0 Applications of Fourier Domain Mode Locked ... Coherence Tomography Imaging 2O9

advertisement
Applications of Fourier Domain Mode Locked Lasers for Optical
Coherence Tomography Imaging
MASSACHUSETTS INS'fE
by
OF TECHNOLOGY
Desmond Christopher Adler
AUG 07 2O9
B.Sc. Electrical Engineering
University of Alberta, Canada, 2001
LIBRARIES
S.M. Electrical Engineering and Computer Science
Massachusetts Institute of Technology, United States of America, 2004
SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING AND COMPUTER
SCIENCE IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY IN ELECTRICAL ENGINEERING AND COMPUTER SCIENCE
AT THE
OF TECHNOLOGY
INSTITUTE
MASSACHUSETTS
JUNE 2009
ARCHIVES
© Massachusetts Institute of Technology. All rights reserved.
The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic
copies of this thesis document in whole or in part in any medium now known or hereafter created.
Signature of Author:
Department of Electrical Engineering and Computer Science
February 27, 2009
Certified by:
G. Fujimoto
\JJames
Professor of Electrical Engineering and Computer Science
Thesis Supervisor
I,
Accepted by:
J
,j7
I
Terry P. Orlando
Students
on
Graduate
Committee
Chair, Department
Applications of Fourier Domain Mode Locked Lasers for Optical Coherence
Tomography Imaging
Submitted to the Department of Electrical Engineering and Computer Science in Partial Fulfillment of the
Requirements for the Degree of Doctor of Philosophy in Electrical Engineering and Computer Science at
the Massachusetts Institute of Technology
by
Desmond Christopher Adler
ABSTRACT
Optical coherence tomography (OCT) is a micrometer-resolution imaging technique that produces
cross-sectional images of sample microstructure by measuring the amplitude and echo time delay of
backscattered light. OCT imaging is performed using low-coherence interferometry, typically with a fiber
optic Michelson interferometer. OCT imaging has recently been performed by measuring the spectrum of
the interference signal in the Fourier domain. In "swept source OCT" implementations, the interference
spectra are generated with a wavelength-swept laser and photodetector. Axial image lines are obtained via
Fourier transformation of the spectra. Fourier domain techniques have extended OCT imaging speeds
from several thousand to hundreds of thousands of axial lines per second, enabling in vivo threedimensional (3D) OCT. Development of the Fourier Domain Mode Locked (FDML) laser has
significantly improved the imaging performance of swept source OCT by providing an unparalleled
combination of high sweep rates, large tuning ranges, narrow instantaneous linewidths, and low phase
noise.
This thesis develops a number of advanced OCT imaging applications using FDML laser technology.
Ultrahigh-speed sub-nanometer phase profilometry is performed by measuring the phase of the OCT
interference signal, taking advantage of the inherent phase stability of FDML lasers. Extending this
concept, phase-sensitive OCT is used to detect gold nanoshell contrast agents with extremely high signalto-noise ratios by inducing photothermal phase modulations in the sample. Working in collaboration with
industrial partners, a 3D-OCT imaging system incorporating an FDML laser is constructed for clinical
research in gastroenterology. Spiral-scanning imaging catheters are developed for use in the human
esophagus and colon, enabling high-density 3D-OCT endomicroscopy of the gastrointestinal tract.
Finally, clinical pilot studies are conducted in collaboration with medical partners to demonstrate the
utility of 3D-OCT endomicroscopy for pathology detection, treatment planning, and follow-up
assessment.
The convergence of 3D spatial resolution, imaging speed, field of view, and minimally invasive
access enabled by 3D-OCT are unmatched by most other biomedical imaging techniques. Though still
early on in its development, 3D-OCT may have a profound impact on human healthcare and industrial
inspection by enabling visualization and quantification of 3D sample microstructure in situ and in real
time.
Thesis Supervisor: James G. Fujimoto
Professor of Electrical Engineering and Computer Science
ACKNOWLEDGEMENTS
I am truly blessed to have had this opportunity. The intellectual stimulation, abundant research
opportunities and fertile environment found in Professor James Fujimoto's Optics and Quantum
Electronics group is vanishingly rare and I am honoured to have been given the opportunity to conduct
my Ph.D. work here. Of course the most valuable assets of an academic research group are the people. I
owe a debt of gratitude to those who taught and mentored me, to those who worked with me in the lab and
in the hospital, and to those who will continue our work long after I depart. I would like to especially
thank Prof. Fujimoto for his guidance and support over the years and for his patience with a student who
is still, at times, a Western Canadian hockey player in spirit. I also thank Robert Huber for his
mentorship, friendship, indomitable spirit, generosity with his time, and for passing on the secrets of the
Fourier Domain Mode Locked laser. I would not have accomplished much without Robert's help and I
won't soon forget his contributions to my career. It was a true privilege to work as one-half of "Team
FDML." To my thesis committee members, Dr. Hiroshi Mashimo and Dr. Kamran Badizadegan, heartfelt
thanks for your thoughtful input and advice. Additional praise goes to Hiroshi, Laren Becker, Marisa
Figueiredo, and the VA nursing staff for putting up with temperamental computers, fragile imaging
probes and an endoscopy suite full of MIT engineers all for the benefit of science.
I would like to thank Tony Ko, Aaron Aguirre, Pei-Lin Hsiung, and Andrew Kowalevic for paving
the way before me and for showing me what a graduate student should aspire to. Thank you very much to
Joe Schmitt and Bob Shearer at LightLab Imaging Inc. for supplying the imaging engine, early catheters,
and advice that made much of the clinical studies possible. For the many therapeutic lunchtime
conversations, I thank Vivek Srinivasan, Kenzie MacIsaac, and Alex Park. The food wasn't terrific but
the company always was. To Maciej Wojtkowski and Iwona Gorczynska, thank you for your enduring
humour in the face of near-constant crisis. There must be something in the water in Poland. I am grateful
for the earlier work of Yu Chen and Paul Herz in endoscopic OCT and for leaving me with a smoothlyrunning collaboration at the VA Boston Healthcare System. More recently it has been a pleasure working
with Chao Zhou, Tsung-Han Tsai, Hsiang-Chieh Lee, Ben Potsaid, and Shu-Wei Huang. Thank you
especially for putting up with my music in the lab. The future of the group is truly in good hands with
you. Jonathan Liu and Yueli Chen have also been good friends and I wish them the best of luck. And of
course I thank Dorothy Fleischer for all of the administrative assistance and ever-present candy dish over
the years.
I must thank the Government of Canada and the Province of Alberta for the funding assistance during
my studies. The Natural Sciences and Engineering Research Council scholarship and the Sir James
Lougheed Award were welcome reminders of my country's support. Finally, I save my last and most
appreciative thanks for my family and friends. I know that the Ph.D. process has been as challenging for
my wife Annie Kim Adler as it has been for me, especially since she agreed to move 4000 km away from
a happy life and warm weather in Vancouver to join me here. Still, Annie has been there every day with
me to celebrate when progress was good and for a shoulder to lean on when progress was less-than-good.
My parents, Desmond Alfred Adler and Eva Adler, and my little sister, Heather Adler, have, along with
Annie, been a source of constant encouragement from the day I decided to embark on this adventure and
my love goes out to them for it. My new family members, John, Susan, Peter and Nickey Kim, my
honourary "second set of parents", David and Nelly Lloyd, and the rest of my family and friends have
also been wonderfully supportive. Finally I thank Scott and Jennifer Rice, Cam and Berengere Moore,
and Darren Dimarco for never letting me forget that, no matter where life takes me, I'll always be a
prairie boy.
My hope as a departing graduate student is that I have left the group in a better position than when I
began. If this is true, it is only because of the people around me.
Before the Snow
Now soon, ah, very soon, I know
The trumpets of the north will blow,
And the great winds will come to bring
The pale, wild riders of the snow.
Darkening the sun with level flight,
At arrowy speed, they will alight,
Unnumbered as the desert sands,
To bivouac on the edge of night.
Then I, within their somber ring,
Shall hear a voice that seems to sing,
Deep, deep within my tranquil heart,
The valiant prophecy of spring.
-- W. Bliss Carman
1861-1929
This thesis is dedicated to my wife, Annie, and to our home, the True North strong and free.
TABLE OF CONTENTS
CHAPTER 1: PRINCIPLES OF SWEPT SOURCE OPTICAL COHERENCE TOMOGRAPHY....... 8
........................ 9
.......................
1.1 Introduction to O C T Im aging ..........................................
9
1.2 Introduction to Swept Source Detection ......................................................
10
1.3 Sw ept Source O CT Theory ...................................................................................
11
1.3.1 Signal to Noise R atio ........................................................................................................
12
...................
........................................
Speed
Imaging
and
1.3.2 Sensitivity
13
........................
......................................
1.3.3 Axial Resolution and Imaging Depth
14
...............
1.4 Com parison to Spectral Dom ain OCT ....................................................................
17
1.5 F igures ..............................................................................
18
.......................................................................................
1.6 References
CHAPTER 2: PRINCIPLES OF FOURIER DOMAIN MODE LOCKED LASERS ............................ 22
2.1 Conventional Wavelength-Swept Lasers ........................................................ 23
2.2 FDML Theory and Typical Performance ........................................................ 24
............... 25
.......................
2.2.1 Theory of FD M L Operation ................................................
26
...................................
2.2.2 Typical FDM L Perform ance .................................
2.3 B uffered C avity D esigns .......................................................................................................... 27
2.3.1 Single-Stage B uffering ....................................................... ................................... 28
................ 29
2.3.2 D ouble-B uffering ........................................... ......................................
30
..........
2.3.3 Future Designs and Higher Order Buffering .....................
31
2.4 Dispersive FDML Cavities and Sigma Ring Designs .............................
......................... 32
...... .. .. ... .................
2.4.1 Polarization Chrom aticity ......................
....................... 32
2.4.2 Dispersive FDML Theory ........................................
............. 33
......................
2.4.3 FD ML Lasers at 1060 nm ..................................................
2.4.4 Low Speed Broadband FDML at 1310 nm .................................................................... 35
.... .... . .... ................................................. 37
2.5 Figures ......................... ........
............................................. 4 4
..............
........
....................
..................................
ces
2 .6 R eferen
CHAPTER 3: PHASE SENSITIVE OCT USING FDML LASERS ................................................... 46
47
.............. ................
.............. .........
3.1 M otivation ..................................
48
..................................
Lasers
FDML
of
Measurements
Stability
3.2 Phase
48
................
..............................................
.....
...
3.2.1 E xperim ental Setup ........................
49
3.2.2 D ata Processing ...................... .. ........ ....... ...................................................
52
3.2.3 Conventional Swept Laser ......................................................
....... ...... ....... ................................. 53
3.2.4 Non-Buffered FDML Laser ..........................
..... 53
....... ...................... ..................
3.2.5 Buffered FD M L Lasers ................................
55
...........
3.3 Sub-N anom eter D ynam ic Sensing .................................... .......... ...........................
55
3.4 Sub-N anom eter 3D Phase Profilom etry ........................................ ..................................
57
......................................
3 .5 F igures ..................................................................................
61
.......................................................
........
........
..
3 .6 Referen ces ...................................................
CHAPTER 4: PHOTOTHERMAL DETECTION OF GOLD NANOSHELLS WITH OCT ................ 63
. ............ ......... .............. 64
4.1 M otivation ...........................................
.. .................................... 65
4.2 Sources of Contrast in O CT ......................................................
67
......................................
Modulation
Phase
Photothermal
with
4.3 Imaging Gold Nanoparticles
67
.............................................
..........
...
4.3.1 Experim ental Setup ......................................
................................................ 68
4.3.2 Sample Preparation
.......................... 68
...............
4.3.3 D ata Processing ...................... ......
69
4.4 Thermal Modeling ...........................................................
4.4.1 Phase to Tem perature Conversion ........................................................................ 69
71
.......................... 7...............................
4.4.2 Estim ated Therm al R esponses ...........................
72
........................................................................................
esults
R
ental
4.5 Experim
74
..................................................
Frequency
Modulation
4.5.1 Signal to Noise Ratio versus
75
........................................
4.5.2 Effect of Measurement Time on Signal to Noise Ratio
...................... 75
4.5.3 Comparison to Model Results ......................................
76
............................ ...........................................
4.6 Lim itations ........................................
79
............................................................................
.......
..
4 .7 F igures .......................................
83
..........................
..........................
..........
4 .8 Referen ces ........................................ ...........
CHAPTER 5: 3D-OCT PLATFORM FOR CLINICAL GASTROENTEROLOGY ..........................
................................ ..... ....................................
5.1 M otivation ....................................
.........................
........
5.2 System Description ...........................
.... .....................................
..... .......
5.3 Laser D esign and Optim ization .......................
5.3.1 Sw eep Rate ................................................................. ... .... ............ ........
......
5.3.2 Buffered versus Non-Buffered Cavity Selection ......................................
5.3.3 Sweep Linearization ..............................................
....... . ........................
5.4 Optical Frequency Clock Optim ization .......................................
5.4.1 Tradeoffs Between Imaging Depth and System Noise ...............................................
5.4.2 Tradeoffs Between Imaging Depth and Axial Resolution .............................................
5.5 System C haracterization ................................................. .......... ....... ..................................
5.5.1 A xial Resolution .......... ........... ............ .......................................
...................................
5.5.2 Sensitivity Rolloff .....................................
5.5.3 True Spatial Resolution ...............................................
.......................
5.6 D ata Post-Processing ......................... ... ............. ........ ...........................
..............................
5.6.1 Fram e Flattening .........................................................................
5.6.2 JPE G C onversion .......................... ......... ..... .................................................................
5.7 Im age V isualization .................................................... .............................................................
.....
5.7.1 3D Rendering ............................................................................
...
5.7.2 Orthoplane Sectioning ........................................
................................................
5.7.3 Projection Viewing
........................................
5.7.4 Linear En Face Images .....
......................................
.........
5.8 Figures ........................
. ........... ...........................................................
5.9 R eferen ces ............................................
88
89
89
91
92
93
94
96
96
97
98
98
99
99
100
100
102
102
102
103
103
105
105
112
CHAPTER 6: HIGH SPEED IMAGING PROBES FOR CLINICAL 3D-OCT .................................
...........................
....... ...........
............................
6.1 M otivation ...................................
6.2 Fused Fiber Lens Systems .........................................
6.2.1 ZEMAX Modeling - Standard Multimode Fiber .......................................................
6.2.2 ZEMAX Modeling - Custom Multimode Fiber ........................................
6.2.3 Polishing ........................ ...............................................................
................ . .....................
....... ......
6.2.4 Fabrication Tolerances .......................
........................................
6.2.5 Measured Performance ......
................................................
6.3 Micro-Optic Lens Systems
................................................
6.3.1 ZEMAX Modeling
......................
...
..
........
6.3.2 Minimization of Backreflection .................
.......................................
6.3.3 Measured Performance .......
6.4 Mechanical Design ..........................................................................................
6.4.1 Torque C oil Selection ............................ ... .......... .......... ............... .......................
6.4.2 Proximal Joint ............................................. .................... .....................
......... .......................
.............................................
6.4.3 Fiber Connector and Flush Port ......
114
115
115
116
117
117
117
118
119
119
120
120
121
122
122
123
........................................
6.4.4 Future Probe Design .......
6.5 Imaging Performance Comparison ..........................................
6 .6 F igures .. . ......... ...........................................................................................................................
.........................................
6.7 References ...
123
123
12 5
129
CHAPTER 7: CLINICAL 3D-OCT IN THE UPPER GASTROINTESTINAL TRACT ................... 130
. 131
.................................
.....................
7.1 M otivation ............
132
...........................................
..........................................................
....
7.2 C linical Protocol
..
134
7.3 N orm al Esophagus ......................... ... ........ ...................................................................
134
........................................
7.3.1 Characteristic Features ......
............................. 136
7.3.2 C om parison to H istology ................................................................
136
........................................
Ablation
Pre-Radiofrequency
Esophagus:
Barrett's
7.4
137
..............................
....
...................
..
.........
7.4.1 C haracteristic Features..........................
138
...
...
.. .................................
7.4.2 Comparison to Histology.....................................
139
7.5 Barrett's Esophagus: Post-Radiofrequency Ablation .............................
..... ....................... 139
7.5.1 Characteristic Features................................................... ..
................ ............................................... 141
7.5.2 Comparison to Histology.......................
........................... 14 1
7.6 E sophageal N odules ...................... ........... .............................................
7.6.1 C haracteristic Features.............................. . .. .............. .............................................. 142
...... ... .... ........................ 143
7.6.2 Comparison to Histology............. ......................
................... 143
...........................
Resection
Mucosal
Post-Endoscopic
7.7 Esophageal Nodules:
143
........................
.....
...........
7.7.1 Characteristic Features.....................
.............. 144
7.7.2 Comparison to Histology..............................
146
...... ................ .........................................
7.8 Figures ..................... ........ ..
157
..
....................................
..............
.........
.....
...............
.....
7.9 R eferences .......................
CHAPTER 8: CLINICAL 3D-OCT IN THE LOWER GASTROINTESTINAL TRACT .................
. .. .. ..............................................
...........................
8.1 M otivation ..
......
..................................
.......................
Protocol
8.2 Clinical
.........................
..
.....................
......
8.3 N orm al C olon ....................... .. ...................
.. ..................... ............. .......................
8.3.1 Characteristic Features .........................
8.3.2 Comparison to Histology ..........................................
......................................
8.4 Inflammatory Bowel Diseases .......
.................
8.4.1 C haracteristic Features .......................... ..... ...........................................
.... ..................................
8.4.2 Com parison to H istology ................................... ....... ......
8.5 Radiation Proctitis: Pre-Radiofrequency Ablation ........................................
.......................................
8.5.1 Characteristic Features .......
........................
...........
.................................
8.5.2 Comparison to Histology
8.6 Radiation Proctitis: Post-Radiofrequency Ablation ........................................
........................................
8.6.1 Characteristic Features ......
8.6.2 Comparison to Histology ............................................
.... .......... ...............................
.... .... ................
.. .
..............
8.7 Figures ..........
8.8 References ............................................. ................. ................................
159
160
161
162
163
164
164
165
166
166
166
168
168
168
169
171
178
CHAPTER 9: CONCLUSIONS, FUTURE WORK, AND PUBLICATIONS .................................
................. .................
............
9.1 Summ ary of Thesis W ork .............................
.....................................................
.
..
.........
......
9.2 Future Work .........................
9.3 Publications Produced During Thesis W ork .........................................................................
179
180
181
182
CHAPTER 1
1.0
Principles of Swept Source Optical Coherence Tomography
1.1 Introduction to OCT Imaging
OCT is an imaging technology that enables micron scale, cross-sectional and 3D imaging of sample
microstructure in real time [1-3]. For biomedical applications, OCT can function as a type of "optical
biopsy". Tissue microstructure can be imaged with resolutions approaching that of excisional biopsy and
histopathology without the need to remove and process tissue specimens [4-6]. For material inspection
applications, OCT can provide nondestructive 3D analysis of depth-resolved sample features. OCT is
analogous to ultrasound B mode imaging, except that imaging is performed by measuring the echo time
delay and intensity of back-reflected or backscattered light rather than sound. An optical beam is scanned
across the sample and echoes of backscattered light are measured as a function of axial range (depth) and
transverse position, as shown in Figure 1.1. Three-dimensional imaging can be conducted by performing
a two-dimensional scan pattern at different transverse positions. Three-dimensional OCT (3D-OCT)
enables powerful methods for visualizing tissue architecture. 3D-OCT generates comprehensive,
volumetric data sets which can be used to construct arbitrary cross-sectional images, projections along
arbitrary axes, or 3D renderings similar to those used in MRI or CT. Unlike confocal microscopy, 3D2
OCT enables imaging over a comparatively large field of view (up to -200 mm ) while maintaining
resolutions on the order of 10 tm, and can therefore provide architectural context in addition to high-
magnification views of focal abnormalities.
OCT is based on a technique known as low coherence interferometry, which has been previously
applied in photonic devices as well as in biological systems to perform optical ranging [7-9]. OCT
measurements are performed using a Michelson interferometer with a low coherence length light source
as shown in Figure 1.2. One arm of the interferometer contains a modular probe that scans the light beam
over the sample and collects the backscattered light. Fiber optic catheters and endoscopes have been
developed for imaging inside the body [6], and microscopes can be used for imaging excised tissue
specimens or material samples. A second arm of the interferometer has a scanning reference path delay
that is mechanically translated over the desired imaging depth in a classical "time domain OCT"
configuration. Optical interference between the light from the sample and reference occurs only when the
optical delays match to within the coherence length of the light [10, 11]. "Coherence" refers to a temporal
property of the light, which is inversely related to its wavelength or frequency bandwidth. Low coherence
interferometry enables the echo delay time and magnitude of backscattered light from internal tissue
microstructures to be measured with high time resolution and sensitivity.
1.2 Introduction to Swept Source Detection
Within the past 3 to 4 years, dramatic advances in OCT technology have resulted in 10 to 100 fold
increases in imaging speed [12-14]. These advances are key to enabling 3D-OCT for in vivo biomedical
applications and real-time material inspection applications. Volumetric data acquisition with 3D-OCT
will enable new visualization and processing techniques such as the generation of cross-sectional images
with arbitrary orientation, the generation of projection views similar to en face microscopy images,
improved quantitative measurements of morphology, improved image processing techniques to reduce
speckle and enhance contrast, and virtual manipulation of sample geometry for the visualization of
structural morphology.
