3 vs. 1.98×10-3). The measured and predicted spectra agree reasonably

well, especially the amplitude, linewidth, and group delay of the peaks, which gives credence

to the fitted profiles. The FWHM of the inferred index profiles (4.5 mm) is a little narrower

than the beam width (5 mm) because of the nonlinear response of the fiber index to the

fluence, which is taken into account in the sigmoid of Eq. 6.2 but likely with not quite the

same intensity dependence as in the actual FBG.

Page 112: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

100  

The mode volume is Vm = Aeff Leff, where Aeff is the area of the fiber transverse mode and Leff

the effective length of the FBG’s longitudinal mode. The mode area is the area integral of the

intensity across the fiber [23]

Aeff = 2π exp −2 2r /D( )2( )r dr = πD2

80

∫ (6.4)

where r is the radial direction in the fiber’s cylindrical coordinates and D is the 1/e field

diameter of the transverse fiber mode. For the SMF-28 fiber, D = 10.5 µm, which gives

Aeff = 43.3 µm2. The mode effective length was obtained by calculating the intensity

distribution of the fundamental mode along the z axis of the FBG, as described in relation to

Fig. 6.1 and using the inferred index-modulation profile (inset of Fig. 6.2b). Figure 6.3 plots

the envelope of the amplitude of the intensity distribution, normalized to the input intensity,

for this mode (red curve). (The actual intensity profile has a high-frequency sinusoidal

variation that follows the FBG’s index modulation.) The mode is concentrated symmetrically

near the center of the FBG. Its effective length is obtained by integrating numerically this

intensity profile along z. This gave a value of Leff = 750 µm, much shorter than the FBG

length, as expected. The mode volume is then Vm = Aeff Leff = 3.28×104 µm3. Inserting this

value and the measured Q factor (1.33x107) in Eq. 6.1 gives a Purcell factor of 38.7. This

compares favorably to the Purcell factor of ~30 observed in a bitapered fiber FP [9], since the

device reported here is simpler, smaller, significantly more robust, and much easier to

fabricate.

Figure 6.3. Calculated distribution of the intensity distribution along the FBG for the three lowest

modes.

Page 113: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

101  

For comparison, we also plotted in Fig. 6.3 the intensity distribution of the next two

longitudinal modes. The maximum intensity in the fundamental mode is ~1000 times higher

than the input intensity. The second mode exhibits two maxima, and its maximum intensity is

slightly higher (~1100), as observed and explained in [13]. Specifically, the second mode has

two lobes because as we explained in Chapter 4 it corresponds to a longer effective FP, and

the second mode in an FP has two maxima too. The maximum intensity is higher because, as

explained in Chapter 4, this peak corresponds to an effective FP with lower reflectivities since

the Δnac is smaller. The third mode has three maxima, all about a third as strong as the

fundamental mode. Higher order modes have a longer effective length, and a lower Purcell

factor.

Since the solid curves in Fig. 6.1 were generated for the parameter values of a different FBG,

these curves were re-simulated using the values of Δn0, a, b, a, and γmax inferred from our

measurements. The new simulations (dashed curves in Fig. 6.1) show similar trends, with only

slightly shorter optimum FBG widths and slightly lower maxima, for both the Purcell factor

and the intensity enhancement. The top vertical axis is the FWHM of the FBG (which is

always a little shorter than the beam width). The value of the FBG’s Purcell factor inferred

from measurements (38.7, filled circle in Fig. 6.1a) is in good agreement with the predicted

value (31.8, open circle) predicted for the actual FBG length of 4.5 mm (see inset of Fig.

6.2b). These two values differ because the fitted resonance predicts a group delay (and

therefore a Q factor) ~20% lower than the measured value (see Fig. 6.2b).

The second FBG was fabricated under nominally identical conditions. However, it was

expected to differ slightly because of small differences in the alignment of the beam with

respect to the fiber, and in the amount of time the D2-loaded fibers were left at room

temperature before the FBGs were written. The second FBG was annealed up to 300°C in

steps of 100°C for 30 minutes each. Compared to the first FBG, its inferred index-modulation

profile has a similar shape, its maximum ac index modulation is slightly higher (2.43x10-3)

and its peak loss is the same (0.44 m-1). It is also ~20% shorter (3.9 mm), likely due to the

aforementioned differences. Its measured Q is 7.3x106 (lower by ~2). The reason is that it is

shorter, hence its reflectivity is lower, and its fundamental mode experiences a higher effective

round-trip loss. Its mode volume, calculated from its inferred index profile, is 3.2×104 µm3,

almost the same as the first FBG. Its measured Purcell factor is 21 (solid square in Fig. 6.1a),

Page 114: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

102  

which is lower by ~2, because its Q factor is half as large. Because this FBG has sensibly the

same loss and index modulation as the first FBG, its Purcell factor and intensity enhancement

are also predicted by the dashed curves in Fig. 6.1a. The predicted Purcell factor (from Fig.