These new OCT detection techniques, known as Fourier domain detection, can achieve very high
speeds and sensitivities by measuring backscattered light in the Fourier domain [12-15]. Conventional
OCT directly measures the interference signal, while Fourier domain OCT measures the spectrum of the
interference signal. The OCT axial scan is then constructed by Fourier transformation. While this requires
spectral measurement and additional signal processing steps, it has the advantage that all depth positions
in the sample are measured simultaneously rather than sequentially as in conventional OCT detection
techniques. Several groups working independently demonstrated in 2003 that Fourier domain detection
enables 10 to 100 fold improvements in detection sensitivity, which gives corresponding improvements in
the imaging speed [12-14].
Fourier domain OCT can be performed using two complementary techniques known as spectral
domain OCT and swept source OCT (also known as optical frequency domain imaging or OFDI).
Spectral domain detection uses a spectrometer and high-speed line scan camera to measure the spectrum
of the OCT interference signal. Spectral domain OCT typically operates at 800 nm wavelengths with axial
imaging rates of 29,000 - 75,000 lines per second (29 - 75 kHz) [16-18]. This technology has had a
powerful impact on ophthalmic OCT imaging because it enables ultrahigh image resolutions as well as
3D-OCT imaging of retinal pathologies [19, 20]. In contrast, swept source OCT uses a wavelength-swept
laser light source and a balanced pair of photodetectors to measure the interference spectrum [21-24] as
shown in Figure 1.3. Swept source OCT technology has the advantage that it can perform imaging at
longer wavelengths of 1000 nm and 1300 nm. Imaging at these wavelengths is important because it
reduces optical scattering and improves image penetration depths [5].
1.3 Swept Source OCT Theory
Swept source OCT is similar in nature to optical frequency domain ranging (OFDR) techniques that
are commonly used in telecommunications applications. In the most basic swept source OCT
implementation using a single photodetector as shown in Figure 1.3, the current output of the detector is
given by
idet (t)r
(P,+ P fr2 (z)dz+2 pr
r(z)F(z)cos(2k(t)z+ B(z))dz)
(1.1)
where iq is the photodetector sensitivity, q is the electrical charge constant, hv is the photon energy, Pr
is the reference arm power returned to the detector, and P is the power incident on the sample [23]. The
sample and light source characteristics are represented by the relative axial coordinate z, the amplitude
and phase of the sample reflectance profile r(z) and b(z), the instantaneous coherence function of the
laser source F(z), and the time-varying wavenumber of the laser output k(t)= 2n~ / (t). At z= 0,
the optical path length of the sample arm is matched to the reference arm.
The first term in Equation (1.1) results in a constant DC photocurrent due to reference arm power on
the photodetector. The second term results in a DC photocurrent that varies with the intensity of the
backscattered sample light and is commonly referred to as the "autocorrelation term." The third term is
the desired interferometric photocurrent that encodes the axial reflectance profile of the sample. The
complex sample reflectance profile can be recovered by Fourier transformation of the photocurrent, which
is typically carried out following digitization with a discrete Fourier transform (DFT) algorithm such as
the Fast Fourier Transform (FFT).
1.3.1 Signal to Noise Ratio
Swept source OCT systems provide order-of-magnitude improvements in signal to noise ratio (SNR)
compared to previous time domain OCT methods. Theoretical SNR performance, as described in ref.
[23], requires an understanding of the signal power F2 and the time-averaged noise power (F,2) where
F and F, are the DFT values of the signal and noise photocurrents is and i,. For a single sample
2
reflector located at z = zo and having a fixed reflectivity r , and assuming that the instantaneous
linewidth of the laser source is sufficiently narrow to give F (z), the signal current becomes
is (t)='q
2
(1.2)
IPcos(2k(t)z o )
where P = r2 p0 is the optical power reflected by the sample and returned to the photodetector. The timeaveraged noise power (i
2
(t)) is given by
+2
=ith2
4
+)
+
RN(Pr +
2
)2 BW
(1.3)
The three terms in Equation (1.3) represent thermal, shot, and relative intensity noise (RIN) of the laser
source in Hz -'. Here, i, is the thermal noise current in the photodetector and BW is the detector
bandwidth in Hz.
The SNR of a swept source OCT system is given by SNR = IF (z o ) / F,2 ) where F and F are
the DFTs of the signal and noise currents, respectively. The DFT value of either photocurrent at z = z, is
given by
i(km)e-j2zlm/N,
F(z,)=
(1.4)
m=0
Here i(k) is the sampled photocurrent and N s is the number of samples acquired per source wavelength
sweep. For the signal component, the DFT is zero-valued at all axial positions except for z i = z0 . Using
Parseval's theorem,
IFs (z0 )2
F2=N
i2
, the magnitude-squared of the signal DFT is given by
(N 2 /2) (i2). The factor of
arises due to the positive and negative frequency peaks
2
associated with the DFT of the real-valued signal photocurrent. The power of the noise DFT is given by
(F 2) = N2 (i2). The SNR for swept source detection therefore becomes
1(i 2
SNR =
(i2)
(1.5)
2 (i,2
Equation (1.5) illustrates one fundamental advantage of swept source detection over conventional
time domain OCT methods. In time domain OCT the SNR is simply the ratio (i2) / (i2), so swept source
detection gives an N, /2 benefit in terms of SNR. In most swept source OCT setups a dual-balanced
detector is used instead of a single photodiode, which removes excess noise from the laser source and
effectively doubles the signal level, giving an additional -3 dB improvement to SNR. If the detection
system is shot noise limited, with negligible thermal noise and laser RIN, and when P, <<Pr , the SNR
for a swept source system can be approximated as
SNR
7Ps
hv
(1.6)
where T, = 1/ fs is the wavelength sweep duration and f, is the frequency at which wavelength sweeps
are generated by the laser source. The sweep frequency is also equivalent to the rate at which axial image
lines are generated, as shown in Equation (1.1).
1.3.2 Sensitivity andImaging Speed
The sensitivity of an OCT system is defined as the minimum detectable sample reflectivity. The
minimum detectable reflectivity results in a signal power that is equal to the noise power, or, equivalently,
results in SNR = 1. Using P, = r 2 P0 and the condition SNR = 1 in Equation (1.6), the minimum
detectable reflectivity rmin 2 is given by
rmin
2
=
hvf
s
,JP
(1.7)
Sensitivity is typically expressed in dB as
, log
Sens = -10
hT
(1.8)
Compared to time domain OCT, swept source detection also conveys an NS /2 sensitivity benefit
directly linked to the SNR benefit. This advantage can be used to detect very weak sample reflections if
the system is operated at low sweep frequencies. In practice, however, the system's finite dynamic range
limits the ability to simultaneously detect strong reflections and extremely weak reflections. The
sensitivity advantage has found more utility in applications where high typical reflectivity levels are
detected but extremely high imaging rates are required.
Swept source OCT was first demonstrated at MIT in 1997, but performance was limited by available
laser technologies [21, 22]. Recent advances in wavelength-swept lasers have enabled much higher speed
imaging. Swept source OCT with axial imaging rates of 19 kHz was demonstrated in 2003 [25], and 115
kHz axial imaging rates were achieved in 2005 [26]. Using advanced Fourier Domain Mode Locked
(FDML) laser technology, our group recently achieved record axial imaging rates of 370 kHz, -100 times
faster than standard OCT [27, 28]. These order-of-magnitude increases in imaging speed have enabled
many new biomedical and industrial applications of three-dimensional OCT (3D-OCT), several of which
were developed in this thesis work.
1.3.3 Axial Resolution and Imaging Depth
Axial resolution and maximum ranging depth are important parameters that determine the utility of a
swept source OCT system. The theoretical value of axial resolution for swept source OCT is the same as
in traditional time domain OCT [1], and is related to the spectral bandwidth of the laser source used in the
imaging system. Theoretical axial resolution Az for a laser with a Gaussian power spectrum is given by
Az
Az =
= 21n2
r
02
nSA/1
(1.9)
(1.9)
where A0 is the centre wavelength of the laser source, ns is the group refractive index of the sample
medium, and AA is the full-width-at-half-maximum (FWHM) spectral bandwidth of the laser source.
Typical axial resolutions for swept source OCT systems operating at 1310 nm are 7 - 14 pm [23, 29, 30],
while systems operating at 1060 nm have achieved axial resolutions of 10 - 20 gm in air [31-33].
Imaging depth limitations in swept source OCT are fundamentally different than in time domain
OCT. In both types of OCT, optical attenuation in the sample decreases the number of backscattered
photons returning from deeper surfaces, which decreases the signal level and limits the useable imaging
range or "penetration depth". Multiple scattering effects in turbid samples such as biological tissue
compound this problem for both swept source and time domain OCT. The maximum sample depth that
the system can interrogate, independently of optical effects in the sample, is referred to as the "imaging
depth." In time domain implementations the imaging depth is determined by the maximum path
differential AL that can be obtained between the reference arm and the sample arm. Aside from the
practical constraints of reference arm scanner construction, there is no fundamental limit on imaging
depth for time domain OCT systems.
In a swept source implementation, interference fringes are acquired in the Fourier domain and
images are generated via a DFT. Deeper positions in the sample are encoded in progressively higherfrequency interference fringes, so imaging depth is therefore limited by the detection and data acquisition
electronics. The photodetector system must have sufficient bandwidth to detect the interference fringe
frequency 2k (t)
Zmax
associated with the maximum imaging depth zmax . The digitizer used to sample the
photodetector signal must also have sufficient bandwidth and sampling rate to avoid attenuation or
aliasing of the highest frequency fringe. Maximum imaging depth is given by [23] as
(1.10)
=
Z
max
4n
A
where 62, is the wavelength sample spacing following digitization. Typical imaging depths for swept
source OCT systems are 1 - 5 mm, depending on the imaging rate and data acquisition performance.
1.4 Comparison to Spectral Domain OCT
Spectral domain OCT is similar in concept to swept source OCT, except that a non-swept laser is
used as the source and a spectrometer is used in the detection arm to record the entire interference fringe
spectrum at once. This is in comparison to swept source OCT, where the spectrum is recorded point-bypoint as a function of time. Spectral domain systems typically use silicon CCD detectors with high
sensitivity near 800 nm, rolling off dramatically at wavelengths longer than 1000 nm. Although spectral
domain OCT systems can provide nearly optimal performance in low-scattering organs such as the eye
[19], these systems are not well suited for imaging scattering tissues since scattering is significantly
increased at shorter wavelengths. Imaging epithelial tissues, for example, is of great interest since the
majority of human cancers originate in epithelium that lines hollow organs such as the colon, esophagus,
and breast ducts. In epithelial tissue, the high density of cellular organelles such as mitochondria, nuclei,
and cellular membranes results in a highly scattering medium that limits the penetration depth of light at
800 nm wavelengths. Similarly, many non-biological materials also exhibit high optical scattering at short
wavelengths. Swept source OCT systems can be constructed at 1000 nm [31-33], 1310 nm [23, 29, 30],
and 1550 nm [34] wavelengths by using commercially-available telecommunications devices such as
semiconductor optical amplifiers, singlemode fiber components, and InGaAs photodetectors. The ability
to image at longer wavelengths is an important benefit for swept source OCT, although emerging InGaAs
camera technology [35] may make spectral domain OCT at 1000 nm and 1310 nm possible in the near
future.
Swept source OCT also provides performance advantages in terms of sensitivity rolloff compared to
spectral domain OCT. In both types of system the sensitivity decreases at longer sample delays. In
spectral domain OCT this effect is due to the finite spectral resolution of the spectrometer used to record
the interference signal. Deeper sample positions are associated with higher frequency interference fringes,
as shown in Equation (1.1). Since the spectrometer must integrate the interference signal over a finite time
period, random sample motion over this period causes phase jitter in the fringes that contributes to
averaging or "fringe washout." The uncertainty in the fringe measurement is more severe at higher
frequencies, causing a corresponding drop in sensitivity. In swept source OCT, sensitivity rolloff is
caused by the finite instantaneous coherence length of the swept laser in a manner analogous to the finite
spectrometer resolution. Swept lasers can be constructed with instantaneous coherence lengths of 0.06 0.2 nm [23, 29, 36] resulting in 6 dB sensitivity rolloffs at 3 - 7 mm. Spectrometer resolutions in spectral
domain OCT can be comparable, but fringe washout effects result in 6 dB sensitivity rolloffs at 1 - 3 mm
[12, 37, 38].
Finally, swept source OCT has so far provided faster imaging speeds than spectral domain OCT. This
has been a result of rapid developments in wavelength-swept laser technology, including the FDML laser
[29], along with improvements in high-speed commercial data acquisition hardware. Axial imaging rates
of up to 370 kHz have been demonstrated with swept source OCT and FDML lasers [27] and the core
technology remains scalable to higher speeds. Spectral domain OCT, on the other hand, is limited by the
sensitivity, integration time, and readout rates of commercial line-scanning detector arrays. While recent
improvements in camera technology have enabled spectral domain OCT at speeds up to 312.5 kHz, axial
resolution and sensitivity rolloff performance are markedly degraded. Progress is also limited by the
ability and willingness of commercial camera manufacturers to develop higher-speed products for what
are initially limited research markets. Swept source OCT, on the other hand, requires comparatively fewer
simultaneous commercial advances in order to scale up imaging rates. High-speed digitizer hardware is
the main limitation to further increases in swept source OCT speeds, but improved data acquisition
systems are desired by a wide variety of industries including aerospace and telecommunications
companies. Currently, swept source OCT has proven its value for applications requiring the fastest
possible imaging speeds at wavelengths of 1310 nm and 1060 nm. For applications where slower speeds
are acceptable but axial resolution is of critical importance, such as retinal imaging, spectral domain OCT
remains the dominant technology.
1.5 Figures
Sample
1D
2D
Axial (Z)Scanning
Axial (Z) Scanning
Transverse (X) Scanning
3D
Axial (Z) Scanning
XYScanning
Backscattered Intensity
i7.
Long Coherence Length Light
Figure 1.1. OCT generates cross-sectional and
3D images of tissue microstructure by measuring
the echo time delay and magnitude of
backscattered light. Architectural morphology can
be imaged in vivo and in real time.
Short Coherence Length Light
uses low coherence
Figure 1.2. OCT
interferometry to detect the time delay and
magnitude of backscattered light.
C
Detector output
frequency - distance
Figure 1.3. Swept source OCT enables a 10-100x increase in imaging speed compared to time domain
OCT. A: Michelson interferometer with path length difference AL and a wavelength-swept laser source. B:
Light from the sample (dotted) and reference (dashed) paths are time delayed and interfere. C: A
radiofrequency beat signal proportional to AL is produced on the detector. D: The Fourier transform of the
detected signal recovers the axial reflectance profile of the sample.
1.6 References
[1]
D. Huang, E. A. Swanson, C. P. Lin, J. S. Schuman, W. G. Stinson, W. Chang, M. R. Hee, T.
Flotte, K. Gregory, C. A. Puliafito, and J. G. Fujimoto, "Optical Coherence Tomography,"
Science, vol. 254, pp. 1178-1181, Nov 22 1991.
[2]
J. G. Fujimoto, C. Pitris, S. A. Boppart, and M. E. Brezinski, "Optical coherence tomography: an
emerging technology for biomedical imaging and optical biopsy," Neoplasia, vol. 2, pp. 9-25,
Jan-Apr 2000.
[3]
J. G. Fujimoto, "Optical coherence tomography for ultrahigh resolution in vivo imaging," Nature
Biotechnology, vol. 21, pp. 1361-1367, Nov 2003.
[4]
J. G. Fujimoto, M. E. Brezinski, G. J. Tearney, S. A. Boppart, B. Bouma, M. R. Hee, J. F.
Southern, and E. A. Swanson, "Optical biopsy and imaging using optical coherence tomography,"
Nature Medicine, vol. 1, pp. 970-972, Sep 1995.
[5]
M. E. Brezinski, G. J. Tearney, B. E. Bouma, J. A. Izatt, M. R. Hee, E. A. Swanson, J. F.
Southern, and J. G. Fujimoto, "Optical coherence tomography for optical biopsy. Properties and
demonstration of vascular pathology," Circulation,vol. 93, pp. 1206-13, Mar 15 1996.
[6]
G. J. Tearney, M. E. Brezinski, B. E. Bouma, S. A. Boppart, C. Pitvis, J. F. Southern, and J. G.
Fujimoto, "In vivo endoscopic optical biopsy with optical coherence tomography," Science, vol.
276, pp. 2037-2039, 1997/06/27 1997.
[7]
K. Takada, I. Yokohama, K. Chida, and J. Noda, "New measurement system for fault location in
optical waveguide devices based on an interferometric technique," Applied Optics, vol. 26, pp.
1603-1608, 1987.
[8]
H. H. Gilgen, R. P. Novak, R. P. Salathe, W. Hodel, and P. Beaud, "Submillimeter optical
reflectometry," IEEE JournalofLightwave Technology, vol. 7, pp. 1225-1233, 1989.
[9]
R. Youngquist, S. Carr, and D. Davies, "Optical coherence-domain reflectometry: a new optical
evaluation technique," Optics Letters, vol. 12, pp. 158-160, March 1987.
[10]
E. A. Swanson, D. Huang, M. R. Hee, J. G. Fujimoto, C. P. Lin, and C. A. Puliafito, "High-speed
optical coherence domain reflectometry," Optics Letters, vol. 17, pp. 151-153, 1992.
[11]
E. A. Swanson, J. A. Izatt, M. R. Hee, D. Huang, C. P. Lin, J. S. Schuman, C. A. Puliafito, and J.
G. Fujimoto, "In vivo retinal imaging by optical coherence tomography," Optics Letters, vol. 18,
pp. 1864-1866, 1993/11/01 1993.
[12]
M. A. Choma, M. V. Sarunic, C. H. Yang, and J. A. Izatt, "Sensitivity advantage of swept source
and Fourier domain optical coherence tomography," Optics Express, vol. 11, pp. 2183-2189, SEP
8 2003.
[13]
J. F. de Boer, B. Cense, B. H. Park, M. C. Pierce, G. J. Tearney, and B. E. Bouma, "Improved
signal-to-noise ratio in spectral-domain compared with time-domain optical coherence
tomography," Optics Letters, vol. 28, pp. 2067-2069, Nov 1 2003.
[14]
R. Leitgeb, C. K. Hitzenberger, and A. F. Fercher, "Performance of Fourier domain vs. time
domain optical coherence tomography," Optics Express, vol. 11, pp. 889-894, 2003/04/21 2003.
[15]
A. F. Fercher, C. K. Hitzenberger, G. Kamp, and S. Y. Elzaiat, "Measurement of Intraocular
Distances by Backscattering Spectral Interferometry," Optics Communications, vol. 117, pp. 4348, MAY 15 1995.
[16]
B. Cense, N. Nassif, T. C. Chen, M. C. Pierce, S. Yun, B. H. Park, B. Bouma, G. Tearney, and J.
F. de Boer, "Ultrahigh-resolution high-speed retinal imaging using spectral-domain optical
coherence tomography," Optics Express, vol. 12, pp. 2435-2447, 2004.
[17]
N. Nassif, B. Cense, B. H. Park, S. H. Yun, T. C. Chen, B. E. Bouma, G. J. Tearney, and J. F. de
Boer, "In vivo human retinal imaging by ultrahigh-speed spectral domain optical coherence
tomography," Opt Lett, vol. 29, pp. 480-2, Mar 1 2004.
[18]
Y. Zhang, B. Cense, J. Rha, R. S. Jonnal, W. Gao, R. J. Zawadzki, J. S. Werner, S. Jones, S.
Olivier, and D. T. Miller, "High-speed volumetric imaging of cone photoreceptors with adaptive
optics spectral-domain optical coherence tomography," Optics Express, vol. 14, pp. 4380-4394,
May 2006.
[19]
M. Wojtkowski, V. Srinivasan, J. G. Fujimoto, T. Ko, J. S. Schuman, A. Kowalczyk, and J. S.
Duker,
"Three-dimensional retinal
imaging with high-speed ultrahigh-resolution
optical
coherence tomography," Ophthalmology, vol. 112, pp. 1734-46, Oct 2005.
[20]
V. J. Srinivasan, M. Wojtkowski, A. J. Witkin, J. S. Duker, T. H. Ko, M. Carvalho, J. S.
Schuman, A. Kowalczyk, and J. G. Fujimoto, "High-definition and 3-dimensional imaging of
macular pathologies with high-speed ultrahigh-resolution optical coherence tomography,"
Ophthalmology, vol. 113, pp. 2054-2065, Nov 2006.
[21]
S. R. Chinn, E. A. Swanson, and J. G. Fujimoto, "Optical coherence tomography using a
frequency-tunable optical source," Optics Letters, vol. 22, pp. 340-342, Mar 1 1997.
[22]
B. Golubovic, B. E. Bouma, G. J. Tearney, and J. G. Fujimoto, "Optical frequency-domain
reflectometry using rapid wavelength tuning of a Cr4+:forsterite laser," Optics Letters, vol. 22,
pp. 1704-1706, Nov 15 1997.
[23]
S. H. Yun, G. J. Tearney, J. F. de Boer, N. Iftimia, and B. E. Bouma, "High-speed optical
frequency-domain imaging," Optics Express, vol. 11, pp. 2953-2963, Nov 3 2003.
[24]
W. Y. Oh, S. H. Yun, B. J. Vakoc, G. J. Tearney, and B. E. Bouma, "Ultrahigh-speed optical
frequency domain imaging and application to laser ablation monitoring," Applied Physics Letters,
vol. 88, pp. -, Mar 6 2006.
[25]
S. H. Yun, G. J. Tearney, B. E.Bouma, B. H. Park, and J. F. de Boer, "High-speed spectraldomain optical coherence tomography at 1.3 mu m wavelength," Optics Express, vol. 11, pp.
3598-3604, DEC 29 2003.
[26]
W. Y. Oh, S. H. Yun, G. J. Tearney, and B. E.Bouma, "115 kHz tuning repetition rate ultrahighspeed wavelength-swept semiconductor laser," Optics Letters, vol. 30, pp. 3159-3161, DEC 1
2005.
[27]
R. Huber, D. C. Adler, and J. G. Fujimoto, "Buffered Fourier domain mode locking:
unidirectional swept laser sources for optical coherence tomography imaging at 370,000 lines/s,"
Optics Letters, vol. 31, pp. 2975-2977, Oct 2006.
[28]
D. C. Adler, R. Huber, and J. G. Fujimoto, "Phase-sensitive optical coherence tomography at up
to 370,000 lines per second using buffered Fourier domain mode-locked lasers," Optics Letters,
vol. 32, pp. 626-628, Mar 2007.
[29]
R. Huber, M. Wojtkowski, and J. G. Fujimoto, "Fourier Domain Mode Locking (FDML): A new
laser operating regime and applications for optical coherence tomography," Optics Express, vol.
14, pp. 3225-3237, Apr 17 2006.
[30]
D. C. Adler, Y. Chen, R. Huber, J. Schmitt, J. Connolly, and J. G. Fujimoto, "Three-dimensional
endomicroscopy using optical coherence tomography," Nature Photonics, vol. 1, pp. 709-716,
Dec 2007.
[31]
E.C. W. Lee, J. F. de Boer, M. Mujat, H. Lim, and S. H. Yun, "In vivo optical frequency domain
imaging of human retina and choroid," Optics Express, vol. 14, pp. 4403-4411, May 15 2006.
[32]
R. Huber, D. C. Adler, V. J. Srinivasan, and J. G. Fujimoto, "Fourier domain mode locking at
1050 nm for ultra-high-speed optical coherence tomography of the human retina at 236,000 axial
scans per second," Optics Letters, vol. 32, pp. 2049-2051, Jul 2007.
[33]
V. J. Srinivasan, D. C. Adler, Y. Chen, I. Gorczynska, R. Huber, J. Duker, J. S. Schuman, and J.
G. Fujimoto, "Ultrahigh-speed Optical Coherence Tomography for Three-Dimensional and En
Face Imaging of the Retina and Optic Nerve Head," 2008, pp. iovs.08-2127.
[34]
K. Asaka and K. Ohbayashi, "Dispersion matching of sample and reference arms in optical
frequency domain reflectometry-optical coherence tomography using a dispersion-shifted fiber,"
Optics Express, vol. 15, pp. 5030-5042, Apr 2007.
[35]
B. Povazay, B. Hermann, A. Unterhuber, B. Hofer, H. Sattmann, F. Zeiler, J. E. Morgan, C.
Falkner-Radler, C. Glittenberg, S. Blinder, and W. Drexler, "Three-dimensional optical coherence
tomography at 1050 nm versus 800 nm in retinal pathologies: enhanced performance and
choroidal penetration in cataract patients," Journal of Biomedical Optics, vol. 12, pp. -, Jul-Aug
2007.
[36]
R. Huber, K. Taira, T. H. Ko, M. Wojtkowski, V. Srinivasan, and J. G. Fujimoto, "High-speed,
amplified, frequency swept laser at 20 kHz sweep rates for OCT imaging," in Conference on
Lasers and Electro-Optics/Quantum Electronics and Laser Science and Photonic Applications,
Systems and Technologies 2005, Baltimore, 2005, p. JThE33.
[37]
M. Wojtkowski, R. Leitgeb, A. Kowalczyk, T. Bajraszewski, and A. F. Fercher, "In vivo human
retinal imaging by Fourier domain optical coherence tomography," Journal of Biomedical Optics,
vol. 7, pp. 457-463, 2002/07/ 2002.
[38]
B. Potsaid, I. Gorczynska, V. J. Srinivasan, Y. L. Chen, J. Jiang, A. Cable, and J. G. Fujimoto,
"Ultrahigh speed Spectral/Fourier domain OCT ophthalmic imaging at 70,000 to 312,500 axial
scans per second," Optics Express, vol. 16, pp. 15149-15169, Sep 15 2008.
CHAPTER 2
2.0
Principles of Fourier Domain Mode Locked Lasers
2.1 Conventional Wavelength-Swept Lasers
Swept source OCT imaging places significant performance demands on the wavelength-swept laser.
The imaging speed, axial resolution, and ranging depth of the OCT system are determined by the sweep
rate, tuning range, and instantaneous linewidth of the laser, respectively. The phase stability of the laser
also affects the flow or displacement sensitivity of phase-sensitive OCT systems. Therefore tuning ranges
of > 150 nm, instantaneous linewidths of < 0.1 nm, and sweep rates of 50 - 500 kHz are simultaneously
desired [1] with an average output power of > 10 mW. Conventional wavelength-swept laser sources
typically consist of a broadband gain medium with a tunable optical band-pass filter inside the cavity as
shown in Figure 2.1. During laser startup, lasing must build up from amplified spontaneous emission
(ASE) or fluorescence background over several cavity roundtrips. As the tunable filter switches to a new
centre wavelength position, lasing collapses at the previous position and must build up again at the new
position. The maximum achievable sweep rate is therefore limited by the characteristic time constant for
building up laser activity inside the cavity [1]. This non-stationary operation, corresponding to a
temporally varying distribution of energy between the longitudinal modes of the laser cavity, has many
drawbacks, including increased amplitude noise, low power, and broad instantaneous linewidth.
Conventional swept lasers experience decreased performance as the sweep rate is increased. Output
power begins to decrease significantly when light at a given wavelength does not make enough cavity
roundtrips to fully saturate the gain medium. This has been termed the "saturation limit" [1]. The
corresponding sweep frequency fat at the saturation limit for a laser with a sinusoidally-driven tunable
filter is given by
sat
"(2.1)
log
(2.1)
log(Gp) Af c/
atAsw L ncAlFW
f PASE
Here, G is the small-signal gain of the gain medium, p is the fraction of energy fed back into the cavity
after each round trip, AAf is the FWHM linewidth of the tunable filter, c is the free-space speed of light,
the factor 1/i corrects for the increased tuning speed (nm/s) of a sinusoidal drive compared to a linear
drive, Pa,, is the saturation power of the gain medium, AFW,,
is the full-width tuning range of the laser,
PASE is the total ASE power of the gain medium, L is the physical cavity length, and n c is the refractive
index of the cavity material. The sweep frequency at the saturation limit represents the frequency at which
the filter moves over its FWHM linewidth in a time period equal to that required for saturated lasing to
build up from ASE. A full derivation of this condition can be found in reference [1].
At sweep rates above fat , the laser output power rapidly declines until the point where light makes
only a single roundtrip in the cavity before being blocked by the tunable filter as it shifts to a new centre
wavelength. This has been termed the "single roundtrip limit" [1]. The corresponding sweep frequency
fsgi
at the single roundtrip limit for a laser with a sinusoidally-driven tunable filter is given by
Scz
gi AAFW L nc
(2.2)
At frequencies near fsngi, the laser output consists mainly of filtered ASE from the gain medium and
output power decreases by 2-3 orders of magnitude relative to the saturation limit.
The drop in power versus sweep rate for a typical swept laser constructed with a 2.4 m fiberoptic ring
cavity, a fiberoptic Fabry-Perot tunable filter (FFP-TF), and a semiconductor optical amplifier (SOA)
gain chip is shown in Figure 2.2 [1]. Normalized output power traces are shown both with and without
extra-cavity amplification by a second SOA. The two vertical lines mark the saturation limit (left line)
and the single roundtrip limit (right line). Although in this example the saturation limit occurs at -8 kHz,
it is possible to increase
fat, by
decreasing the cavity length. At extremely short cavity lengths the
spacing between cavity modes becomes wider and a more discrete stepwise tuning rather than a smooth
sweep can arise, potentially generating aliasing artifacts during OCT imaging.
Several research groups and commercial entities have investigated conventional swept lasers for OCT
imaging. One of the more successful designs uses a fixed diffraction grating and rotating polygon mirror
as the tunable filter element. Sweep rates of 15.7 kHz - 115 kHz and full-width tuning ranges of 74 - 125
nm have been demonstrated using these lasers [2-5], although performance at the higher sweep rates is
significantly degraded. Conventional swept lasers are typically restricted to sweep rates of 50 - 60 kHz
for most OCT imaging applications, which can be insufficient for acquiring densely sampled 3D datasets
in vivo [5].
2.2 FDML Theory and Typical Performance
Fourier Domain Mode Locked (FDML) lasers are a class of wavelength-swept laser that overcomes
limitations in the maximum speed rate while simultaneously providing a broad tuning range, narrow
instantaneous linewidth, and high output power. FDML lasers were originally developed by Dr. Robert
Huber in 2005 while he was a visiting scientist in Prof. James Fujimoto's group at MIT [6, 7]. FDML
lasers generate very low-noise frequency sweeps, equivalent to train of highly chirped laser pulses [7-10].
Figure 2.3 shows a schematic representation of the FDML concept. A dispersion managed delay line is
incorporated into the cavity and the filter is tuned synchronously to the cavity round-trip time (or a
harmonic of the round-trip time). This results in a quasi-stationary mode of operation. Light at one
wavelength propagates through the cavity and returns to the filter input in the same amount of time
required for the filter to tune through one cycle and return to the same wavelength position. Therefore
lasing occurs simultaneously at all wavelengths in the sweep and does not have to continuously build up
from ASE or fluorescence background. In other words, an entire wavelength sweep is optically stored
within the dispersion managed delay line in the laser cavity.
2.2.1 Theory ofFDML Operation
FDML operation requires precise synchronization between the optical roundtrip time of light in the
cavity and the tuning period of the intracavity filter [7]. This condition is given by
A rf = ncL
(2.3)
where rf is the tuning period of the tunable filter and A is a positive integer representing the harmonic
of the cavity roundtrip time. The synchronization condition can also be expressed in terms of the gating
time during which the tunable filter transmits light at a given wavelength. The required gating time rg is
given by
(2.4)
rg =
;T
fdA,FW
where fd is the filter drive frequency and f, = 2 fd is the effective sweep frequency for a sinusoidal
drive waveform. In practice rf must be synchronized to better than one part in lx10 relative to ncL /c
in order to ensure optimal operation.
Due to the stringent synchronization requirements of FDML operation, chromatic dispersion in the
long laser cavity can cause decreased performance by altering the optical roundtrip time for different
wavelengths in the sweep. Since the tunable filter must be operated at a single drive frequency it can only
be perfectly synchronized to one wavelength in the presence of dispersion, resulting in additional loss at
the desynchronized wavelengths and subsequent narrowing of the tuning range. The variation in roundtrip
time Ardisp due to the non-zero dispersion slope at wavelength A in the cavity is given as
Ardisp
= (A - 1313nm) 2 .L 0.086ps/km.nm 2
(2.5)
To minimize dispersion effects, initial FDML lasers were designed to operate at centre wavelengths near
1313 nm. This is the zero-dispersion wavelength for standard Coming SMF-28 singlemode fiber, which
can be used to form the FDML cavity and is also typically used in fiberoptic components such as couplers
and isolators. Dispersion effects are covered more completely in Section 2.4.2 below.
Under ideal operation, sequential sweeps in and FDML laser have the same phase evolution and are
mutually coherent. The filter dissipates very little energy since, at any point in time, the light incident on
the filter input contains only a narrow band of wavelengths that are matched to the transmission window
of the filter at that moment. In the frequency domain this requires destructive interference of all
longitudinal modes that are not transmitted through the narrowband filter at a given time. Thus, the phases
of the longitudinal modes must be locked. Traditional mode-locked lasers have longitudinal modes that
are locked together with constant phase, corresponding to the generation of a train of short pulses at a
repetition rate equal to the cavity round-trip time. FDML lasers have modes that are locked together with
a different phase relationship. The laser output is not a train of short pulses, but is a train of wavelength
sweeps or highly chirped, very long pulses. The tunable narrowband filtering in an FDML laser is
equivalent to an infinite number of narrowband amplitude modulators that are slightly out of phase.
Fourier domain mode locking is performed by periodic spectral modulation, rather than amplitude
modulation. This can be viewed as the Fourier domain analog of mode locking for short-pulse generation.
2.2.2 Typical FDML Performance
Figure 2.4 shows a schematic diagram of an FDML laser. The laser uses a fiber ring geometry with an
SOA as the gain medium and an FFP-TF as the tunable optical bandpass filter. The cavity is formed from
a 7 km length of SMF-28 singlemode fiber giving an optical roundtrip time of 34 pts, an FFP-TF drive
frequency of 29 kHz, and an effective sweep rate of 58 kHz. Depending on the birefringence properties of
the fiber ring, the SOA can be polarized (low cavity birefringence) or polarization-insensitive (high cavity
birefringence). Polarized SOAs generally provide higher gain, broader amplification bandwidths, and
increased output power compared to polarization-insensitive SOAs. The FFP-TF is selected to have a free
spectral range slightly greater than the desired tuning range. The FFP-TF is driven with a sinusoidal
waveform created by a high-precision digital function generator. The waveform is amplified by an
electronic power amplifier for driving the low-impedance capacitive load of the lead zirconate titanate
(PZT) FFP-TF actuator. The optical isolators (ISO) eliminate extraneous intracavity reflections and
ensure unidirectional lasing of the ring cavity. A fiber splitter acts as the output coupler with the coupling
ratio (typically 20% - 45%) controlling the tradeoff between output power and tuning range. For OCT
imaging the laser output can be further amplified with a second SOA. The inset in Figure 2.4 illustrates
the alternating series of long-to-short (backward) and short-to-long (forward) wavelength sweeps created
by the sinusoidal FFP-TF drive.
Figure 2.5 shows the transient intensity profiles of the 7 km FDML laser operating at different
harmonics of the cavity roundtrip time [7] as measured by Dr. Robert Huber during initial development of
the FDML concept. Average output power was 19 mW and the total tuning range was 105 nm. The data is
shown for the direct laser output without booster amplification in order to ensure that the transient
intensity profiles are not obscured or shaped by saturation effects [1]. In contrast to other high-speed
swept lasers, the forward and backward sweeps of the FDML laser have the same intensity profile and the
same maximum power for a wide range of sweep rates.
Since FDML lasers operate in a quasi-stationary regime, they are inherently less noisy than
conventional swept lasers. This is illustrated in Figure 2.6. Figure 2.6 (top) shows the RF interference
fringes produced by an asymmetric Michelson interferometer when an FDML and a standard swept laser
are used as the light source. The mutually-coherent sweeps in the FDML laser provide superior phase
stability, which is discussed in more detail in Chapter 3. Due to the narrower instantaneous linewidth
caused by multi-passing of the FFP-TF in the FDML laser, the OCT point spread functions shown in
Figure 2.6 (bottom) roll off more slowly than when the conventional swept laser is used. Imaging depths
of up to 7 mm are possible using FDML lasers, whereas a considerable drop in sensitivity is observed
over only 3 mm with the conventional swept laser. From the roll off of the point spread functions, a
linewidth of 0.06 nm can be calculated for the FDML laser. This is much narrower than the filter
bandwidth of 0.25 nm. Broader spectral filters can therefore be applied than in conventional swept lasers,
reducing component costs and losses in the cavity.
2.3 Buffered Cavity Designs
For OCT imaging in highly-scattering biological tissue, detection sensitivities of better than -100 dB
are typically required to ensure reasonable image quality. With a maximum incident power on the sample
of 20 mW, typical detector efficiency of 50%, and accounting for typical sample arm losses of 3 dB
(single-pass), Equation 1.1 indicates that sweep rates of I - 2 MHz are possible before OCT sensitivity
becomes prohibitively low. FDML lasers are an ideal technology to advance OCT imaging rates towards
the MHz range since they maintain performance levels as sweep rate is scaled up.
In order to increase OCT imaging rates using FDML lasers the sweep speed in nm/s must be
increased. The most straightforward way to accomplish this is to simply decrease the cavity length and
drive the FFP-TF with a higher-frequency sine wave. This approach has several drawbacks. First, the
FFP-TFs commonly used in FDML lasers exhibit strong mechanical resonances. The voltage required to
tune the FFP-TF over one free spectral range (FSR) increases dramatically at sweep rates above 200 kHz
and results in sweep instability. Second, despite the fact that both forward and backward sweep directions
have approximately the same transient power characteristics, as the effective sweep rate is increased
beyond 300 kHz the two sweep directions begin to exhibit different noise characteristics. The forward
sweep exhibits a higher noise floor, decreasing the dynamic range for OCT imaging. Finally, it is
desirable to produce unidirectional sweeps in order to reduce signal processing requirements for OCT.
With bidirectional sweeping, for example, every second interference fringe must be reversed prior to
Fourier transformation and image formation. A cavity architecture called "buffered FDML" addresses
these issues, enabling ultrahigh-speed sweeping and unidirectional operation.
2.3.1 Single-Stage Buffering
Buffered FDML lasers use a cavity design that optically replicates the low-noise backward sweep
and removes the undesired forward sweep by using a combination of time multiplexing and gain
modulation. This concept is illustrated in Figure 2.7. The interference fringes produced by an asymmetric
Mach-Zehnder interferometer are shown with the desired backward sweep circled in green and the
undesired forward sweep crossed out in red. A buffered FDML cavity schematic is shown in Figure 2.8.
Two output couplers are placed at evenly-spaced locations within the cavity. Each output coupler extracts
an optical copy of the propagating sweep, with the second coupler extracting a copy that is time-delayed
by exactly one half of the cavity round-trip time. During the time normally occupied by the forward
sweep, the intracavity SOA is modulated off using direct current modulation. The two copies of the
remaining backward sweep are combined outside of the cavity using a 50/50 fiber splitter and boosted by
an external SOA. For the configuration shown in Figure 2.8, the sweep is copied once and the layout is
referred to as "single-stage buffering." The result is a series of unidirectional, low-noise wavelength
sweeps generated at twice the FFP-TF drive frequency as shown in the inset of Figure 2.8.
The single-stage buffered FDML laser shown here was developed in collaboration with Dr. Robert
Huber as part of this thesis project [9]. The cavity length was 1.1 km, broken into two 550 m sections.
This resulted in a roundtrip time of 5.4 ps and an FFP-TF sweep rate of 185 kHz. Unidirectional
wavelength sweeps were generated at a record 370 kHz. This record has since been surpassed by
stretched-pulse laser sources operating in the MHz range [11, 12] but high excess noise levels make these
sources generally unsuitable for imaging in biological tissue. The average output power was 36 mW and
the total tuning range was 100 nm with an instantaneous linewidth of 0.1 nm, giving a peak sensitivity of
-98 dB and a rolloff of 12 dB over a 4 mm ranging depth in air. The shot noise limited sensitivity was 109 dB assuming a detector efficiency of 50%, sweep duty cycle of 90%, and 10 mW of average power
on the sample.
A comparison of dynamic range versus imaging depth for a buffered and non-buffered configuration
at effective sweep rates of 370 kHz is shown in Figure 2.9. The non-buffered FDML configuration
displays a clear degradation in noise performance of the forward wavelength sweep at large depths. With
a sensitivity of -98 dB and typical peak backreflection levels of -50 dB in biological tissue, the desired
dynamic range for most OCT applications is -50 dB. The non-buffered dynamic range falls below this
level at an imaging depth of -500
tpm for the forward sweep. The buffered laser, which produces only
backward sweeps, maintains > 50 dB dynamic range until 2000 pm imaging depths. The buffered FDML
laser was used to demonstrate OCT imaging at axial line rates of 370 kHz as shown in Figure 2.10.
2.3.2 Double-Buffering
Significant degradation in laser stability is observed when the FFP-TF is driven near a mechanical
resonance frequency, making it challenging to exceed effective sweep rates of several hundred kHz
simply by increasing the FFP-TF drive frequency. To overcome this limitation and enable stable, highspeed tuning, the FDML buffering concept can be extended to further multiply the sweep rate. Additional
optical copies are created outside of the cavity, time-delayed appropriately, and recombined. The FFP-TF
is driven near its main mechanical resonance. For the filters used in this thesis work, the main resonant
peak is near 58 kHz. Complete characterization of the FFP-TF frequency response is discussed in detail in
Chapter 5.
The concept of double-buffering is shown in Figure 2.11. To increase the sweep speed (nm/s) while
maintaining a constant FFP-TF drive frequency (Hz), the filter is overdriven with a voltage amplitude
larger than necessary to span the amplification bandwidth of the intracavity SOA. In Figure 2.11 the SOA
gain bandwidth is represented by the region between
A
and
2.
The FFP-TF drive amplitude is chosen
such that the useable wavelength range is scanned in a time equal to V2 of the normal unidirectional sweep
duration ts , where the filter drive frequency fd = 1/ 2t,. The intracavity SOA is modulated off during the
times when the filter is outside of the range A -+ A2 . Optical copies of the sweep are generated, timedelayed, and recombined to create a near-100% duty cycle train of unidirectional wavelength sweeps at
four times the FFP-TF drive frequency. Double buffering has the added benefit of decreasing the portion
of the sine wave used to generate the sweep, which improves the linearity of the sweep in optical
frequency (k) space.