6.1a with an FBG length of 3.9 mm) is 24.2 (open square). It is slightly higher than the

measured value (by 15%), again because the fitted group delay is 15% higher than the

measured group delay. Its intensity enhancement is 1525. It is higher than for the first FBG

because the fundamental mode is closer to a matched FP mode.

6.3 Conclusions

In conclusion, numerical simulations of strong FBGs using index profiles and loss inferred

from measurements show that the slow-light resonances of these devices can exhibit large

intensity enhancements and Purcell factors when their length is optimized (5–6 mm). These

predictions are verified experimentally in two short type-I FBGs fabricated in D2-loaded fiber

and annealed to reduce their loss. For each grating the measured transmission and group-delay

spectra are fitted to a model to infer the index-modulation profiles and loss, from which the

Purcell factor and intensity enhancement are calculated. The fundamental slow-light mode

was observed in both gratings, for the first time. The fundamental mode of the first grating has

a measured Purcell factor of 38.7. The intensity enhancement of the second grating is 1525.

Both values establish new records for an all-fiber device.

References

 [1] G. Skolianos, A. Arora, M. Bernier, and M. J. F. Digonnet, “High Purcell factor in

fiber Bragg gratings utilizing the fundamental slow-light mode,” Optics Letters, 40,

3440 (2015).

[2] T. Baba, and D. Sano, “Low-threshold lasing and Purcell effect in microdisk lasers at

room temperature,” IEEE J. Sel. Top. Quantum Electronics, 9, 1340 (2003).

[3] T. J. Kippenberg, S. M. Spillane, and K. J. Vahala, “Demonstration of ultra-high-Q

small mode volume toroid microcavities on a chip”, Applied Physics Letters, 85, 6113

(2004).

Page 115: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

103  

[4] J. Bravo-Abad, A. Rodriguez, P. Bermel, S. Johnson, J. D. Joannopoulos, and M.

Soljacic, “Enhanced nonlinear optics in photonic-crystal microcavities,” Optics

Express, 15, 16161 (2007).

[5] E. M. Purcell, “Spontaneous emission probabilities at radio frequencies,” Physics

Review, 69, 681 (1946).

[6] J. Canet-Ferrer, L. J. Martínez, I. Prieto, B. Alén, G. Muñoz-Matutano, D. Fuster, Y.

González, M. L. Dotor, L. González, P. A. Postigo, and J. P. Martínez-Pastor, “Purcell

effect in photonic crystal microcavities embedding InAs/InP quantum wires,” Optics

Express, 20, 7901 (2012).

[7] K. M. C. Fu, P. E. Barclay, C. Santori, A. Faraon, and R. G. Beausoleil, “Low-

temperature tapered-fiber probing of diamond nitrogen-vacancy ensembles coupled to

GaP microcavities,” New Journal of Physics, 13, 055023 (2011).

[8] D. Hunger, T. Steinmetz, Y. Colombe, C. Deutsch, T. W. Hänsch, and J. Reichel, “A

fiber Fabry–Perot cavity with high finesse,” New Journal of Physics, 12, 065038

(2010).

[9] C. Wuttke, M. Becker, S. Bruckner, M. Rothhardt, and A. Rauschenbeutel,

“Nanofiber Fabry-Perot microresonator for nonlinear optics and cavity quantum

electrodynamics,” Optics Letters, 37, 1949 (2012).

[10] G. Skolianos, M. Bernier, R. Vallée, and M. J. F. Digonnet, “Observation of ~20-ns

group delay in a low-loss apodized fiber Bragg grating,” Optics Letters, 39, 3978

(2014).

[11] G. Skolianos, A. Arora, M. Bernier, and M. J. F. Digonnet, “Slowing down light to

300 km/s in a deuterium-loaded fiber Bragg grating,” Optics Letters, 40, 1524 (2015).