Sweep copies are generated and reconstructed using a laser design constructed for this thesis work as
shown in Figure 2.12. The cavity length is 3.4 km and is broken into two sections of 1.7 km each, giving
a roundtrip time of 16.7 [ts and an FFP-TF drive frequency of 60 kHz. The FFP is driven with a -2x
amplitude overdrive as shown in Figure 2.11. Since the FSR of the FFP-TF is only 165 nm, the overdrive
results in "wrapping" of the filter transmission window back to wavelengths containing SOA gain at the
edges of the FFP-TF drive wave. This results in the creation of partial wavelength sweeps bracketing the
desired sweep. These partial sweeps are removed in the same manner as the undesired forward sweep by
modulating off the SOA. This effect, however, can be used in the future to create higher-order buffered
FDML lasers without the need for additional fiber delays or couplers as described in Section 2.3.3.
After the intra-cavity SOA modulation is activated, two time-shifted copies of the backward sweep are
extracted from the cavity using 80%/20% and 70%/30% fiberoptic couplers. The copies extracted from
the cavity are approximately 4.0 ts in duration and are time-shifted by 8.34 ts with respect to one
another. The copies are then routed to an external buffering stage consisting of an unbalanced MachZehnder interferometer with a 0.85 km path imbalance that provides a 4.17 [ts time delay. In the external
stage, two additional copies of the cavity output are created. Two are time-shifted by 4.17 [is and all four
copies are recombined in a final fiberoptic coupler. An oscilloscope screen capture of the laser output is
shown in Figure 2.13. The top trace shows interference fringes from an asymmetric Mach-Zehnder
interferometer, the middle trace shows the FFP-TF drive wave, and the bottom trace shows the intracavity
SOA modulation signal. Four nearly identical sweep copies are produced per FFP-TF drive cycle,
quadruples the effective sweep rate compared to the FFP-TF drive frequency. Although a double-buffered
FDML laser could also be constructed by breaking the cavity into 4 equal segments, the use of an external
buffering stage minimizes power losses during recombination of the sweep copies.
The double-buffered FDML laser has provided good performance for a number of OCT imaging
experiments [13]. Initial results gave an average output power of 62 mW with a total tuning range of 158
nm and a FWHM tuning range of 117 nm, supporting an OCT axial resolution of 8.3 [tm in air. The
sensitivity rolloff was -5.5 dB at an imaging depth of 2 mm in air. The time-averaged spectrum of this
laser is shown in Figure 2.15 and the OCT point spread function rolloff is shown in Figure 2.14.
Acquisition of more advanced highly-polarized SOA chips has recently boosted the output power of this
design to 80 mW and has improved total tuning range to > 180 nm, supporting axial resolutions of 7.3 [tm
in air or -5 jtm in tissue. Peak sensitivities have reached -104 dB with 29 mW of power incident on the
sample compared to a theoretical value of -111 dB after accounting for losses in the microscope sample
arm of -5 dB. Versions of this laser have been constructed for use by collaborators at the University of
Illinois at Urbana-Champagne as well as collaborators at the University of Washington.
2.3.3 Future Designs and Higher Order Buffering
To reach FDML sweep rates of 1 - 2 MHz the buffering concept can be extended by adding additional
external buffering stages. Each stage would be composed of an unbalanced Mach-Zehnder interferometer
with a fiber delay line to provide an appropriate time shift. Each additional stage can further multiply the
sweep rate without adding additional loss beyond the excess loss of the splitter and propagation loss of
the fiber. Preliminary tests indicate that the FFP-TF can tolerate at least an additional factor of 2 increase
in drive amplitude without short-term loss of stability, although no long-term testing has been conducted.
There is therefore reasonable confidence that FDML lasers with sweep rates of 500 kHz, with similar
tuning range and output power performance to the laser described above, can be readily constructed. This
is also a promising route towards breaking the 1 MHz barrier although filter ageing may be accelerated.
Since the FFP-TF is a periodic filter, overdriving the device results in "wrapping" of the transmission
window when the end of one FSR is reached. Current double-buffered designs have configured the FFPTF drive wave such that the filter tuning range is centered in the middle of the SOA gain band.
Overdriving the FFP-TF in the absence of SOA modulation gives two partial sweeps before and after the
desired central sweep due to transmission window wrapping as the filter tunes beyond its FSR. These
partial sweeps have been discarded by modulation of the SOA in the same manner that the undesired
forward sweep is removed. Future buffered designs could make use of the periodicity of the filter to
enable higher-order buffering without the addition of external buffering stages. This concept is illustrated
in Figure 2.16. Using the same FFP-TF drive amplitude as current double-buffered designs but with a
different DC voltage offset to center the sweep between two consecutive filter FSR bands, two backward
sweeps can be generated from
1/2
of the FFP-TF drive cycle. A cavity layout identical to that shown in
Figure 2.8 could be used to time-delay and recombine the sweeps. This design could be readily scaled to
produce 3 or 4 sweeps per '/2 drive cycle, leading to a 6x or 8x increase in effective sweep rate. Unlike
current double-buffered designs, however, no external time delay stages are necessary, leading to a
decrease in insertion loss and laser complexity.
2.4 Dispersive FDML Cavities and Sigma Ring Designs
The FDML laser designs discussed in previous sections have all operated at centre wavelengths near
1310 nm where chromatic dispersion in SMF-28 singlemode fiber is near zero. Wavelength-dependant
variations in fiber birefringence are also low at 1310 nm and are negligible when the cavity length is
shorter than -4 km. There are many applications, however, where operation in a dispersive regime is
required or when the cavity length must be increased in order to reduce sweep rates. For ophthalmic OCT
imaging, the laser must operate at 1060 nm or 800 nm. To use FDML sources with lower-cost
commercial OCT platforms, the sweep rate must be reduced to be compatible with moderate digitizer
sampling rates. In these situations circular cavity layouts have significant drawbacks. In both of these
examples, chromaticity in the birefringent properties of the fiber make it impossible to linearly polarize
the entire sweep prior to entering the intra-cavity SOA. A linear polarization is necessary in order to make
use of the optimal amplification performance of highly polarized SOA chips. At 1060 nm, additional
issues arise due to chromatic dispersion. This section discusses FDML operation in dispersive cavities
and introduces sigma-ring geometries to compensate for chromatic birefringence effects.
2.4.1 PolarizationChromaticity
FDML lasers exhibit unique and unusual behaviour associated with their polarization properties.
Unlike continuous wave or pulsed lasers, where the main problems with polarization management are
thermal drift effects, acoustic vibrations and changing stress in the optical components, in FDML lasers
an inherent and repeatable change of the polarization state is observed depending on the instantaneous
sweep wavelength and position inside the cavity. Immediately after the intra-cavity SOA, which is highly
polarized for optimal performance, the entire sweep is linearly polarized parallel to the gain axis of the
SOA. Following the long fiber delay and immediately prior to entering the SOA, however, the sweep
exhibits wavelength-dependant modulations in polarization state that cannot be removed with simple fiber
loop polarization rotators.
This chromatic variation in polarization is likely caused by wavelength-dependant behaviour in cavity
birefringence linked to the group delay dispersion (GDD) and polarization mode dispersion (PMD) of the
fiber. This "polarization chromaticity" is much more severe at wavelengths near 1060 nm, where
dispersion in standard single-mode fiber is high, and for low-speed FDML lasers where the cavity is
extremely long (> 4 km). Polarization chromaticity makes it impossible to use high-performance
polarized SOAs, placing heavy restrictions on performance at 1060 nm. This effect also contributes to
deviation of the sweep spectrum away from an ideal Gaussian shape due to polarization dependant loss
(PDL) in the laser components, creating sidelobes and broadened main lobes in OCT point spread
functions and degrading image quality. Reduction of polarization chromaticity effects can be achieved by
applying sigma ring cavity designs, where birefringence is cancelled by use of a Faraday device and
double-passing the cavity, as described below.
2.4.2 Dispersive FDML Theory
A series of FDML lasers operating at 1060 nm were developed as part of this thesis work in
collaboration with Dr. Robert Huber and Dr. Vivek Srinivasan [14, 15]. Operation at 1060 nm or 800 nm
is required for ophthalmic imaging due to high water absorption in the eye at 1310 nm. Ophthalmic
imaging is the most developed and widespread application of OCT, and is also an application where
sample motion is a significant limitation to image quality. Therefore an FDML laser that enables
ultrahigh-speed imaging of water-rich samples is highly desirable.
Two effects contribute added difficulty to constructing FDML lasers at 1060 nm. The first effect,
polarization chromaticity, is discussed above and can be counter-acted with a sigma-ring cavity design.
The second effect is group delay dispersion, which is significantly higher at 1060 nm than 1310 nm in
conventional singlemode fiber. Dispersive effects can be approximated by generalizing Equation (2.5) to
Ardisp = L. DAFW
(2.6)
where D is the chromatic dispersion coefficient of the fiber in ps/nm/km. The roundtrip timing mismatch
due to dispersion ATdisp must be less than the gating time zg over which the filter transmits a given
wavelength, as shown in Equation (2.4). This requirement that Adisp < g sets a lower bound on the
filter linewidth given by
A
>
c-A2 2D-z
A2w D 7
(2.7)
nC
Equation (2.7) provides several interesting insights for FDML operation within a dispersive cavity.
First, it is remarkable that to first approximation the minimum filter bandwidth required for
synchronization is independent of the cavity length. A longer cavity causes more dispersion and leads to
increased larger temporal desynchronization between the different wavelengths in the sweep. However,
the gating time of the filter rg also increases proportionally to cavity length, which largely compensates
for this effect. Second, the tuning range A/,F
occurs as a quadratic factor in Equation (2.7). This
indicates that the filter linewidth must be increased dramatically when a larger tuning range is desired. It
is important to emphasize that synchronization for only two optical roundtrips were assumed in this
analysis. For true quasi-stationary operation more effective cavity roundtrips and a longer effective cavity
photon lifetime are desired, which can be achieved by increasing the filter linewidth beyond the limit
described in Equation (2.7).
2.4.3 FDML Lasers at 1060 nm
The first 1060 nm FDML laser developed in this project used a standard circular cavity with an
external buffering stage, operating at an FFP-TF drive frequency of 118 kHz and an effective
unidirectional sweep rate of 236 kHz. The cavity was formed with 1.7 km of Coming HI-1060
singlemode fiber. FDML laser performance was characterized using three FFP-TF filters with different
bandwidths, in order to validate the theory developed above. With a dispersion value of D z 40 ps/nm/km
for HI-1060 fiber, the refractive index n = 1.46, and a desired tuning range of 80 nm, Equation (2.7) gives
a minimum filter linewidth of 0.16 nm for FDML operation. The actual obtained tuning range was 63 nm
using a filter linewidth of 0.15 nm, which supported an OCT axial resolution of 15 [tm in air.
FFP-TFs with linewidths of -0.08, -0.15, and -0.3 nm were tested and the OCT point spread
functions were measured as a function of imaging depth to assess resolution and system sensitivity, as
shown in Figure 2.17. From Equation (2.7), these linewidths should support tuning ranges of 56, 77, and
108 nm respectively. Experimentally obtained tuning ranges varied from 60 - 63 nm, indicating that
polarization effects (discussed below) were acting to limit tuning range in addition to dispersion effects.
The FDML lasers incorporating 0.08 and 0.15 nm filters show comparable sensitivity rolloff at short
delays, whereas the 0.3 nm filter has a more rapid rolloff due to a wider instantaneous linewidth. At
longer delays the 0.15 nm filter provides slightly enhanced performance compared to the 0.08 nm filter,
possibly due to increased losses after one round trip in the case of the 0.08 nm filter. These observations
support the dispersive FDML theory and verify that an optimal filter width of -0.16 nm is required for
operation near 1060 nm. This laser was used for initial imaging of the human retina at axial line rates of
236 kHz [14].
To increase the tuning range of the laser and support improved OCT axial resolutions, a second 1060
nm laser was constructed with a sigma-ring cavity configuration as shown in Figure 2.18 [15]. The cavity
length is 1.65 km and is formed by double-passing a single 825 m section of HI-1060 fiber in the linear
portion of the sigma ring, giving an FFP-TF drive frequency of 124 kHz. A fiberoptic Faraday mirror
(FM) provides a
/ 2 polarization rotation prior to the second pass of the 825 m fiber section. This
effectively cancels out all birefringence effects in the majority of the cavity except for the short circular
portion containing the FFP-TF, polarization controllers (PC), isolators, and SOA. Unidirectional
sweeping at 2x the drive frequency is obtained using an external buffering stage identical to that described
in Section 2.3.2.
Some level of chromatic dispersion compensation was obtained by introducing a Mach-Zehnder
interferometer into the cavity with a 15 cm air-path mismatch between the two arms. While the
mechanisms of action of the interferometer are not completely understood, it can be thought of as
providing two possible paths for photons traveling in the cavity. One path requires an additional 502 ps to
travel than the other, and propagating light will travel the path with the lowest total loss in a process
similar to mode competition. Light at different wavelengths will therefore self-select a path through
multiple roundtrips that optimally matches the filter tuning period, partially correcting for chromatic
dispersion effects. Incorporation of the sigma ring cavity and Mach-Zehnder dispersion compensator
resulted in a tuning range of 80 nm with a FWHM of 68 nm, compared to a previous tuning range of 63
nm with a FWHM of 38 nm as shown in Figure 2.19. The sigma ring FDML laser supports an OCT axial
resolution of 11 tm in air and produced an average output power of 3 mW.
This laser was subsequently used for human retinal imaging at a record 248 kHz [15]. Sample 2D and
3D images of a normal human retina are shown in Figure 2.20. Figure 2.20(a,b) show 16,000 axial line
images of the fovea (a) and optic nerve head (b) each acquired in 64 ms. Compared to conventional
ophthalmic OCT systems operating at 800 nm, the 1060 nm FDML laser enables improved depth
penetration into the choroid due to decreased scattering at 1060 nm. Figure 2.20(c,d) show en face
reconstructions of a 3D dataset acquired over the fovea (a) and optic nerve head (b). Retinal blood vessels
are clearly visible. In Figure 2.20(d), the lamina cribrosa is visible as the large circular feature in the
centre of the image. The lamina cribrosa is the structure where the retinal nerves connecting to individual
photoreceptors pass into the eye.
2.4.4 Low Speed BroadbandFDML at 1310 nm
Although FDML lasers enable extremely high sweep rates and OCT imaging speeds, there are many
situations when reduced speeds are desirable. One example can be found in lower-cost commercial
imaging systems. These systems, such as the Thorlabs OCS1300SS OCT platform, generally contain
lower-speed digitizer cards and data processing software that cannot sustain imaging rates in the hundreds
of kHz range. Nevertheless FDML lasers still provide important benefits in terms of low noise, broad
tuning, and alignment-free operation that can be beneficial to commercial OCT systems. Desired sweep
speeds for systems such as these are generally 20 - 40 kHz, corresponding to cavity lengths of 5 - 10 km.
Although chromatic variations in birefringence are much less severe at 1310 nm in SMF-28 fiber, when
the cavity length exceeds 4 km noticeable intra-sweep polarization modulations appear in the cavity.
These modulations cannot be removed with simple polarization rotator paddles and preclude the use of
high-gain, broadband polarized SOAs. Sigma ring cavity designs can be employed in this situation when
polarized SOAs are required in a long FDML cavity.
To enable high-performance OCT imaging with a commercial Thorlabs data acquisition and
processing platform, a sigma ring FDML laser at 1310 nm was constructed as part of this thesis work.
The cavity layout is shown in Figure 2.21. The design is similar to the 1060 nm FDML laser shown in
Figure 2.18, except the linear fiber section was changed to 4.85 km to give a total cavity length of 9.7 km
and an FFP-TF drive frequency of 21 kHz. An external buffer was used to generate unidirectional sweeps
at 42 kHz. The fiber was SMF-28 and the intra-cavity interferometer was removed since there is no
significant chromatic dispersion at 1310 nm. High-performance polarized SOAs from Covega were use
inside the cavity and for extra-cavity booster amplification.
Table 2.1 compares the performance of this laser to a buffered circular ring FDML laser with the same
cavity length using non-polarized SOA chips. As shown in
Table 2.1, the sigma ring geometry provides enhanced performance by enabling the use of polarized
SOA chips. Output power is increased by 3.6x, total tuning range is increased by 47%, and axial
resolution is improved by 29%. Sensitivity rolloff is slightly worse with the sigma ring design. It is
generally true that rolloff performance suffers as the tuning range is increased, likely due to increased
instantaneous sweep velocities (nm/s) and therefore reduced gating times at a given sweep rate (Hz).
Decreasing the filter gating time rg results in less tolerance to slight mismatches in roundtrip time due to
dispersion, lowering the number of effective cavity roundtrips at the edges of the sweep and reducing the
instantaneous linewidth. Nevertheless, sigma ring cavity designs have been demonstrated to provide
critical performance improvements both for 1060 nm FDML lasers as well as low-speed 1310 nm lasers.
Parameter
Circular Cavity
Sigma Ring
SOA Type
Non-Polarized / 42 kHz
Polarized / 42 kHz
Centre Wavelength
1285 nm
1310 nm
Average Output Power
13 mW
47 mW
Full Width Tuning Range
108 nm
148 nm
FWHM Tuning Range
75 nm
101 nm
Supported Axial Resolution (Air)
9.7 pm
7.5 pm
6 dB Sensitivity Rolloff
4.5 mm
3.7 mm
Table 2.1. Comparison of FDML performance at 42 kHz for a circular cavity with
non-polarized SOAs and a sigma ring cavity with polarized SOAs. The sigma ring
cavity provides enhanced performance.
2.5 Figures
only one
RESONATOR
or severanl
modes are
simultanously
active in the
resonator
outcoupler
*---forward scan - laser
-e-- forward scan - booster
.I .
10
..
.
d
100
- A
.
.
1000
.
I
I
,,
10000
sweep frequency [Hz]
Figure 2.1. Conceptual diagram of conventional
wavelength-swept lasers. Lasing must build up
every time the optical bandpass setting is
changed, and only one wavelength is active in
the cavity at a given time. Originally published in
[7].
Figure 2.2. Output power vs. sweep rate for
conventional wavelength-swept lasers both with
and without extra-cavity amplification by a
booster SOA. First vertical line represents the
saturation limit. Second vertical line represents
the single roundtrip limit. Output power degrades
rapidly as sweep rate is increased. Originally
published in [1].
aNmodmrd
Figure 2.3. Conceptual diagram of Fourier
domain mode-locked (FDML) lasers. The tuning
period of the optical bandpass element is
synchronized to the cavity roundtrip time. All
wavelengths are active in the cavity at all times,
removing the fundamental limitation to sweep
speed. Originally published in [7].
Figure 2.4. FDML laser schematic showing a
circular cavity with 7 km of SMF-28 fiber.
Isolators (ISO) ensure that the laser operates in
one direction around the cavity. Semiconductor
optical amplifiers (SOA) provide gain and a fiber
Fabry-Perot tunable filter (FFP-TF) is used as the
tunable optical bandpass element. Inset: FDML
output is an alternating series of long-to-short
and short-to-long wavelength sweeps when a
sinusoidal waveform is used to drive the FFP-TF.
290kHz
232kHz
116kHz
3
1
0
-8 4
30
10 20
0
-2
8
4
0
2
4
6 -3-2-10 12345
time [s]
Figure 2.5. Transient intensity profiles of an FDML laser operating at 58 - 290 kHz. The profiles remain
unchanged as the sweep rate is scaled up, unlike conventional wavelength-swept lasers that degrade
with increasing sweep rate. Originally published in [7].
FDML
no FDML
time [20nsidiv.]
time [40ns/div.]
1101
90
0i
t
1100
0
0
1
1
20
2
3
4
depth [mm]
5
a
7
1000
20
3000
doph p]
Figure 2.6. FDML lasers provide improved stability compared to conventional swept lasers, resulting in
narrower instantaneous linewidth and larger ranging depth. Top: time-lapse view of interference fringes
acquired with an FDML (left) and conventional swept laser (right). Bottom: OCT point spread functions
versus ranging depth for an FDML (left) and conventional swept laser (right).
t
to OCT
2
6
4
time (~]
Figure 2.7. Buffered FDML lasers provide
unidirectional sweeps at multiples of the cavity's
fundamental frequency. The desired backward
sweep is replicated, and the cavity SOA is turned
off during the forward sweep.
Figure 2.8. Buffered FDML laser schematic for
unidirectional sweeping at 370 kHz. The cavity is
broken into two sections to create two copies of
the backward sweep direction. Sweep copies are
recombined in an extra-cavity fiber coupler.
2
depth [mini
3
4
depth [mm]
Figure 2.9. Dynamic range versus imaging depth for an FDML laser (left) and a buffered FDML laser
(right). The forward sweep in the FDML laser has increased noise compared to the backward sweep, so
dynamic range degrades more quickly. In a buffered FDML laser, the forward sweep is eliminated and the
backward sweep is replicated, eliminating the dynamic range degradation problem. Originally published in
[16].
2 tS
t
Figure 2.10. 3D-OCT imaging at 370 kHz using a
buffered FDML laser. A: Single cross-sectional
image of human skin. 1024 lines acquired in 2.7
ms. B-D: 3D volumetric renderings of human
skin. Originally published in [16].
Figure 2.11. Double-buffering concept. The FFPTF is overdriven in amplitude to increase the
tuning speed and 4 sweep copies are
recombined to quadruple the effective sweep rate
relative to the FFP-TF drive frequency. Shaded
rectangles indicate times and wavelengths
occupied by each sweep copy.
OCT
O
<r)
ocO
C'
Figure 2.12. Double-buffered FDML laser
schematic for operation at a sweep rate of 240
kHz with a tuning range of 158 nm at a center
wavelength of 1315 nm. Sweep rate is
quadrupled by internal and external buffering
stages. Originally published in [13].
Figure 2.13. Screen capture of interference
fringes and drive waves for a double-buffered
FDML laser operating at 240 kHz. Upper trace
shows interference signal from an unbalanced
Mach-Zehnder interferometer. Middle trace
shows the FFP-TF drive waveform. 4 sweeps are
generated for each drive cycle. Bottom trace
shows the SOA modulation signal, which is ~1/4
the duration of the drive cycle.
Point Spread Functions at Various Imaging Depths
0
0
Imaging Depth [mm]
Figure 2.14. OCT point spread functions for 240
kHz double-buffered FDML laser measured at
increasing imaging depths. Sensitivity rolloff is 6
dB at 2.0 mm. Originally published in [13].
Integrated Laser Spectrum, 240 kHz Double-Buffered FDML Laser
1300
Wavelength [nm]
1400
Figure 2.15. Integrated output spectrum for
double-buffered FDML laser, with a total tuning
range of 158 nm and a FWHM bandwidth of 117
nm. Originally published in [13].
2
4
imaging depth [mm]
Figure 2.16. Proposed double-buffering concept
without external buffering stages. The FFP-TF is
overdriven and offset to produce 2 consecutive
sweeps over adjacent filter FSR bands during
each half drive cycle. Shaded rectangles indicate
times and wavelengths occupied by each sweep
copy.
Figure 2.17. OCT point spread functions for
various imaging depths, obtained using an FDML
laser at 1060 nm with three different tunable filter
linewidths (0.08, 0.15, and 0.3 nm). As filter width
increases, a larger tuning range is possible but
the instantaneous linewidth is increased and
imaging range decreases. Originally published in
[14].
0.8
825 m
P
I
0.6
to
OCT
S= 68.32v
S 38
i
0.2
A
100
825 m
Figure 2.18. 1060 nm FDML laser with sigma
ring for polarization chromaticity compensation,
intracavity Mach-Zehnder interferometer for
dispersion compensation, and external buffering.
Sweep rates of 249 kHz with a tuning range of 80
nm are achieved.
1020
1040
ION
1080
1100
1120
wavobr4$ [nm]
Figure 2.19. Integrated output spectra for 1060
nm sigma-ring FDML (FWHM = 68 nm) and
circular cavity FDML (FWHM = 38 nm). Sigma
dispersion
intracavity
with
design
ring
range
tuning
improves
significantly
compensator
and supported OCT axial resolution.
Figure 2.20. Images of the human retina acquired with a 1060 nm FDML laser. A, high-density crosssection through the fovea. B, high-density cross-section through the optic nerve head. C, en face
reconstruction of the fovea. D, en face reconstruction of the optic nerve head. Originally published in [15].
ISO
4.85 km
Figure 2.21. 1310 nm FDML laser with sigma ring for polarization chromaticity compensation and
external buffering stage. Sweep rates of 42 kHz with a tuning range of 148 nm are achieved using
polarized SOA devices.
2.6 References
[1]
R. Huber, M. Wojtkowski, K. Taira, J. G. Fujimoto, and K. Hsu, "Amplified, frequency swept
lasers for frequency domain reflectometry and OCT imaging: design and scaling principles,"
Optics Express, vol. 13, pp. 3513-3528, May 2 2005.
[2]
S. H. Yun, G. J. Teamey, J. F. de Boer, N. Iftimia, and B. E. Bouma, "High-speed optical
frequency-domain imaging," Optics Express, vol. 11, pp. 2953-2963, Nov 3 2003.
[3]
W. Y. Oh, S. H. Yun, B. J. Vakoc, G. J. Tearney, and B. E. Bouma, "Ultrahigh-speed optical
frequency domain imaging and application to laser ablation monitoring," Applied Physics Letters,
vol. 88, pp. -, Mar 6 2006.
[4]
S. H. Yun, G. J. Tearney, B. J. Vakoc, M. Shishkov, W. Y. Oh, A. E. Desjardins, M. J. Suter, R.
C. Chan, J. A. Evans, I. K. Jang, N. S. Nishioka, J. F. de Boer, and B. E. Bouma, "Comprehensive
volumetric optical microscopy in vivo," Nature Medicine, vol. 12, pp. 1429-1433, Dec 2006.
[5]
B. J. Vakoc, M. Shishko, S. H. Yun, W. Y. Oh, M. J. Suter, A. E. Desjardins, J. A. Evans, N. S.
Nishioka, G. J. Tearney, and B. E. Bouma, "Comprehensive esophageal microscopy by using
optical frequency-domain imaging (with video)," GastrointestinalEndoscopy, vol. 65, pp. 898905, May 2007.
[6]
R. Huber, K. Taira, M. Wojtkowski, and J. G. Fujimoto, "Fourier Domain Mode Locked Lasers
for OCT imaging at up to 290kHz sweep rates," in Optical Coherence Tomography and
Coherence Techniques II, Munich, 2005, pp. 245-250.
[7]
R. Huber, M. Wojtkowski, and J. G. Fujimoto, "Fourier Domain Mode Locking (FDML): A new
laser operating regime and applications for optical coherence tomography," Optics Express, vol.
14, pp. 3225-3237, Apr 17 2006.
[8]
D. C. Adler, R. Huber, and J. G. Fujimoto, "Phase-sensitive optical coherence tomography at up
to 370,000 lines per second using buffered Fourier domain mode locked lasers," Opt Lett, vol. 32,
pp. 626-628, 2007.
[9]
R. Huber, D. C. Adler, and J. G. Fujimoto, "Buffered Fourier Domain Mode Locking (FDML):
Unidirectional swept laser sources for OCT imaging at 370,000 lines per second," Optics Letters,
vol. 31, pp. 2975-2977, October 15, 2006 2006.
[10]
R. Huber, D. C. Adler, V. J. Srinivasan, and J. G. Fujimoto, "Fourier domain mode locking at
1050 nm for ultrahigh-speed optical coherence tomography of the human retina at 236,000 axial
scans per second," Optics Letters, vol. 32, p. In Press, 2007.
[11]
S. Moon and D. Y. Kim, "Ultra-high-speed optical coherence tomography with a stretched pulse
supercontinuum source," Optics Express, vol. 14, pp. 11575-11584, Nov 2006.
[12]
Y. Park, T. J. Ahn, J. C. Kieffer, and J. Azana, "Optical frequency domain reflectometry based on
real-time Fourier transformation," Optics Express, vol. 15, pp. 4597-4616, Apr 2007.
[13]
D. C. Adler, S. W. Huang, R. Huber, and J. G. Fujimoto, "Photothermal detection of gold
nanoparticles using phase-sensitive optical coherence tomography," Optics Express, vol. 16, pp.
4376-4393, Mar 31 2008.
[14]
R. Huber, D. C. Adler, V. J. Srinivasan, and J. G. Fujimoto, "Fourier domain mode locking at
1050 nm for ultra-high-speed optical coherence tomography of the human retina at 236,000 axial
scans per second," Optics Letters, vol. 32, pp. 2049-2051, Jul 2007.
[15]
V. J. Srinivasan, D. C. Adler, Y. Chen, I. Gorczynska, R. Huber, J. Duker, J. S. Schuman, and J.
G. Fujimoto, "Ultrahigh-speed Optical Coherence Tomography for Three-Dimensional and En
Face Imaging of the Retina and Optic Nerve Head," 2008, pp. iovs.08-2127.
[16]
R. Huber, D. C. Adler, and J. G. Fujimoto, "Buffered Fourier domain mode locking:
unidirectional swept laser sources for optical coherence tomography imaging at 370,000 lines/s,"
Optics Letters, vol. 31, pp. 2975-2977, Oct 2006.
CHAPTER 3
3.0
Phase Sensitive OCT Using FDML Lasers
3.1 Motivation
Contrast in conventional OCT images results from measuring the amplitude of the interference signal
formed by a reference field and sample field in a Michelson interferometer. It is also possible to extract
the phase of the interference signal, giving an alternative contrast modality that complements the standard
amplitude information. Differential (axial line-to-line) phase information can be used to perform Doppler
flow OCT [1-6] by detecting variations in interference fringe phase at the same spatial location over short
periods of time. It is also possible to map the absolute interference fringe phase relative to a fixed
reference surface, enabling the detection of picometer-scale surface displacements or variations in optical
path length [7]. These techniques are commonly referred to as OCT phase microscopy or, more generally,
phase-sensitive OCT [8-15]. In both cases, small changes in the phase of the interference signal can be
correlated to small, sub-resolution changes in the location of backscattering or backreflecting surfaces
within the sample. These changes are not detectable using conventional OCT analysis since the phase
information is discarded during image formation.
The combination of Fourier domain detection and phase sensitive analysis is very attractive, since it
enables phase-based contrast at extremely high imaging speeds. The displacement sensitivity of a phasesensitive OCT system is a specification of the smallest variation in optical path length that is detectable
by measuring the interference fringe phase relative to a fixed reference surface. The smallest possible
displacement sensitivity is desired for detecting low flow rates or small variations in optical path length.
Rapid imaging speeds are simultaneously desired to minimize motion artifacts, avoid fringe averaging
and 27r phase wrapping artifacts, and enable detection of rapid flows in Doppler OCT applications.
OCT phase microscopy systems have been previously reported based on both spectral domain [8-10]
and swept source [11] detection. Doppler OCT systems have also been reported using both spectral
domain [12-14] and swept source [15] detection. The spectrometer-based systems have provided excellent
displacement sensitivities (18 pm at 30 Hz imaging rates [13] or 25 pm at 29 kHz imaging rates [9]) but
are limited in speed due to the readout rates of the CCD cameras, limiting the largest flow rates and path
changes that can be observed. Prior swept source systems have produced poorer displacement sensitivities
(1300 pm at 12 kHz [11] or 475 pm at 16 kHz [15]) due to the low sweep-to-sweep phase stability of
conventional swept wavelength laser sources. As part of this thesis work, the phase stability properties of
FDML lasers were characterized and compared to conventional swept lasers. Differences in buffered
versus non-buffered FDML cavities were observed, along with differences between buffered FDML
lasers operating at different sweep rates. High-speed phase sensitive OCT was also demonstrated at axial
line rates of a record 117 kHz.
3.2 Phase Stability Measurements of FDML Lasers
As discussed above, phase sensitive OCT imaging requires a light source with extremely high phase
stability since phase jitter in the laser can easily drown out the phase changes associated with variations in
optical path length of the sample. Simultaneously, high-speed imaging is desirable in order to minimize
parasitic sample motion and maximize the range of detectable path changes. FDML lasers operate in a
quasi-stationary regime where each wavelength in the sweep has many characteristics of a narrowband
continuous-wave laser, including high sweep-to-sweep phase stability. The phase of the interference
signal generated during OCT imaging is analyzed by Fourier transforming the fringes and then evaluating
the phase q (z), as opposed to the amplitude A(z), of the backscattered light as a function of sample
depth.
3.2.1 Experimental Setup
The experimental setup for performing phase sensitive OCT measurements is shown in Figure 3.1.
95% of the power from the laser source enters a common path Michelson interferometer. In a common
path topology, the front surface of a 210 [tm thick glass cover slip provides the reference reflection for the
interferometer. A sample, such as a biological cell or photonic device, can be placed on the back surface
of the coverslip for interrogation. Common path setups are commonly used in phase sensitive OCT
systems to prevent phase jitter between the separate reference arm and sample arm in a standard
Michelson interferometer. 5% of the laser output is routed to a second common path interferometer,
which is used to record and calibrate the slow phase drift of the FDML lasers. This step is required since
the phase noise of FDML lasers consists of a low amplitude (milliradian) white Gaussian component and
a larger amplitude (radian) component that drifts slowly. The slow component typically has a period of 1
- 5 ms and is likely caused by thermal drift in the fiber Fabry-Perot tunable filter (FFP-TF) element.
Digitization is carried out by a 5 GSample/s 8-bit digital oscilloscope (for displacement sensitivity
measurements) or a 200 MSample/s 14-bit digitizer card (for imaging experiments). A personal computer
is used for resampling the interference fringes onto a uniform (k) spacing, Fourier transformation, phase
extraction, and calibration.
Displacement sensitivities Zmin using FDML and conventional swept lasers were measured by
recording the phase of the interference signal b(z) originating from the back surface of the 210 gm
coverslip (z = z o = 210 pm ) in the sample arm relative to the front surface of the coverslip. The phase
0 (z, ) was measured continuously over -3 ms. The standard deviation of this measurement Uc((zo,
was used to calculate displacement sensitivity as [13]
t))
Zmin =
4fn
(3.1)
where A0 is the laser centre wavelength in vacuum and n is the refractive index of the sample material.
Differential displacement sensitivity was also measured using each laser source. Differential displacement
sensitivity is a measurement of the minimum path length variation that can be detected at the same axial
position over two consecutive axial lines. This specification is useful to estimate Doppler OCT
performance, since Doppler processing requires differential measurements across time to detect flow.
Differential displacement sensitivity
Azmin
Az
where A0(zo,t)= (zo, t + At) -
is given by
=
"-
(zo, t))
o(A
o47rn
(3.2)
(3.2)
(zo, t) and At is the laser sweep period.
It is also possible to calculate the signal-to-noise (SNR) limited displacement sensitivity, which allows
comparison of each laser's performance to the theoretical optimum value. SNR-limited displacement
sensitivity zSNR is given by [13]
ZSNR -
n SNR
(3.3)
where SNR is measured at z = z o = 210pm and is given by
SNR =
A (zo)2
var (A (z))
(3.4)
with A (z) the amplitude of the OCT interference fringe.
3.2.2 Data Processing
Since the phase noise of FDML lasers is extremely low, great care must be taken to avoid introducing
phase noise artifacts during data processing. Significant phase artifacts can arise from data segmentation,
Fourier transformation, and phase extraction. To prevent phase jitter from the PC digitizer board's sample
clock and record-to-record rearming time jitter, displacement sensitivity measurements were performed
using the 5 GSample/s digital oscilloscope. While rearming jitter is eliminated by storing many
interference fringes back-to-back in a single 3 ms long record, individual fringes must be segmented and
extracted from the record prior to Fourier transformation. The segmentation technique is a critical data
processing step that can easily induce phase errors on the order of the actual FDML laser phase noise.
With typical FDML phase noise levels less than 1 mrad at an interference signal frequency of- 10 MHz,
for example, data segmentation errors on the order of 100 ps will add considerably to the measured phase
noise. 100 ps corresponds to 0.5 samples of a 5 GSample/s signal, indicating a need for custom
segmentation software.
The data processing flowchart for displacement sensitivity measurements is shown in Figure 3.2. Steps
where manual input is required are shown as polygons whereas automated steps are shown as rectangles.
Parallelograms indicate steps where data is created or manipulated. Additional figures illustrating the data
processing procedure are shown in Figure 3.3 to Figure 3.8. These figures were generated using data
acquired with a 21 kHz buffered FDML laser operating at an effective sweep rate of 42 kHz. First, the
entire record including the timebase and sample values from the 5 GSample/s oscilloscope is read into
Matlab and converted to two double precision vectors. Both the time vector and sample value vectors are
roughly 1x10 7 points long and contain 40 - 7400 laser sweeps, depending on the sweep rate. Since the
record may start and end partway through a laser sweep, the points containing partial sweeps are removed
as shown in Figure 3.3.
Next a new timebase is created that will allow perfect segmentation of each sweep. In the original
timebase of the oscilloscope, there are generally not an integer number of samples in each laser sweep
period. Simple segmentation of the original record would therefore lead to an accumulating error in sweep
duration, causing phase walkoff and single-point segmentation jumps that will corrupt the noise
measurement. To avoid this, the processing software calculates the smallest number of samples that
would be required in the entire record to give an integer number of samples per sweep after segmentation.
This requires knowledge of the exact FFP-TF drive frequency
fd
(which is input by the user), the
oscilloscope sampling period At, (which is derived from the timebase vector), and the number of
buffering stages in the laser B (if any) to calculate the effective sweep rate. The desired number of
samples in the long record Ns, is calculated as
N,, = Nces round
where Ncycles = floor (NM0o f . At)
lfd
At,
(3.5)
is the number of complete FFP-TF drive periods contained in the
long record and Nso is the original number of samples in the long record.
Once the new timebase is established, the long record is broken up into 4 equal segments in order to
avoid memory overflow issues in Matlab. Each segment is then linearly interpolated onto the new
timebase and the long record is broken up into segments corresponding to one sweep period. Each new
segment is exactly of length N,, / Ncycles . At this point the user must check the segmentation results by
examining the first sweep and the last sweep. The sweeps should overlap perfectly to avoid phase walkoff
or k-space recalibration errors. If segmentation drift is observed, the user must adjust the FFP-TF drive
frequency value and repeat steps 3 - 8 as shown in Figure 3.2.
After segmentation, the user must strip the unused "dead points" from the beginning and end of each
sweep as shown in Figure 3.4. The process of k-space recalibration can now begin. Here, a recalibration
vector is used to resample each interference fringe onto a timebase that is uniformly spaced in optical
frequency k. If the laser under test is a buffered FDML laser, each independent path through the buffering
stages will generate a slightly different phase evolution. One recalibration vector must therefore be
generated for each optical copy that is created (ie, a single-buffered laser with two output couplers in the
cavity requires two recalibration vectors, whereas a double-buffered laser requires four recalibration
vectors).
To generate low-noise fringes for recalibration, all of the fringes are grouped according to buffered
output index and then averaged. A 6 th order polynomial fit is subtracted from the averaged fringes in order
to reduce the asymmetry about 0 V arising due to unbalanced photodetection. This step is necessary to
reduce phase artifacts in the subsequent Hilbert transform, which is used to extract instantaneous phase
versus time curves for each calibration fringe. A typical phase versus time curve is shown as the blue
trace in Figure 3.5. A
5 th
order polynomial fit is then matched to the instantaneous phase curve. Each
sample of this polynomial is proportional to the k value of each corresponding sample in the original
interference fringes. A linear phase vector is generated through the minimum and maximum values of the
polynomial as shown in Figure 3.5. This linear curve represents the desired linear k spacing of each
fringe. Spline interpolation is used to resample each segmented interference fringe onto the linear k vector
assuming initial k values defined by the polynomial. The resulting fringes have samples that are perfectly
linearly spaced in k, giving optimal point spread functions for OCT imaging and preventing phase
distortions as the ranging depth is varied.
After recalibration each interference fringe is windowed with a Hanning function and Fourier
transformed using a fast Fourier transform (FFT) algorithm. The FFT is not zero-padded and does not
provide sufficient axial sampling density to accurately track the measured phase over long time periods,
since a single point change in the FFT peak location can cause significant phase spikes. The FFT is used
only to generate a rough axial amplitude trace of the sample in order to locate the coverslip's rear surface.
Following identification of the rear surface, a chirped Z transform (CZT) is used to perform a finelysampled frequency transform centered around the rear surface. The CZT typically contains 213 points and
spans a range of +/- 2x the axial position of the coverslip reflection. This gives sufficient axial sampling
density to accurately extract phase over long periods of time.
The sample phase is measured at the CZT peak corresponding to the coverslip's rear surface in each
interference fringe as shown in Figure 3.6. The large peak at sample 2000 is the coverslip surface, while
the small peak at sample 4000 is an artifact due to path length ambiguity in the common path geometry.
Once the phase has been extracted for each fringe in the dataset, it can be plotted as a function of sweep
number as shown in Figure 3.7. If the laser is a buffered FDML configuration, the slight mismatch in
optical path length between the two laser output paths results in a small (-2 mrad in this case) constant
phase offset between consecutive measurements. This can be corrected by separating the measurement
into even and odd data points and then setting the mean of each set to zero. The slight linear component to
the phase profile in Figure 3.7 is caused by miniscule segmentation error that slowly builds up over each
sweep. This can also be removed by subtracting a linear fit from the data.
The final phase measurement is shown in Figure 3.8. This measurement is obtained after subtracting
the lowpass-filtered calibration fringe from the separate common-path calibration arm shown in Figure
3.1. The calibration fringe is processed in an identical manner to that described above, except it is
lowpass-filtered at the end to remove all phase noise and leave only the long-term phase drift from the
laser. The measured standard deviation from the sample fringe is 0.617 mrad over 123 sweeps (3 ms),
corresponding to a displacement sensitivity of 42.1 pm in glass with a refractive index of 1.5 at a centre
wavelength of 1285 nm. Identical measurements were made, as described below, for a series of lasers
including conventional swept sources, non-buffered FDML lasers, and buffered FDML lasers operating at
a range of sweep rates.
3.2.3 Conventional Swept Laser
To serve as a control in this experiment, a conventional wavelength-swept laser was constructed using
a fiber ring, semiconductor optical amplifier (SOA) gain chip, and FFP-TF. The total tuning range was
128 nm at a centre wavelength of 1285 nm with an average output power of 14 mW. The FFP-TF was
operated at a drive frequency of 1 kHz, with forward and backward sweeps giving an effective sweep rate
of 2 kHz. Output power, tuning range, and phase noise increased significantly at speeds above 2 kHz as
described in Chapter 2.