[12] I. V. Kabakova, C. Martijn de Sterke, and B. J. Eggleton, “Bistable switching and

reshaping of optical pulses in a Bragg grating cavity,” Journal of Optical Society of

America B, 27, 2648 (2010).

[13] J. Upham, I. D. Leon, D. Grobnic, E. Ma, M. N. Dicaire, S. A. Schulz, S. Murugkar,

and R. W. Boyd, “Enhancing optical field intensities in Gaussian-profile fiber Bragg

gratings,” Optics Letters, 39, 849 (2014).

[14] M. Asobe, “Nonlinear optical properties of chalcogenide glass fibers and their

application to all-optical switching,” Optical Fiber Technology, 3, 142 (1997).

Page 116: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

104  

[15] D. Grobnic, R. B. Walker, S. J. Mihailov, C. W. Smelser, and L. Ping, “Bragg

gratings made in highly nonlinear bismuth oxide fibers with ultrafast IR radiation,”

Photonics Technology Letters, 22, 124 (2010).

[16] A. Bhardwaj, A. Hreibi, C. Liu, J. Heo, J.-M. Blondy, and F. Gérome, “High

temperature stable PbS quantum dots,” Optics Express, 21, 24922 (2012).

[17] M. Bernier, F. Trépanier, J. Carrier, and R. Vallée, “High mechanical strength fiber

Bragg gratings made with infrared femtosecond pulses and a phase mask,” Optics

Letters, 39, 3646 (2014).

[18] C. Smelser, S. Mihailov, and D. Grobnic, “Formation of type I-IR and type II-IR

gratings with an ultrafast IR laser and a phase mask,” Optics Express, 13, 5377

(2005).

[19] M. Troparevsky, A. Sabau, A. Lupini, and Z. Zhang, “Transfer-matrix formalism for

the calculation of optical response in multilayer systems: from coherent to incoherent

interference,” Optics Express, 18, 24715 (2010).

[20] H. Wen, G. Skolianos, M. J. F. Digonnet, and S. Fan, “Slow Light in Fiber Bragg

Gratings,” Procedings SPIE 7949, 79490E (2011).

[21] T. Erdogan, “Fiber grating spectra,” Journal of Lightwave Technology, 15, 1277

(1997).

[22] V. Mizrahi, and J. E. Sipe, “Optical properties of photosensitive fiber phase gratings,”

Journal of Lightwave Technology, 11, 1513 (1993).

[23] L. C. Andreani, G. Panzarini, and J. M. Gérard, “Strong-coupling regime for quantum

boxes in pillar microcavities: Theory,” Physical Review B, 60, 13276 (1999).

Page 117: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

105  

Chapter 7: Conclusions and future work.

In this thesis we studied slow-light FBGs mainly to develop a strain sensor with

unprecedented resolution, and to observe and quantify thermal phase noise in a passive FBG

for the first time. In order to meet these goals we first studied theoretically the magnitude and

frequency dependence of the thermal phase noise in Fabry-Perot resonances. This study

showed that the phase noise is proportional to the group index of the resonance, and inversely

proportional to the square root of the FBG length (when expressed in units of strain). Second,

we showed how we can increase the group index theoretically and experimentally in an FBG

in order to achieve high strain sensitivity in short FBGs, a needed condition to observe thermal

phase noise. Specifically, we were able to increase the group delay by a factor of ~8, to 42 ns

in a 12.5-mm grating, by using a strongly apodized FBG written with a near-infrared

femtosecond laser in a deuterium-loaded fiber, then thermally annealed. Third, we reduced

several sources of noise in our experimental sensor and setup: we probed the sensor with a

laser that had a much narrower linewidth (<200 Hz) and therefore a significantly lower

frequency noise; we used a low-noise photodetector; and we isolated our sensor from the

environment by placing it in an anechoic enclosure.

These various improvements, along with our theoretical study of thermal phase noise in FP

resonances, enabled us to observe for the first time thermal phase noise in a passive FBG. The

device was 5-mm long, and it was dominated by phase noise between 1 kHz and 10 kHz. The

phase noise was found to be exceedingly small, of the order of ~70 fε/√Hz at 5 kHz, or

equivalently a single-pass phase noise of only 1.6 nrad/√Hz, in perfect agreement with the

magnitude predicted by our model. This measurement helped resolve a longstanding

controversy about the magnitude of the phase noise in a multi-pass interferometer such as a

Fabry-Perot.