The phase noise of the phase sensitive OCT system was measured to be 3.32 mrad using the method
described in Section 3.2.1 and Section 3.2.2. This corresponds to a displacement sensitivity of 226 pm,
which is comparable to previously-reported displacement sensitivities using conventional swept lasers
[11, 15]. The SNR at the back surface of the coverslip was 77 dB, giving an SNR-limited displacement
sensitivity of 9 pm. The 25x difference between the actual sensitivity and the SNR limit highlights a
major shortcoming of conventional swept lasers. Non-stationary operation leads to greatly increased
phase noise, lowering the utility of these sources for phase sensitive OCT. Differential displacement
sensitivity was 361 pm, which is roughly a factor of 2 more than the single-line displacement
increases
sensitivity. This is to be expected, since differential measurements of noisy signals result in in measurement uncertainty.
3.2.4 Non-Buffered FDML Laser
A non-buffered FDML laser was constructed by adding a 9.7 km length of SMF-28 optical fiber to the
cavity used for the conventional swept laser in Section 3.2.3. The same SOAs, FFP-TF, and fiber output
couplers were used in both cases. The FDML tuning range was 112 nm and the average output power was
13 mW at a center wavelength of 1285 nm. The FFP-TF drive frequency was 21 kHz and the effective
sweep rate was 42 kHz using both forward and backward sweeps. The measured phase noise
r (A
(z, t)) was 1.04 mrad, corresponding to a displacement sensitivity of 71 pm. The SNR at the
back surface of the coverslip was 68 dB, giving an SNR-limited displacement sensitivity of 26 pm. As
expected the differential displacement sensitivity was 115 pm, approximately a factor ofxT
higher than
the displacement sensitivity.
These results illustrate several important features of FDML lasers. First, the quasi-stationary FDML
operation leads to significant improvements in phase noise and displacement sensitivity compared to
conventional swept lasers. The 42 kHz FDML laser obtained a >3x improvement in displacement
sensitivity compared to the conventional swept source, even though the FDML laser was operating at a
41x higher effective sweep rate. Second, the displacement sensitivity of the FDML laser comes within
2.7x of the theoretical SNR limit compared to 25x for the conventional swept laser. This is further
evidence that the excess phase noise of FDML lasers is much lower than other swept sources.
3.2.5 Buffered FDML Lasers
To investigate differences between buffered and non-buffered FDML lasers, a series of buffered
cavities were constructed with fiber lengths of 9.7 km, 3.5 km, and 1.1 km. Effective unidirectional sweep
rates were 42 kHz, 117 kHz, and 368 kHz respectively. In all cases the tuning range was > 110 nm and
the average output power was > 11 mW at a centre wavelength of 1285 nm. The 42 kHz buffered laser
produced a phase noise reading of 0.57 mrad. The displacement sensitivity, SNR-limited displacement
sensitivity, and differential displacement sensitivity were 39 pm, 22 pm, and 43 pm respectively.
Again, several interesting observations can be made by comparing the buffered and non-buffered 42
kHz FDML laser results. Even though the SNR-limited displacement sensitivities were similar for the
buffered and non-buffered lasers (22 pm and 26 pm respectively), the measured displacement sensitivity
for the buffered laser was a factor of 1.8x smaller than the non-buffered laser at the same sweep rate.
Additionally, the differential displacement sensitivity for the buffered laser was within 10% of the
displacement sensitivity. The expected
-.2 increase in noise was not observed in the buffered case. This
can be explained by the fact that, in buffered lasers, each output coupler produces a time-delayed optical
copy of the propagating sweep at evenly-spaced positions in the cavity. The propagating sweep does not
undergo amplification or filtering between the output couplers, so the phase correlation between these
pair of consecutives sweeps is preserved. This improves the single-measurement displacement sensitivity
and also the differential displacement sensitivity, which could significantly enhance Doppler OCT
measurements in the future. Similar improvements should be observed in double-buffered FDML lasers.
Results for the 117 kHz and 368 kHz buffered FDML lasers, along with the rest of the lasers tested in
this section, are shown in Table 3.1. As the sweep rates of the buffered FDML lasers were increased by
shortening the cavity length, the displacement sensitivity degraded moderately from 39 pm at 42 kHz to
102 pm at 368 kHz. This compares favorably with previously reported displacement sensitivities of 25
pm at 29 kHz for spectrometer-based phase sensitive OCT systems [9]. However, buffered FDML lasers
enable significant increases in data acquisition rate far beyond the speed limits of spectrometer-based
systems. The increased speeds possible with buffered FDML lasers could also be used to perform data
averaging, further improving displacement sensitivity. In profilometry applications, displacements of a
single surface are measured. In this application, buffered FDML lasers would be capable of measuring
displacements in a continuous range between the minimum displacement sensitivity (39 - 102 pm) and
the laser coherence length (> 4 mm), corresponding to roughly 8 orders of magnitude.
The buffered FDML lasers operating at 42 kHz - 368 kHz come within 1.1x - 2.0x of their respective
SNR-limited displacement sensitivities. All of the buffered lasers also provided differential displacement
sensitivities that are within -10%
of the displacement sensitivity, confirming that buffered cavities
provide improved sweep-to-sweep phase correlation through extraction of optical sweep copies. Even at a
sweep rate of 368 kHz, 184x higher than the conventional swept laser, the buffered FDML still produced
>2x lower phase noise and displacement sensitivity than the conventional source. These observations
indicate that buffered FDML lasers are ideal sources for high-speed phase-sensitive OCT applications.
Laser Type Used
Swept Laser (2 kHz)
FDML (42 kHz)
Buffered FDML (42 kHz)
Buffered FDML (117 kHz)
Buffered FDML (368 kHz)
Displ. Sens.
[pm]
226
71
39
52
102
SNR-Limited
Displ. Sens. [pm]
9
26
22
38
50
Differential
Displ. Sens. [pm]
361
115
43
56
119
Table 3.1. Displacement sensitivity, SNR-limited displacement sensitivity, and differential
displacement sensitivity using phase-sensitive OCT with a conventional swept laser, nonbuffered FDML laser, and buffered FDML lasers at several effective sweep rates. Buffered
FDML lasers provide superior phase-sensitive performance at the highest possible sweep
rates.
3.3 Sub-Nanometer Dynamic Sensing
One potential application of high-speed phase-sensitive OCT is to capture small, rapid, transient
events to characterize MEMS devices. The buffered FDML laser running at 117 kHz was used to
demonstrate dynamic phase-sensitive OCT measurements that may be similar to those required for
MEMS characterization. A gold mirror was mounted to a lead-zirconate-titanate (PZT) piezoelectric
transducer positioned 1 mm behind a 3 mm glass coverslip in the sample arm. A 3 mm coverslip was used
so that the back surface generated a reference reflection while the front surface was outside the beam
focus to prevent unwanted interference. The incident beam waist was 15 jim at the 1/e intensity point,
giving a Rayleigh length of 138 jim. The PZT was sinusoidally actuated at 5 kHz, while the phase at a
single point on the mirror was monitored over a time period of 3 ms.
Figure 3.9 shows the measured PZT displacement versus time. The OCT interference fringes were
processed in the same manner as described in Section 3.2.2, except that the measured phase at the gold
mirror relative to the glass reference surface was plotted as a function of time as the final step. Removal
of the residual linear phase (Step 23 in Figure 3.2) would also remove any linear drift in the mirror
position relative to the reference surface, which is not of interest for this experiment. The 5 kHz PZT
motion is clearly visible in Figure 3.9 as a periodic modulation at +/- 3 nm amplitude. The excellent phase
stability of the laser yields a low noise measurement of displacement. The slow modulation at -500 Hz is
likely due to motion of the PZT mount relative to the reference surface. This result demonstrates that
phase sensitive OCT with buffered FDML lasers can resolve nanometer-scale surface displacements in
microsecond time scales.
3.4 Sub-Nanometer 3D Phase Profilometry
Another application of high-speed phase-sensitive OCT is to perform 3D measurements of sample
optical path length with sub-nanometer precision. This capability, which can be referred to as 3D-OCT
phase microscopy, could be useful for material characterization, cellular biology studies, or industrial
inspection. 3D-OCT phase microscopy was performed on a 210 jtm glass coverslip by scanning the
incident beam over a 1 mm x 1 mm area using a pair of galvanometers. The 3D dataset consists of 230 x
230 axial scans and was acquired in 0.45 s. The phase of the back surface of the coverslip relative to the
front surface was measured and displayed as a physical path difference in a false color image. Again, data
processing was performed in the same manner as described in Section 3.2.2, except that the measured
phase of each axial line was mapped into a spatial location according to the corresponding galvanometer
position.
The results of this imaging experiment are shown in Figure 3.10. The exceptionally low phase noise of
the system results in good displacement sensitivity, enabling visualization of sub-nanometer variations in
optical path length. Surface defects consistent with microgrooves are visible on the left hand side of the
coverslip. Larger, more homogenous variations in optical path length are visible on the right hand side of
the coverslip. These features may be variations in physical sample thickness (+/- 2 nm) or refractive index
(+/- 1.4 x 10 ). Small circular features consistent with pockmarks are scattered throughout the sample.
This result demonstrates that phase sensitive OCT with buffered FDML lasers can perform 3D surface
measurements with nanometer sensitivities at speeds significantly higher than spectrometer-based
systems.
3.5 Figures
M
TRIG
CHI CH2
D Conversion
ADb Convron
slide
Calibration Arm
PD
Amp
2 10 um
Amp
repsampl
FFT
phase extraction
* phase calibration (CH 1 - Ch 2)
Figure 3.1. Common-path Michelson interferometer used for phase-sensitive OCT.
Combining the sample and reference paths into a single path reduces phase noise and
improves displacement sensitivity. Originally published in [16].
VECTOR
TRAES
Figure 3.2. Flowchart showing data processing steps for displacement sensitivity
measurements and phase-sensitive OCT imaging.
is
0
10
20
30
40
70
60
so
90
80
of
Figure 3.3. Data processing Step 2. Partial sweeps are removed from the start and end
record.
ms
2
a
of
us
90
first
the oscilloscope record. Figure shows
5
Wo
The [us7-
Figure 3.4. Data processing Step 10. Unusable
points are removed from each sweep following
segmentation. Figure shows one sweep that has
been segmented out from the 2 ms record.
Figure 3.5. Data processing Steps 14-16. Phase
vs. time curves for the original calibration trace
(blue) and desired linear phase curve (red).
IO
FPm PMe
Figure 3.6. Data processing Step 21. Sample
is extracted at the FFT peak
phase
corresponding to the glass coverslip's rear
surface for each interference fringe.
20
16
10
20
40
so
80
100
120
140
Swep Numbr
Figure 3.7. Data processing Steps 22-23. Phase
is corrected for path length imbalances in
buffered FDML lasers and residual linear phase
0
10
_.2
4
0 20
40
00
o80 100
120
S
14
Sweep Number
phase
Final
22-24.
Steps
Figure 3.8.
measurement shows structureless noise with a
standard deviation of 0.617 mrad, giving a
displacement sensitivity of 42 pm.
0.6
1
1.6
Tim [ms]
2
2.6
Figure 3.9. Dynamic measurement of PZT
transducer motion using phase-sensitive OCT
system and 117 kHz buffered FDML laser. PZT
was driven at 5 kHz over +/- 3 nm. Originally
published in [16].
3D OCT Phase Microscopy Innage of 210um Glass Coverslip
0
100
200
Aoo3WO
S400
0
S 600
-2
.4
400
3
600
1000
-10
X Distance [uMn]
Figure 3.10. 3D-OCT phase microscopy image of a glass coverslip performed using a 117
kHz buffered FDML laser. Sub-nanometer variations in optical path length are visible as
small striations, large modulations, and pockmarks. Originally published in [16].
3.6 References
[1]
S. Yazdanfar, M. D. Kulkarni, and J. A. Izatt, "High resolution imaging of in vivo cardiac
dynamics using color Doppler optical coherence tomography," Optics Express, vol. 1, 1997/12/22
1997.
[2]
Z. Chen, T. E. Milner, D. Dave, and J. S. Nelson, "Optical Doppler tomographic imaging of fluid
flow velocity in highly scattering media," Optics Letters, vol. 22, pp. 64-6, 1997/01/01 1997.
[3]
Z. Ding, Y. Zhao, H. Ren, J. S. Nelson, and Z. Chen, "Real-time phase-resolved optical
coherence tomography and optical Doppler tomography," Optics Express, vol. 10, 2002/03/11
2002.
[4]
V. X. D. Yang, M. L. Gordon, B. Qi, J. Pekar, S. Lo, E. Seng-Yue, A. Mok, B. C. Wilson, and I.
A. Vitkin, "High speed, wide velocity dynamic range Doppler optical coherence tomography
(Part I): System design, signal processing, and performance," Optics Express, vol. 11, pp. 794809, Apr 7 2003.
[5]
V. X. D. Yang, M. L. Gordon, E. Seng-Yue, S. Lo, B. Qi, J. Pekar, A. Mok, B. C. Wilson, and I.
A. Vitkin, "High speed, wide velocity dynamic range Doppler optical coherence tomography
(Part II): imaging in vivo cardiac dynamics of Xenopus laevis," Optics Express, vol. 11,
2003/07/14 2003.
[6]
V. X. D. Yang, M. L. Gordon, S. J. Tang, N. E. Marcon, G. Gardiner, B. Qi, S. Bisland, E. SengYue, S. Lo, J. Pekar, B. C. Wilson, and I. A. Vitkin, "High speed, wide velocity dynamic range
Doppler optical coherence tomography (Part III): in vivo endoscopic imaging of blood flow in the
rat and human gastrointestinal tracts," Optics Express, vol. 11, pp. 2416-2424, Sep 22 2003.
[7]
M. Sticker, M. Pircher, E. Gotzinger, H. Sattmann, A. F. Fercher, and C. K. Hitzenberger, "En
face imaging of single cell layers by differential phase-contrast optical coherence microscopy,"
Optics Letters, vol. 27, pp. 1126-8, 2002/07/01 2002.
[8]
M. A. Choma, A. K. Ellerbee, C. H. Yang, T. L. Creazzo, and J. A. Izatt, "Spectral-domain phase
microscopy," Optics Letters, vol. 30, pp. 1162-1164, May 15 2005.
[9]
C. Joo, T. Akkin, B. Cense, B. H. Park, and J. E. de Boer, "Spectral-domain optical coherence
phase microscopy for quantitative phase-contrast imaging," Optics Letters, vol. 30, pp. 21312133, Aug 15 2005.
[10]
H. Li, B. A. Standish, A. Mariampillai, N. R. Munce, Y. X. Mao, S. Chiu, N. E. Alarcon, B. C.
Wilson, A. Vitkin, and V. X. D. Yang, "Feasibility of interstitial Doppler optical coherence
tomography for in vivo detection of microvascular changes during photodynamic therapy,"
Lasers in Surgery and Medicine, vol. 38, pp. 754-761, Sep 2006.
[11]
M. V. Sarunic, S. Weinberg, and J. A. Izatt, "Full-field swept-source phase microscopy," Optics
Letters, vol. 31, pp. 1462-1464, May 15 2006.
[12]
B. R. White, M. C. Pierce, N. Nassif, B. Cense, B. H. Park, G. J. Tearney, B. E. Bouma, T. C.
Chen, and J. F. de Boer, "In vivo dynamic human retinal blood flow imaging using ultra-highspeed spectral domain optical Doppler tomography," Optics Express, vol. 11, pp. 3490-3497,
DEC 15 2003.
[13]
M. A. Choma, A. K. Ellerbee, S. Yazdanfar, and J. A. Izatt, "Doppler flow imaging of
cytoplasmic streaming using spectral domain phase microscopy," Journal of Biomedical Optics,
vol. 11, pp. -, Mar-Apr 2006.
[14]
H. W. Ren, T. Sun, D. J. MacDonald, M. J. Cobb, and X. D. Li, "Real-time in vivo dblood-flow
imaging by movingscatterer-sensitive spectral-domain optical Doppler tomography," Optics
Letters, vol. 31, pp. 927-929, Apr 2006.
[15]
B. J. Vakoc, S. H. Yun, J. F. de Boer, G. J. Tearney, and B. E. Bouma, "Phase-resolved optical
frequency domain imaging," Optics Express, vol. 13, pp. 5483-5493, Jul 11 2005.
[16]
D. C. Adler, R. Huber, and J. G. Fujimoto, "Phase-sensitive optical coherence tomography at up
to 370,000 lines per second using buffered Fourier domain mode-locked lasers," Optics Letters,
vol. 32, pp. 626-628, Mar 2007.
CHAPTER 4
4.0
Photothermal Detection of Gold Nanoshells with OCT
4.1 Motivation
Optical coherence tomography (OCT) has been proven to be a powerful tool for assessing tissue
architectural morphology. OCT has enabled three-dimensional (3D) imaging of biological samples with
micrometer resolutions, and analysis of internal organs has been made possible through the use of
minimally invasive miniaturized probes. Although OCT is a very powerful imaging modality it has not
been able to leverage recent advances in molecularly-sensitive contrast agents that are revolutionizing
other areas of biomedicine by enabling selective interaction between synthetic materials and specific cell
types. The synthetic material, such as a pharmaceutical compound, fluorescent molecule, or nanoparticle,
is conjugated to a targeting agent such as an antibody or peptide. When the conjugated substance is
introduced into a cell culture or organism, it binds only to cells where the targeting agent has high affinity
for a specific protein expressed on the cell surface. In this way, a therapeutic or diagnostic agent can be
targeted to individual cells exhibiting a certain phenotype or pathology.
Molecular targeting technology has led to dramatic advances in drug delivery, fluorescence imaging,
and photothermal therapy. Gold nanoparticles, consisting of a silica core and a gold outer shell, are
especially attractive for imaging applications due to their customizable absorption and scattering
properties, biocompatibility, and ease of conjugation to antibodies and peptides. Gold nanoparticles can
be designed to have high absorption at wavelengths where tissue absorption is low, which has also led to
their use as photothermal therapy agents. A clear synergy exists between OCT and molecularly-sensitive
contrast agents, although it has been challenging to detect low concentrations of these contrast agents
using OCT in the past. If these limitations could be overcome, the structure and pathologic state of tissue
could be studied in 3D, in vivo, in real time, and with micron-scale spatial resolutions.
A technique for performing ultrahigh-speed 3D-OCT with the ability to detect pathologic tissue or
cells based on molecular markers would have a profound impact on biomedicine. Development of
molecularly specific contrast mechanisms for OCT would fundamentally transform the field by enabling
3D imaging of biological function in addition to architectural and cellular structure. Development of
molecular contrast agents for OCT would achieve cross-disciplinary impact by creating a bridge between
the OCT, molecular biology and nanotechnology fields. The impact of molecular biology and
nanotechnology research would be multiplied by enabling access to new imaging platforms for 3D
structural and cellular imaging. The ability to integrate molecular and 3D structural imaging would have
powerful applications for small animal studies, accelerating cancer research. As molecular contrast agents
become clinically available, this also promises to improve the sensitivity and specificity for early cancer
detection in clinical applications. This work will also serve as a launching point for research on other
pathologies associated with abnormal protein expressions, such as neurodegenerative and cardiovascular
diseases, that have typically not taken advantage of advances in OCT. Methods for more sensitive and
specific detection of disease can improve patient outcome and reduce healthcare costs, impacting a major
public issue in the United States and the international community.
In this section of the thesis work, a novel method for detecting gold nanoparticles using phasesensitive OCT was developed. Photothermal phase modulations were induced in a sample by localized
heating with an amplitude-modulated laser diode. These phase modulations were detected with a highspeed phase-sensitive OCT system similar to the one described in Chapter 3. Extremely high signal to
noise ratios (SNR) were achieved by detection of the photothermal signal only at the known modulation
frequency, similar in effect to lock-in detection, suggesting the possibility of molecularly-specific OCT in
the near future.
4.2 Sources of Contrast in OCT
OCT produces cross-sectional and 3D images of tissue microstructure by interferometrically
measuring the amplitude and echo time delay of backscattered light [1]. OCT imaging can derive contrast
from sources that are either endogenous or exogenous to the tissue being imaged. The most commonly
utilized source of endogenous contrast is spatial variations in the scattering properties of the tissue, which
produces contrast in conventional OCT images. In this case, only the amplitude of the interference signal
is analyzed to form the image. Another source of endogenous contrast is velocity or flow within the
sample. Typically referred to as Doppler OCT or optical Doppler tomography (ODT), these techniques
analyze phase changes in the interference signal over brief time periods to detect vascular blood flow [27]. Endogenous OCT contrast can also be derived from variations in the size of scattering particles in the
tissue [8-10] or wavelength-dependant absorption of different tissue components [11, 12] using
spectroscopic OCT [13-15]. Finally, non-centrosymmetric endogenous tissue components, such as
collagen, can be detected using nonlinear methods such as second harmonic OCT [16-21].
Exogenous contrast agents have not typically been used in OCT, but recently have been studied more
closely. As in other biomedical imaging modalities, OCT contrast agents promise to enable enhanced
visualization of selected features such as microvasculature, epithelial structures, and diseased or abnormal
tissue. Agents such as methylene blue, rhodamine, and indocyanine green can be detected by signatures in
their electron relaxation times using pump-probe OCT [22-24]. OCT contrast enhancement has also been
demonstrated using scattering microspheres [25], near-infrared (NIR) dyes [26], iron oxide microparticles
[27], and, more recently, nanoparticles [28-36].