To further validate this observation, and to demonstrate a slow-light FBG significantly more

sensitive to strain than reported earlier by our research group, we developed a second FBG

sensor with a longer length (20 mm) so that its phase noise would be much lower, as

anticipated by our model. As a result of this noise reduction, this FBG was found to be the

most sensitive passive FBG strain sensor ever reported, with a measured MDS of 110 fε/√Hz

Page 118: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

106  

at 2 kHz and 30 fε/√Hz at 30 kHz. Its noise was limited by the frequency noise of our laser,

like its predecessor, but at a much lower level. This is about one order of magnitude smaller

than the previous reported MDS at high frequencies, and two times smaller at 2 kHz. Both

figures represent new world records. Figure 7.1 shows the evolution of MDS over the years; it

is the same as Fig. 1.2 but with this new MDS data point included (far left blue arrow). When

this work started the record MDS in a single FBG was 280 fε/√Hz. Also, prior to this thesis

strain resolutions at the level reported in this work (around or smaller than 100 fε/√Hz) had

been attained only in active fiber sensors. One of the main goals of this thesis—to match

and/or exceed the resolution of active FBG sensors—has been met. Strain sensors based on

strong FBGs and ultra-narrow lasers now outperform active fiber strain sensors.

Figure 7.1 Evolution of the MDS of selected strain sensors using FBGs reported over the years.

Such sensitive fiber sensors can be used to sense other quantities, such as humidity or acoustic

waves, that have very weak but measurable effects on the fiber’s properties and hence on the

spectrum of the FBG. To demonstrate this potential, we tested the performance of the same

20-mm FBG as an acoustic sensor by placing it in an acoustically isolated chamber and

subjecting it to sound waves from a speaker. The FBG acoustic sensor was found to have a flat

band between roughly 500 Hz and 2 kHz, with an average minimum detectable pressure

(MDP) of 600 µPa/√Hz over this range. Between 3 kHz and 6 kHz the average MDP was as

low as ~50 µPa/√Hz. These values are about one order of magnitude lower than the best

previously reported values for an FBG acoustic sensor. They are still on average about a factor

of ~100 higher than reported in diaphragm-based fiber FP sensors, but this result was achieved

in a much more robust device that does not involve a fragile diaphragm as a transducer. These

preliminary results are very promising. They can readily be improved significantly in the

Page 119: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

107  

future, for example by specifically designing the fiber in which the FBG is written to be more

sensitive to acoustic waves.

Finally, we showed that with such FBGs a high Purcell factor and intensity enhancement can

be achieved as a result of the strong temporal and spatial confinement that these devices offer.

Specifically, after optimizing the length of our FBG, we demonstrated a Purcell factor of 38.7

(the highest in an all-fiber device) and an intensity enhancement of 1545, which is the highest

in an FBG. Both numbers are in good agreement with the values predicted from our model.

The development of ultrasensitive strain sensors reported in this thesis could be extended in at

least two directions, namely reducing the MDS of slow-light FBGs at high frequency beyond

what we have achieved, and demonstrating MDSs at the fε/√Hz level at much lower

frequencies (below ~10 Hz). In order to achieve these goals, the frequency noise of the laser

needs to be further reduced. This can be done by switching to a laser with an even smaller

frequency noise (recently new laser from OEwaves claims to have developed a commercial

laser with a sub-Hz linewidth), or by locking our current laser to an ultra-stable reference

frequency, i.e., an ultra-stable cavity, a frequency comb, etc., using the Pound-Drever-Hall

technique. With this technique, it has been demonstrated that sub-Hz linewidths are

achievable. With such ultra-narrow linewidths, the FBG phase noise will become dominant, as

we have demonstrated, and it will have to be reduced. This can be accomplished in one of at

least three ways: (1) decreasing the FBG temperature, at least for a proof of concept; (2)

increasing the FBG length; and (3) canceling out the thermal phase noise using signal

processing techniques. The first two methods take advantage of the physics of the thermal

phase noise and reduce it directly. The third one would involve measuring the thermal phase

noise independently, then subtracting it from the signal. This can be achieved for example by

having two identical FBGs and only one subjected to the strain. Then by subtracting their

outputs either optically or electronically the signal can be isolated from the noise. More

sophisticated techniques can be used as well, such as encoding the output signal in such a way

that it can be separated from the noise, e.g., the sensor is susceptible to strain only for a small

period of time whereas the phase noise is always present. Again the phase noise would be

reduced by using two measured signals, one with and one without the strain, and subtracting

them.

Page 120: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

108  

Last but not least, we can take advantage of these exceptional properties to enable whole new

applications for these FBGs by writing FBGs in fibers doped with active materials such as a

rare-earth ion or quantum dots to produce ultra-narrowband lasers. These FBGs can also be

written in highly nonlinear fibers to enhance nonlinear effects such as four-wave mixing, self-

phase modulation, and cross-phase modulation. The devices with high Purcell factors that we

have developed can be used to fabricate low-threshold lasers as well. These new FBGs can be

very useful in optical communication and in all-optical information processing. Finally, these

FBGs can be used as a platform to study more fundamental physics involving light-matter

interactions or weak phenomena that could not be observed before. For example, in FBGs

cooled to cryogenic temperature one can investigate whether the thermodynamic description

of thermal-phase noise developed by Wanser and in this work breaks down at low

temperatures, and how the incident light affects the thermodynamic state of the fiber.  

 

Page 121: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

109  

Appendix: Power spectral density on resonance

In this appendix we derive the PSD of the phase noise for the signal transmitted by a Fabry

Perot interferometer on resonance (Eq. 3.15), discussed in Chapter 3, starting from the

expression of the intra-cavity field Ec2 (Eq. 3.9) and assuming that the autocorrelation function

of the phase noise φ(z, t) is described by Eq. 3.5. Making the generally valid assumption that

the phase noise terms are small, the exponential terms in Eq. 3.9 can be expanded to first order

in the phase noise to give

Ec2

E0

= t1e− iΦ0 e

−i φ (z,t− zv

0

L

∫ )dz

{1+ Rn

n=1

∑ [1− i fmm=1

n

∑ ]} (A1)

where fm = (φ(z, t − 2mL − zv

)+φ(z, t − 2mL + zv

))dz0

L

∫ and R = r1r2e− i2Φ1 . Summing the first two

terms (independent of phase noise) in the {} term gives

Ec2

E0

= t1e− iΦ0 e

−i φ (z,t− zv

0

L

∫ )dz

{1

1− R− i Rn

n=1

∑ fmm=1

n

∑ } (A2)

 With straightforward manipulation, the product of the two series can be re-written as a double

sum of a product

Ec2

E0

= t1e− iΦ0 e

−i φ (z,t− zv

0

L

∫ )dz

{1

1− R− i Rm fm

m=1

∑ Rn

n=0

∑ } (A3)

 Since this last infinite series is equal to (1 - R)-1, it can be factored out of the bracket

Ec2

E0

=t1e

− iΦ0

1−Re−i φ (z,t− z

v0

L

∫ )dz

{1− i Rm fmm=1

∑ } (A4)

Finally, expanding also the exponential term to first order in the phase noise and neglecting

Page 122: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

110  

higher order terms in the product gives

Ec2

E0

=t1e

− iΦ0

1−R(1− i φ(z, t − z

v0

L

∫ )dz)− i Rm fmm=1

∑ ) (A5)

The phase fluctuations resulting from random temperature fluctuations are given by the

argument of Ec2/E0. In general, because R is complex neither this argument nor its

autocorrelation have a simple closed form, and there is no simple analytical expression for the

PSD. However, in the particular case of a signal on resonance, Φ1 is a multiple of π and

R = r1r2 is real. The argument of Ec2/E0 (see Eq. A5) is then simply

θ (t) = φ(z, t − zv0

L

∫ )dz)+ (r1r2 )mm=1

∑ ⋅ (φ(z, t − 2mL − zv

)+φ(z, t − 2mL + zv

))dz0

L

∫ (A6)

Using the short-hand notation

φm± (z) = φ(z, t −2mL ± z

v)

φm± (z ') = φ(z ', t −2mL ± z '

v+ τ )

(A7)

 

we can write the autocorrelation < θ (t)θ (t + τ ) > as a sum of nine integrals Ik

< θ (t)θ (t + τ ) >= Ikk=1

9

∑ (A8)

given by:

 

I1 = < φ0+ (z)φ0+ (z ') > dz dz '0

L

∫0

L

∫ (A9a)

I2 = (r1r2 )n

n=1

∑ < φ0+ (z)φn− (z ') > dz dz '0

L

∫0

L

∫ (A9b)