Gold nanoparticles are particularly attractive contrast agents since they can be targeted to biochemical
markers associated with specific types of disease such as cancer [37, 38], which suggests the possibility of
highly sensitive and specific OCT detection of early neoplasia. Gold nanoshells consist of an inner silica
core surrounded by a thin gold shell. By changing the relative dimensions of the core and shell, the optical
resonance frequency of the particles can be tuned from ultraviolet to near infrared wavelengths [28]. This
allows customized tailoring of the optical scattering and absorption properties of the particles to suit the
needs of the specific application. Gold nanoshells are highly biocompatible, water-soluble, and
commercially available. Nanoshells can be designed with high absorption, targeted to cancer cells, and
used for photothermal therapy with minimal damage to surrounding tissue [39]. Other types of gold
nanoparticles such as nanorods [33] and nanocages [29] exhibit similar properties to nanoshells and can
also be used for exogenous OCT contrast enhancement.
Methods for detecting exogenous contrast agents using OCT can be divided into two general
categories: passive techniques and active techniques. Passive detection techniques rely on time-invariant
differences in the optical properties of the agent compared to the tissue to generate contrast. Differences
in the absorption of near infrared dyes compared to tissue have been passively detected using
spectroscopic OCT [26]. Similar spectroscopic methods have been applied to detect gold nanoparticles,
where the absorption of the nanoparticles caused a blue shift of the OCT signal [29]. Optical scattering
can also be used to detect gold nanoparticles using conventional amplitude-based OCT [28, 30, 32-34, 36]
since the peak scattering or absorption wavelength of the nanoparticles can be selected to overlap with the
OCT imaging wavelength. Passive contrast agent detection may be difficult to apply in vivo, however,
since the signal is not background-free and variations in the optical properties of heterogeneous tissue can
mask the scattering and absorption characteristics of the agents.
Active contrast agent detection techniques modulate a property of the agent to enhance visualization
against a heterogeneous tissue background. One example of active contrast agent detection is
magnetomotive OCT [27, 31, 35]. In this technique, superparamagnetic iron oxide (SPIO) nanoparticles
are taken up by cells in the sample tissue and are then exposed to an external magnetic field of 0.06 - 0.5
T that is modulated at 3 - 50 Hz. Modulation of the external magnetic field causes localized motion in
regions of the tissue that have taken up the SPIO, and this motion is detected by fluctuations in the
amplitude [31] or phase [35] of the OCT interference signal. Magnetomotive OCT achieves a high signalto-noise ratio (SNR) for detecting SPIO nanoparticles since active modulation of the contrast agent results
in a detection scheme that is less susceptible to background noise. However, this technique requires the
application of fairly strong magnetic field gradients (up to 11 T/m) and is limited in imaging speed due to
the relatively slow mechanical response of SPIO-laden tissue. These factors may make magnetomotive
OCT challenging to apply for in vivo imaging in humans.
4.3 Imaging Gold Nanoparticles with Photothermal Phase Modulation
As part of this thesis work, an active contrast agent detection technique for high-speed OCT imaging
based on photothermal modulation was developed and demonstrated. The technique used gold nanoshells
designed to have high absorption at 808 nm where tissue absorption is inherently low. A multimode laser
diode operating at 808 nm was used to induce small-scale, localized temperature gradients in regions of
the sample that contain the contrast agent. These temperature variations altered the optical path length in
the sample. Changes in path length were detected using a swept source OCT phase microscopy system
[40-43] built using a double-buffered Fourier domain mode locked (FDML) laser operating at 1315 nm
and a sweep rate of 240,000 sweeps per second (240 kHz). By modulating the 808 nm laser diode at a
known frequency and observing variations in optical path length that occur only at that frequency, the
contrast agent can be detected in a way that significantly reduces background noise. Contrast agent SNR's
of up to 131 are obtained using modulation frequencies of 500 Hz - 60 kHz. The technique described here
can be integrated with 3D-OCT imaging to provide contrast-enhanced images of tissue architectural
morphology. In the future, photothermal detection of gold nanoshells using high-speed, phase-sensitive
OCT may enable targeted in vivo imaging of disease with high sensitivity and specificity.
4.3.1 ExperimentalSetup
A phase-sensitive swept source OCT system, double-buffered FDML laser, and amplitude modulated
808 nm laser diode were used to detect gold nanoshells in this experiment. The setup is shown in Figure
4.1. The swept source OCT phase microscope is similar to the design previously described in Chapter 3
but was modified to collinearly direct the 808 nm laser diode beam and the 1315 nm OCT beam onto the
sample. 95% of the FDML output was routed to a common path interferometer, designated as the "sample
interferometer." The liquid sample was held in a glass cuvette where the first glass/liquid interface
provides the reference reflection for the interferometer. The output of the fiber-pigtailed 808 nm laser
diode was combined with the OCT beam using a dichroic mirror. The diode had a maximum continuous
wave output power of 300 mW and was pigtailed to a multimode fiber with a 50 [tm core diameter. The
808 nm beam was collimated by a 15 mm focal length lens and the 1315 nm OCT beam was collimated
by a 20 mm focal length lens.
An XY pair of galvanometer mirrors with a 6 mm clear aperture was used to aim the combined 808
nm / 1315 nm beam on the sample. The beams were focused using a 30 mm focal length achromatic
2
objective lens. The 808 nm beam diameter was -140 Vm at the 1/e intensity point, while the 1300 nm
beam diameter was 15 tm at the 1/e2 intensity point. The 808 nm spot size was measured with a CCD
camera while the 1300 nm spot size was estimated using a resolution test target. A smaller diameter is
generally more desirable for the 808 nm beam in order to increase the energy density and induce larger
optical path changes in the sample. However, the larger beam diameter ensured uniform heating in the
volume interrogated with the OCT system and also simplified alignment of the 808 nm beam to the OCT
beam.
The remaining 5% of the FDML laser output was routed to a second common path interferometer,
designated as the "calibration interferometer," which used a 210 [m thick glass slide as the sample. The
front and back air/glass boundaries generate two fields that interfere to produce a calibration signal for
resampling the sample fringes onto a linear k spacing, and for removing slow phase drift caused by the
FDML laser [43]. The sample and calibration data were acquired simultaneously using a 2 GS/s, 8 bit
digital oscilloscope. Data processing was performed post-acquisition using a personal computer as
described in Section 4.3.3. The 808 nm laser diode was modulated by a digital pulse generator that was
synchronized to the beginning of each wavelength sweep.
The phase noise of the system was measured by placing a 210 [tm glass slide in the sample
interferometer and recording the position of the back surface relative to the front surface over 30 ms. The
slow component of the phase drift caused by the FDML laser was recorded using the calibration
interferometer and subtracted from the sample data as described in Chapter 3. The results of this
measurement are shown in the inset of Figure 4.1. The phase noise was measured to be 2.2 mrad,
corresponding to a displacement sensitivity of 153 pm.
4.3.2 Sample Preparation
Gold nanoshells with a 120 nm diameter core, 16 nm shell thickness, and peak absorption wavelength
of 780 nm were obtained commercially. The nanoshells were mixed with deionized water and diluted to a
concentration of 1010 mL' (16.6 pM). At this concentration, the nanoshell absorption coefficient at 808
nm was approximately 3.88 cm-' with a FWHM bandwidth of -400 nm. A glass cuvette with a sample
path of 200 jim was filled with the solution, and the cuvette was placed in the sample interferometer of
the OCT phase microscope. As shown in Figure 4.2, the glass/fluid interface between the cuvette cover
and the nanoshell solution was used as the reference reflection for the common path interferometer. The
fluid/glass interface between the nanoshell solution and the cuvette body was monitored for small changes
in optical path length, corresponding to localized absorption and heating of the nanoshells during
exposure to the 808 nm laser beam. For control experiments, pure deionized water was placed in the same
cuvette instead of the gold nanoshell solution.
4.3.3 Data Processing
Data processing was very similar to the method described for making phase-sensitive measurements in
Chapter 3. The interference fringes from the sample interferometer was recorded as a function of time in a
single long record using a digital oscilloscope as described above. The interference fringes generated by
the calibration interferometer were simultaneously recorded on a separate oscilloscope channel. Data was
segmented into individual sweeps and the measured phase as a function of time was extracted using the
method described in Chapter 3. The slow long-term phase drift from the FDML laser extracted from the
calibration data was low-pass filtered and subtracted from the sample phase. Once the phase response of
the liquid phantom sample was obtained, a Fourier transform was applied to measure the induced phase
modulations as a function of frequency.
4.4 Thermal Modeling
Two separate thermal models were constructed for this portion of the thesis work. The first model was
used to estimate the temperature rise in the sample based on the observed phase variations. This model
was required since it was impossible to directly measure temperature oscillations in the sample on the
order of - 1' C over a sample volume of 3 x 10-3 mm 3 and millisecond time scales. This model was also
necessary in order to make maximum temperature rise estimates for biosafety calculations for future in
vivo applications. The second model was used estimate the expected thermal response of the sample based
on the thermo-optic properties of the sample material and 808 nm laser. This model was used to verify the
results of the first model.
4.4.1 Phase to Temperature Conversion
The temperature change in the sample is estimated from the measured phase variations by modeling
the effects of temperature on optical path length. Although the mechanical and thermal dynamics of the
phantom used in these experiments are fairly complex, some understanding of the system behavior can be
achieved by modeling two major effects that work to change the optical path length in opposing directions
as the sample temperature is increased. First, the refractive index of water decreases with increasing
temperature, which tends to decrease the optical path length. Second, the volumetric thermal expansion
coefficient of water is positive near room temperature, which tends to increase the optical path length. For
the phantom apparatus used in these experiments, the measured optical path length of the sample, z(T),
varies with temperature T and can be modeled as:
z(T)= L(T)n(T)
(4.1)
Here, L(T) is the physical path length and n(T) is the refractive index. A change in optical path
length Az that occurs due to a change in temperature AT relative to an initial condition T can be
written as:
Az = z(To +AT)-z(To)= z(To +AT)-zo
(4.2)
Az = L(To +AT)n(T o +AT)-zo
(4.3)
Here, zo is the initial optical path length at To . With a volumetric coefficient of expansion p, an
initial physical path length Lo, an initial refractive index no , and a variation in refractive index with
temperature dn / dT , the change in optical path length can be expressed using:
(4.4)
L(To + AT)= L o x(1+ jAT)
dn
dT
(4.5)
n(T o +AT)= no +-AT
Az = Lo(1+
AT) no + -AT
dT
(4.6)
-zo
This formulation assumes that the fluid column illuminated by the 808 nm laser is free to expand only
in the axial direction. Axial expansion could occur since the cuvette cover was not tightly fixed to the
body. Note that the absolute change in optical path length associated with one thermal modulation is <
120 nm for a 500 Hz modulation frequency and < 7 nm for a 60 kHz modulation frequency. These small
size scales complicate the dynamic response of the expanding and contracting water column, so the model
used here may not precisely reflect the actual behaviour of the system. In this model, dn / dT is assumed
to be constant with temperature. The expression for Az can be expanded to give:
Cn
dn_
Az = Lono + Lo dn AT + LoPATno + LoAT 2 dn
dT
dT
dn
dn
AT 2
Az = Ldn AT + LonoAT + Lof
dT
dT
The swept source OCT phase microscope measures phase changes Aq
(4.7)
(4.8)
in the OCT interference
fringes. A0 is related to Az through the expression:
(4.9)
Az = -o A
4zC
Therefore the estimated sample temperature variation can be calculated by solving the following
quadratic expression for AT:
+Lon
Lodn AT2 + Lo
dT
dT
o
AT
4z
=0
(4.10)
Although it is possible to explicitly solve Equation (10) for AT, in reality P is a function of
temperature as well, P=P(T). For water, the variation in P is significant and changes by -50%
between 20 and 30C. Therefore Equation (4.10) was solved numerically using a mathematics package.
Figure 4.3 and Figure 4.4 show the estimated sample temperature T in the illuminated volume versus
0
the observed OCT signal phase change A0 for a room temperature of 20 C and two different measured
-i,
To = 20 OC, Lo = 200 lpm, and A0 = 1315
phase ranges. For this model, values of dn/dT = -91 x 10 6 OC
0
-1
nm were used. p values ranged from 207 x 10-6 OC-1 at 20 0 C to 385 x 10 6 oC at 40 C. Since P is
several times larger than dn / dT , sample expansion dominates the system and the net optical path length
changes is positive. The quadratic term in Equation (4.10) does not contribute substantially to the solution
but was taken into account nonetheless.
4.4.2 EstimatedThermal Responses
A second model was applied to understand photothermal contrast behavior as the 808 nm power and
spot size are scaled. This model enables calculation of a "forward estimate" of thermal response from the
thermo-optic properties of the sample and laser. The system was modeled as an absorbing Gaussian
cylinder surrounded by an infinite and homogenous medium. This simplified model does not take into
account the effects of scattering on the 808 nm beam shape, but provides a reasonable estimate of the
relationship between induced temperature change and 808 nm beam parameters. The model also assumes
that the spot size is much smaller than the penetration depth due to absorption, which is true for the
parameter space considered below.
The heat conduction equation in a cylindrical geometry is given by:
aAT(t,z,r)
at
=
pa(Z, r)
+a
pc
a2AT(t, z, r)
az2
2
+
AT(t,z,r)
ar 2
+
a 2AT(t, z,r)
rr
9
(4.11)
(4.11)
Here, t is time, z is the axial distance from the top of the cylinder, r is radial distance from the center of
the cylinder, pa is the absorption coefficient, p is the fluence rate of the laser, p is the density of the
medium, c is the specific heat of the medium, and a is the thermal diffusivity of the medium. For small
spot sizes relative to the absorption depth, the heat conduction equation is dominated by radial heat
transfer [44]. Combined with a cylindrical geometry, this allows the heat conduction equation to be
solved in closed form. For one 808 nm modulation period, the temperature variation can therefore be
modeled as [44]:
AT(t,r=0)=E
pc
8a )
In 1
(
, W < 1/Pat
<
t
(4.12)
(t-t,,rE=
-AT
=
VE
In 1+ W
28 +
(a- t t
p)
W<
Pla,t 2 tp
(4.13)
Here, E is the irradiance, W is the 1/e2 beam radius, and tp is the exposure duration. E and a are
calculated from fundamental material properties of the sample, given by:
E = 2PLs / trW2
(4.14)
a = k /pc
(4.15)
Here, Pps is the pulse power of the 808 nm laser and k is the thermal conductivity of the surrounding
medium.
Equations (4.12) and (4.13) were used to model the expected temperature profile for an 808 nm
modulation frequency of 500 Hz over one pulse period. The following parameter values were used for the
model: p = 1000 kg/m3, c = 4186 J/kg K, t, = 1 ms, Pa = 388 m-', and k = 0.6 W/m K. The value for
was obtained from measurements performed by the nanoshell manufacturer. Thermal responses were
modeled for 808 nm beam radii of 5 - 40 jtm. The pulse power was chosen to give a constant temperature
increase of 1 'C for each beam radius, corresponding to average powers of 2.6 - 18.2 mW at radii of 5 40 [tm. The results of the model are shown in Figure 4.5, with the legend indicating the beam radius and
pulse power used to generate each curve. During the second half of each modulation cycle (t = 1 ms to t =
2 ms), the simulation indicates that temperatures decrease to 18% - 62% of their peak values. This
suggests that a slow temperature drift may be difficult to avoid. Decreasing the beam radius, however,
enables a given thermal increase to be achieved with less incident power, and also results in more rapid
cooling. This would result in a larger phase modulation and lower tissue exposure, emphasizing the need
to reduce the 808 nm beam diameter for in vivo applications.
4.5 Experimental Results
OCT phase microscopy was performed on two samples: one containing pure deionized water and one
containing a diluted gold nanoshell solution. In both cases measurements were taken both with and
without exposure to the 808 nm laser source. The results of these experiments are shown in Figure 4.6. In
Figure 4.6 all plots show the interference fringe phase associated with the second fluid/glass interface,
measured over time, at a single spatial location in the cuvette. Figure 4.6(A) shows the measured phase
when the sample is pure deionized water and the 808 nm laser is disabled. The phase profile is generally
featureless and corresponds to noise. Figure 4.6(B) shows the measured phase when the sample is a 1010
mL' gold nanoshell solution and the 808 nm laser is disabled. Phase noise is increased due to increased
scattering in the sample, but no systematic pattern is observed. This nanoshell concentration is consistent
with estimates of concentrations that may be attainable in tumor tissue following systemic administration
of antibody-labeled nanoshells [32].
Figure 4.6(C) shows the measured phase from the deionized water sample when the 808 nm laser is
activated. The 808 nm laser was set to provide 276 mW and was modulated with a 500 Hz square wave
with a 50% duty cycle, giving an average power of 138 mW. The 1310 nm FDML laser provided an
additional 20 mW of power on the sample. The 808 nm laser modulation pattern is shown at the top of the
plot, and the vertical line indicates the time at which the 808 nm laser was switched on at t = 4 ms. No
change in the phase is observed compared to Figure 4.6(A), indicating that the absorption of water at 808
nm is not high enough to cause localized heating and induce optical path changes.
Figure 4.6(D) shows the measured phase from the nanoshell solution with the 808 nm laser activated.
The modulation parameters were identical to those used for the deionized water sample, and the vertical
line indicates the time at which the laser was switched on at t = 4 ms. In this case, a strong phase response
is observed. The high absorption of the gold nanoshells at 808 nm causes localized heating of the
solution, which in turn increases the optical path length of the sample. The phase response of the sample
shows the same modulation pattern as the 808 nm laser. Each phase modulation Aq is -1.1 rad peak-topeak, corresponding to a physical path difference AL of -87 nm using AL = ,0 -A0 / 47rn where
0=
1315 nm is the center wavelength of the FDML laser and n = 1.33 is the refractive index of water. Since
there is insufficient time for the solution to fully cool using these 808 nm modulation parameters, there is
a slow increase in temperature producing a cumulative increase in optical path of -7.5 rad or 590 nm over
30 ms.
Although it clear from Figure 4.6 that gold nanoshells can be detected by direct inspection of the
measured phase under certain conditions, an enhanced SNR can be achieved by detecting modulations in
the phase signal. One approach is to Fourier transform the measured phase and search for a peak at the
precisely-known 808 nm modulation frequency. This concept is illustrated in Figure 4.7. The plots in
Figure 4.7 show the frequency spectrum of the measured phase for each test condition in Figure 4.6. In
each case, the phase data from t = 0 to t = 4 ms was removed. The slow phase increase from gradual
heating in Figure 4.6(D) was removed prior to Fourier transformation by subtracting a quadratic fit from
the measurement. When the 808 nm laser is disabled (Figure 4.7(A,B)), whether the sample contained
deionized water (Figure 4.7(A)) or a nanoshells solution (Figure 4.7(B)), the only characteristic feature of
the frequency spectra is I/f noise. The same is true when the 808 nm laser is modulated at 500 Hz but the
sample contains deionized water (Figure 4.7(C)). However, when the 808 nm laser is modulated at 500
Hz and the sample contains nanoshells, a strong peak is seen in the frequency spectrum of the measured
phase at exactly 500 Hz (Figure 4.7(D)). Smaller harmonic peaks are also visible at 1 kHz frequency
increments, consistent with the frequency spectrum of a triangular waveform repeating at 500 Hz.
The "signal" in these measurements is defined as the peak Fourier transform amplitude within ± 20 Hz
of the nominal 808 nm modulation frequency. This range was selected to allow for a small absolute error
in the modulation frequency. The "noise" is defined as the peak FT amplitude within the same ± 20 Hz
window measured using the first 4 ms of data. During this time the 808 nm laser was disabled in all cases,
allowing for an accurate and consistent estimate of phase noise at the 808 nm modulation frequency. The
SNR measurements were repeated five times for each combination of sample type (with and without
nanoshells) and 808 nm laser state (disabled or enabled for t > 4 ms). The SNR is defined as the ratio of
the signal peak to the noise value. In Figure 4.7 SNR values are shown as the mean of five repeated
measurements, plus or minus one standard deviation. As shown in Figure 4.7(A,C), SNR values are
insignificant when nanoshells are not present. This indicates the lack of a photothermal modulation signal.
SNR values are also insignificant when nanoshells are present but the 808 nm laser is switched off
(Figure 4.7(B)). However, when nanoshells are present in the sample and the 808 nm laser is modulated at
500 Hz, the SNR is 131 ± 91.
4.5.1 Signal to Noise Ratio versus Modulation Frequency
As the modulation frequency of the 808 nm laser is increased, the frequency of the phase modulation
in the nanoshell solution also increases. This has the benefit of shifting the photothermal modulation peak
to higher frequencies where 1/f noise is reduced. However, the sample volume illuminated with the 808
nm beam has less time to heat and cool, resulting in a smaller optical path modulation and lower Fourier
transform peak amplitude. To investigate this tradeoff three sets of additional experiments were
conducted using 808 nm modulation frequencies of 1, 15, and 60 kHz. Each experiment was repeated five
times under the same conditions to test for consistency. The results are shown in Figure 4.8. Figure 4.8(AC) shows the measured phase from the nanoshell solution at one transverse position on the cuvette. The
vertical line in each plot indicates the time when the 808 nm beam was switched on at t = 4 ms. The insets
show enlarged views of 500 ps (Figure 4.8 (B)) and 90 [ts (Figure 4.8(C)) segments of the measured
phase with phase modulations clearly visible. For all three cases a gradual path change is visible due to
slow heating of the sample, although this effect is lower at 15 and 60 kHz.