I3 = (r1r2 )n

n=1

∑ < φ0+ (z)φn+ (z ') > dz dz '0

L

∫0

L

∫ (A9c)

Page 123: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

111  

I4 = (r1r2 )m

m=1

∑ < φm− (z)φ0+ (z ') > dz dz '0

L

∫0

L

∫ (A9d)

I5 = (r1r2 )m

m=1

∑ < φm+ (z)φ0+ (z ') > dz dz '0

L

∫0

L

∫ (A9e)

I6 = (r1r2 )m

m=1

∑ (r1r2 )n

n=1

∑ < φm− (z)φn+ (z ') > dz dz '0

L

∫0

L

∫ (A9f)

I7 = (r1r2 )m

m=1

∑ (r1r2 )n

n=1

∑ < φm− (z)φn− (z ') > dz dz '0

L

∫0

L

∫ (A9g)

I8 = (r1r2 )m

m=1

∑ (r1r2 )n

n=1

∑ < φm+ (z)φn− (z ') > dz dz '0

L

∫0

L

∫ (A9h)

I9 = (r1r2 )m

m=1

∑ (r1r2 )n

n=1

∑ < φm+ (z)φn+ (z ') > dz dz '0

L

∫0

L

(A9i)

 

For symmetry reasons, in all integrals with odd index the signs in front of z and z' have the

same sign, and they can all be evaluated the same way. The general form of the solution is

Jm±,n± = < φm± (z)φn± (z ') > dz dz '0

L

∫0

L

∫ = Rφ (τ + 2(m − n)Lv

)δ(z − z ')dz dz '0

L

∫0

L

= Rφ(τ + 2(m − n)L

v)L

(A10)

 

The remaining four integrals (even index), in which the signs in front of z and z’ are opposite,

also of the same form, given by

Jm∓,n± = < φm∓ (z)φn± (z ') > dz dz '0

L

∫0

L

∫ = Rφ (τ + 2(m − n)L ∓ z

v)δ(z − z ')dz dz '

0

L

∫0

L

= ∓v2

Rφ (y)dyτ+2 ((m−n )L/v )

τ+2 ((m−n∓1)L/v )

∫ (A11)

The phase-noise PSD is the Fourier transform of the sum of all nine Ik integrals (Eq. A8).

These individual Fourier transforms can be calculated more expediently by recognizing that

the Fourier transform of the Jm±,n± and Jm∓,n± terms can be expressed as

Page 124: SLOW LIGHT IN FIBER BRAGG GRATINGS FOR SENSING … › file › druid:gd710fc6670 › ... · I would also to thank Arushi Arora for her help in the last years of my Ph.D. journey,

 

   

112  

Sm±,n± = S f e2 jω (m−n)L

v

Sm∓,n± = S f e2 jω (m−n)L

v e∓ jω L

v sincωLv

(A12)

Using these relationships and denoting the Fourier transform of Ik as Sk, we have

S1 =F J0+,0+( ) = S f (A13a)

S3 + S5 = r1r2( )nF J0+,n+( )

n=1

∑ + r1r2( )mF Jm+,0+( )

m=1

∑ = 2Sf

1− r1r2 cos2ωLv

1+ r1r2( )2− 2r1r2 cos

2ωLv

−1

$

%

&&&

'

(

)))

(A13b)

S2 + S4 = r1r2( )nF J0+,n−( )

n=1

∑ + r1r2( )mF Jm−,0+( )

m=1

= 2Sf [1+ r1r2( )sinc 2ωLv − 2r1r2sinc

4ωLv

1+ r1r2( )2− 2r1r2 cos

2ωLv

− sinc 2ωLv]

(A13c)

Skk=6

9

∑ = r1r2( )m

m=1

∑ r1r2( )n

n=1

∑ .F Jm−,n+ + Jm−,n− + Jm+,n− + Jm+,n+( )= 2Sf

1+ sinc 2ωLv

$%&

'() r1r2( )

2

1+ r1r2( )2− 2r1r2 cos

2ωLv

(A13d)

Summing all the Sk (Eqs. A13) yields the final expression for the PSD of the phase noise

SFPt = Skk=1

9

∑ = Sf 1+cos 2ωL

v+ sinc 2ωL

v"#$

%&'2r1r2

1+ r1r2( )2− 2r1r2 cos

2ωLv

"

#

$$$$

%

&

''''

(A14)

which  is  Eq.  3.15.