Figure 4.8(D-F) shows the Fourier transform of the measured phases beginning at 4 ms and with the
slow phase component removed by subtracting a quadratic fit. The SNR is measured at each modulation
frequency as described in Section 4.5 above. An optimum SNR of 112 ± 45 is achieved at a modulation
frequency of 15 kHz. At a modulation frequency of 1 kHz, the SNR suffers from low frequency 1/f noise
near the baseband. At a modulation frequency of 60 kHz, the benefits of decreased 1/f noise are
outweighed by the decrease in peak modulation amplitude. It is expected that the optimal modulation
frequency will vary depending on the optical properties of the sample, nanoshell concentration, and 808
nm laser power level.
Although other techniques used to detect gold nanoparticles with OCT have used different phantom
systems, which makes direct comparison difficult, reported SNR's have typically ranged from 1.5 - 5 [29,
32-34, 36]. There has been one previous report of SNR values of 79 - 631 using highly backscattering
nanoshells in a non-scattering water sample, but the SNR decreased to 5 when a scattering tissue phantom
that more closely approximated biological tissue was used [32]. The phantom used in the photothermal
experiments described here does not include scattering but accurately reflects the absorption properties of
biological tissue [45] and achieves mean SNR's of 2 - 131. The photothermal modulation method is
expected to perform well in scattering tissue since the technique is less sensitive to background noise
because phase changes are detected at a specific modulation frequency away from baseband. In addition,
this technique uses absorption rather than scattering to generate contrast. The high SNR also suggests that
lower concentrations of nanoparticles could be detectable using this method than other methods, which is
relevant for in vivo applications where the agent is administered systemically and accumulates at lower
levels in diseased tissue [34].
4.5.2 Effect of Measurement Time on Signal to Noise Ratio
For in vivo imaging applications, it is important to determine how long one region of the sample must
be measured in order to obtain a reasonable contrast agent SNR. Longer measurement times lead to
increased SNR but decreased overall frame rates. This tradeoff was evaluated by shortening the time
window used in the Fourier transform for an 808 nm modulation frequency of 15 kHz. One representative
dataset was used for this test, and the results are shown in Figure 4.9. A linear relationship between the
observation time and the SNR is observed in the measured data. As the observation time window is
decreased from 28 ms to 2 ms, the SNR decreases linearly from 108 to 7. These results suggest that
observation times of only a few ms may be required per transverse position in order to obtain reasonable
nanoshell contrast using Fourier transform analysis. Other data analysis techniques may be developed in
the future that can obtain similar contrast in shorter time periods.
4.5.3 Comparison to Model Results
The phase measurements described in Section 4.5 were compared to the models developed in Section
4.4 in order to estimate the temperature response associated with the measured phase responses in the
sample. The estimated temperature response was then compared to the "forward" thermal model based on
the thermo-optic properties of the sample and 808 nm laser. Figure 4.10 shows a comparison between the
forward model results (blue "expected" curve) and the thermal response estimated from the actual phase
measurements shown in Figure 4.6(d) for t = 4 ms to 6 ms. The measured phase was converted to an
estimated temperature increase using Equation (4.10). The expected model parameters were adjusted to
reflect the actual experimental conditions. A beam radius of 70 tm was used with an average incident
power of 138 mW. The expected and estimated modles show good correlation, indicating that the thermal
models described here are reasonable.
Experimental results showed average phase modulations of ±575, ±369, ±93, and ±32 mrad at 808 nm
laser modulation frequencies of 0.5, 1, 15, and 60 kHz, respectively, as shown in Figure 4.6(D) and
Figure 4.8. By comparing the measured phase modulations to the model results shown in Figure 4.3, the
estimated temperature fluctuations are ±1.47, ±0.98, ±0.26, and ±0.090 C at 808 nm laser modulation
frequencies of 0.5, 1, 15, and 60 kHz, respectively. If the 808 nm laser is held at one transverse position
for 26 ms, slow phase increases of -4 - 8 rad are observed, corresponding to temperature increases of
-7.4 - 12.3 0 C. This slow temperature increase can be minimized in in vivo imaging applications by
translating the beam faster, since only a few ms of observation time per transverse location may be
required. Additionally, the use of a more tightly focused 808 nm beam would permit more rapid cooling
as shown in Figure 4.5. This result also highlights the potential of this technique for conducting
photothermal therapy as well as OCT imaging, since even larger temperature gradients could be induced
by increasing the 808 nm exposure levels.
4.6 Limitations
Although good performance was achieved in this phantom experiment, additional studies are required
to demonstrate the utility of the technique in vivo. The phantom used here was selected to test the concept
of photothermal detection with the most critical optical property of the sample, absorption, isolated from
other effects. Although there are challenges for applying this method in vivo, OCT phase microscopy has
previously been applied for studies of living cells and tissue by placing a thin coverslip in contact with the
sample [40, 46-48]. Similar approaches may be used for the technique described here, with the top surface
of a thin glass window or endoscope sheath in light contact with the tissue providing the reference field.
Other groups have demonstrated that optically-induced thermal gradients can result in physical tissue
displacements ex vivo that are detectable with phase-sensitive OCT [49]. Although these previous studies
induced large temperature gradients and measured correspondingly large phase changes, they indicate that
the photothermal modulation contrast mechanism described here should remain valid in tissue using
smaller temperature gradients and measuring smaller phase modulations.
More fundamentally, recent reports have indicated that it is difficult to accurately measure small phase
changes < 0.1 rad in vivo using OCT [50]. This phase uncertainty is due to speckle noise and sample
motion, and can be reduced through spatial averaging and other signal processing techniques. With
photothermal detection of gold nanoshells, phase variations of + 0.098 rad can be induced at a modulation
frequency of 15 kHz and much larger variations are possible at lower modulation frequencies. This
indicates that observation of the photothermal modulation signal should be possible in vivo. The necessity
to spatially average the phase measurements over the axial or transverse dimensions in order to detect the
photothermal modulation could decrease the effective spatial resolution or imaging speed of this
technique. It would still be possible, however, to obtain conventional 3D-OCT images at high resolution
concurrently with the contrast-enhanced images by applying standard OCT signal processing.
Another challenge for in vivo applications is ensuring that the thermal changes induced by the 808 nm
laser are small enough to preclude inadvertent tissue damage. The thermal changes have two components:
a small-amplitude component associated with the 808 nm modulation over tens of microseconds, and a
slow rise in temperature associated with gradual heating over tens of milliseconds. For this discussion, it
is assumed that in vivo temperature increases will be similar to those obtained in the phantom
experiments. The results shown here indicate that for optimal SNR conditions, the temperature rise
associated with the 808 nm modulation is -0.52 OC. This is an extremely small thermal change that is not
expected to cause tissue damage. The gradual temperature increase associated with slow sample heating
at optimal SNR conditions is -8 'C over 26 ms. For in vivo applications, however, the observation time at
one transverse spot could be reduced to 5 ms while maintaining an SNR of -20 which would cause an
overall temperature increase of only -3.5 0 C. Since the tissue would only be exposed to this elevated
temperature for several milliseconds as the beam is scanned in the transverse dimension, no significant
tissue damage is expected. By comparison, photothermal and photodynamic therapies typically require
temperature increases of several tens of degrees for many tens of seconds to induce permanent tissue
damage. Observation times may be further decreased by developing new analysis algorithms that more
efficiently detect the photothermal phase modulation. The use of a single mode laser diode instead of the
multimode 808 nm laser diode would allow thermal gradients to be induced over a more localized area.
This would reduce the power required to create a given phase modulation and would simultaneously
increase cooling rates, further increasing safety margins.
One final consideration for in vivo applications is the effect of scattering in biological tissue.
Scattering may decrease the effectiveness of the photothermal modulation technique for large tissue
depths, since 808 nm light penetration will be reduced. In the phantom experiments performed here, the
phase measurement was performed at a fluid/glass interface at a single transverse position. In an in vivo
application, the OCT and 808 nm beams would be continuously scanned across the sample in a transverse
direction. Due to the spatial distribution of scattering centers in biological tissue, phase measurements
would be obtained at each axial position in the sample in a manner similar to Doppler OCT techniques.
The photothermal phase modulation may be detected by comparing the measured phase at each axial
point to the measured phase in subsequent axial lines, or by comparison to a fixed phase reference such as
a glass window or endoscope sheath. The combination of transverse beam scanning and axially
distributed phase measurements will add speckle noise to the photothermal signal. Since nanoshell
contrast is derived from a phase modulation at a precisely known frequency, however, the effects of
random background noise such as speckle and sample motion are expected to be minimized. Increasing
the 808 nm modulation frequency may also have important benefits in scattering systems since this will
allow more phase modulation periods to be captured over each speckle cell. Higher modulation
frequencies could therefore increase OCT frame rates and take maximum advantage of the speed
advantage offered by FDML lasers.
Although the initial results shown here suggest that the photothermal modulation technique may
provide performance benefits in vivo, more study is required to verify this. Future experiments will focus
on evaluating SNR performance in solid tissue phantoms that more closely approximate the scattering,
thermal conductivity, and mechanical properties of biological tissue. Minimum detectable nanoshell
concentrations will be studied in these phantoms and assessed for benefits compared to scattering- or
absorption-based contrast. Finally, experiments in in vitro and in vivo biological tissue are needed to
validate the photothermal contrast technique.
4.7 Figures
System Phase Noise
0.01
0.005
0
-0.005
0.010
10
20
30
Time [ms]
Figure 4.1. Swept-source OCT phase microscope with photothermal modulation system. C1, C2, C3,
collimating lenses. OBJ, objective lens. DCM, dichroic mirror. X,Y, galvanometer mirrors. PD, photodiode.
A, amplifier. TRG, sweep trigger. CH 1, OCT signal input. CH 2, calibration signal input. DAQ, data
acquisition. Inset shows measured phase noise of 2.2 mrad. Originally published in [51].
15 pm
II
Cuvette
Cover
Reference
Surface
II
Estimated Temperature vs. Phase Change
Nanoshell
Solution
Cuvette
Body
\
I
Sample
Surface
I
I
I
140 pm
Figure 4.2. Sample holder and beam geometries for
photothermal detection of gold nanoparticles. Beam
widths are approximate l/e' points of optical intensity.
Originally published in [51].
_0
2
4
6
8
10
Observed Phase Change [rad]
Figure 4.3. Thermal modeling results
showing estimated sample temperature
calculated from observed phase changes
from 0 - 10 rad. Originally published in [51].
Modeled Thermal Response, 5-40 Lm Beam Radii
Estimated Temperature vs. Phase Change
,O
1.5
1
0.5
Observed Phase Change [rad]
'0
2
1
1.5
2
Time [ms]
Figure 4.5. Thermal modeling results showing
expected temperature increases for 808 nm
beam radii of 4 - 40 pm. Originally published in
[51].
Figure 4.4. Thermal modeling results showing
estimated sample temperature calculated from
observed phase changes from 0 - 2 rad.
Originally published in [51].
Nanoshells (+)
Nanoshells (-)
nAu
0.5
a
0.3
-0.
3
51
-0.
10
20
Time [ms]
2
Time [ms]
0.5
C
0.3
0.1
-0.1
-0.3
II
Time [ms]
15
Time [ms]
Figure 4.6. Measured phase from the back surface of the cuvette for various experimental configurations.
1
laser
A: Deionized water with 808 nm laser deactivated. B: 1 x 1010 mL- nanoshell solution with 808 nm
laser
the
shows
train
pulse
Red
Hz.
500
at
deactivated. C: Deionized water with 808 nm laser modulated
Red
Hz.
500
modulation signal. D: 1 x 1010 mL-1 nanoshell solution with 808 nm laser modulated at
contains
sample
pulse train shows laser modulation signal. Phase modulations are visible only when
nanoshells and when the 808 nm laser is activated. Originally published in [51].
Nanoshells (+)
Nanoshells (-)
Ln
4)
I.
0
0
(U
E
C
Co
0
2000
3000
4000
5000
2000
1000
-0
+
5000
3000
Frequency [Hz]
Frequency [Hz]
2
[dl
1500)
1
E
SNR = 131 ±91
a
E 1000
00
o
0
5001
"'
II
3000
2000
Frequency [Hz]
5000
"0
l~n~_A
1000
2000 3000
Frequency [Hz]
1L-'
4000
A
5000
Figure 4.7. Fourier transform of phase vs. time curves shown in Figure 4.6. A: Deionized water with 808
1
nm laser deactivated. B: 1 x 1010 mL- nanoshell solution with 808 nm laser deactivated. C: Deionized
1
water with 808 nm laser modulated at 500 Hz. d, 1 x 1010 mL- nanoshell solution with 808 nm laser
modulated at 500 Hz. Strong peak is observed at 500 Hz when nanoshells are present and the 808 nm
laser is activated. Originally published in [51].
15 kHz Modulation
1 kHz Modulation
60 kHz Modulation
1b
8-
I
4-
0.
10
0
10
20
Time [ms]
Time [ms]
1000.
SNR =70t 24
d
e
3(
Time [ms]
SNR=112 a 45
800
600
400
200
-0
2
4
6
8
10
10
12
Frequency [kHz]
14
16
18
Frequency [kHz]
20
35
57
59
61
63
Frequency [kHz]
Figure 4.8. Measured phase (A-C) and Fourier transforms (D-F) of measured phase at various 808 nm
laser modulation frequencies. Red lines in (A-C) show time when 808 nm laser was activated. Insets in
(B,C) show enlarged views of the measured phase. 808 nm laser modulation frequencies were 1 kHz
(A,D), 15 kHz (B,E), and 60 kHz (C,F). Originally published in [51].
Estimated and Expected Temperature Increases
Signal-to-Noise Ratio vs. Observation Time
0
5
10
15
20
Observation Time [ms]
25
Figure 4.9. Measured gold nanoshell signal-tonoise ratio (SNR) as the sample observation time
is increased for an 808 nm laser modulation
frequency of 15 kHz. Originally published in [51].
30
3
Time [ms]
Figure 4.10. Comparison of estimated thermal
response from phase measurements and
expected thermal response from thermo-optic
sample properties. Models used a modulation
frequency of 500 Hz, beam radius of 70 pm, and
average power of 138 mW.
4.8 References
[1]
D. Huang, E. A. Swanson, C. P. Lin, J. S. Schuman, W. G. Stinson, W. Chang, M. R. Hee, T.
Flotte, K. Gregory, C. A. Puliafito, and J. G. Fujimoto, "Optical Coherence Tomography,"
Science, vol. 254, pp. 1178-1181, Nov 22 1991.
[2]
S. Yazdanfar, M. D. Kulkarni, and J. A. Izatt, "High resolution imaging of in vivo cardiac
dynamics using color Doppler optical coherence tomography," Optics Express, vol. 1, 1997/12/22
1997.
[3]
Z. Chen, T. E. Milner, D. Dave, and J. S. Nelson, "Optical Doppler tomographic imaging of fluid
flow velocity in highly scattering media," Optics Letters, vol. 22, pp. 64-6, 1997/01/01 1997.
[4]
Z. Ding, Y. Zhao, H. Ren, J. S. Nelson, and Z. Chen, "Real-time phase-resolved optical
coherence tomography and optical Doppler tomography," Optics Express, vol. 10, 2002/03/11
2002.
[5]
V. X. D. Yang, M. L. Gordon, B. Qi, J. Pekar, S. Lo, E. Seng-Yue, A. Mok, B. C. Wilson, and I.
A. Vitkin, "High speed, wide velocity dynamic range Doppler optical coherence tomography
(Part I): System design, signal processing, and performance," Optics Express, vol. 11, pp. 794809, Apr 7 2003.
[6]
V. X. D. Yang, M. L. Gordon, E. Seng-Yue, S. Lo, B. Qi, J. Pekar, A. Mok, B. C. Wilson, and I.
A. Vitkin, "High speed, wide velocity dynamic range Doppler optical coherence tomography
(Part II): imaging in vivo cardiac dynamics of Xenopus laevis," Optics Express, vol. 11,
2003/07/14 2003.
[7]
V. X. D. Yang, M. L. Gordon, S. J. Tang, N. E. Marcon, G. Gardiner, B. Qi, S. Bisland, E. SengYue, S. Lo, J. Pekar, B. C. Wilson, and I. A. Vitkin, "High speed, wide velocity dynamic range
Doppler optical coherence tomography (Part III): in vivo endoscopic imaging of blood flow in the
rat and human gastrointestinal tracts," Optics Express, vol. 11, pp. 2416-2424, Sep 22 2003.
[8]
D. C. Adler, T. H. Ko, P. R. Herz, and J. G. Fujimoto, "Optical coherence tomography contrast
enhancement using spectroscopic analysis with spectral autocorrelation," Optics Express, vol. 12,
pp. 5487-5501, Nov 1 2004.
[9]
C. Y. Xu, P. S. Carney, and S. A. Boppart, "Wavelength-dependent scattering in spectroscopic
optical coherence tomography," Optics Express, vol. 13, pp. 5450-5462, Jul 11 2005.
[10]
S. D. Dyer, T. Dennis, L. K. Street, S. M. Etzel, T. A. Germer, and A. Dienstfrey, "Spectroscopic
phase-dispersion optical coherence tomography measurements of scattering phantoms," Optics
Express, vol. 14, pp. 8138-8153, Sep 2006.
[11]
D. J. Faber, E. G. Mik, M. C. Aalders, and T. G. van Leeuwen, "Light absorption of (oxy)hemoglobin assessed by spectroscopic optical coherence tomography," Opt Lett, vol. 28, pp.
1436-8, Aug 15 2003.
[12]
D. J. Faber, E. G. Mik, M. C. G. Aalders, and T. G. van Leeuwen, "Toward assessment of blood
oxygen saturation by spectroscopic optical coherence tomography," Optics Letters, vol. 30, pp.
1015-17, 2005/05/01 2005.
[13]
U. Morgner, W. Drexler, F. X. Kartner, X. D. Li, C. Pitris, E. P. Ippen, and J. G. Fujimoto,
"Spectroscopic optical coherence tomography," Optics Letters, vol. 25, pp. 111-113, 2000/01/15
2000.
[14]
R. Leitgeb, M. Wojtkowski, A. Kowalczyk, C. K. Hitzenberger, M. Sticker, and A. F. Fercher,
"Spectral measurement of absorption by spectroscopic frequency-domain optical coherence
tomography," Optics Letters, vol. 25, pp. 820-2, 2000/06/01 2000.
[15]
C. Y. Xu, C. Vinegoni, T. S. Ralston, W. Luo, W. Tan, and S. A. Boppart, "Spectroscopic
spectral-domain optical coherence microscopy," Optics Letters, vol. 31, pp. 1079-1081, Apr
2006.
[16]
C. Vinegoni, J. S. Bredfeldt, D. L. Marks, and S. A. Boppart, "Nonlinear optical contrast
enhancement for optical coherence tomography," Optics Express, vol. 12, pp. 331-341, JAN 26
2004.
[17]
Y. Jiang, I. Tomov, Y. M. Wang, and Z. P. Chen, "Second-harmonic optical coherence
tomography," Optics Letters, vol. 29, pp. 1090-1092, MAY 15 2004.
[18]
B. E. Applegate, C. Yang, A. M. Rollins, and J. A. Izatt, "Polarization-resolved second-harmonicgeneration optical coherence tomography in collagen," Optics Letters, vol. 29, pp. 2252-4,
2004/10/01 2004.
[19]
S. Yazdanfar, L. H. Laiho, and P. T. C. So, "Interferometric second harmonic generation
microscopy," Optics Express, vol. 12, pp. 2739-2745, JUN 14 2004.
[20]
M. V. Sarunic, B. E. Applegate, and J. A. Izatt, "Spectral domain second-harmonic optical
coherence tomography," Optics Letters, vol. 30, pp. 2391-3, 15 Sept. 2005 2005.
[21]
J. P. Su, I. V. Tomov, Y. Jiang, and Z. P. Chen, "High-resolution frequency-domain secondharmonic optical coherence tomography," Applied Optics, vol. 46, pp. 1770-1775, Apr 2007.
[22]
K. D. Rao, M. A. Choma, S. Yazdanfar, A. M. Rollins, and J. A. Izatt, "Molecular contrast in
optical coherence tomography by use of a pump-probe technique," Optics Letters, vol. 28, pp.
340-342, Mar 1 2003.
[23]
B. E. Applegate and J. A. Izatt, "Molecular imaging of endogenous and exogenous chromophores
using ground state recovery pump-probe optical coherence tomography," Optics Express, vol. 14,
pp. 9142-9155, Oct 2006.
[24]
Z. Yaqoob, E. McDowell, J. G. Wu, X. Heng, J. Fingler, and C. H. Yang, "Molecular contrast
optical coherence tomography: a pump-probe scheme using indocyanine green as a contrast
agent," Journalof Biomedical Optics, vol. 11, p. 12, Sep-Oct 2006.
[25]
T. M. Lee, A. L. Oldenburg, S. Sitafalwalla, D. L. Marks, W. Luo, F. J. J. Toublan, K. S. Suslick,
and S. A. Boppart, "Engineered microsphere contrast agents for optical coherence tomography,"
Optics Letters, vol. 28, pp. 1546-1548, Sep 1 2003.
[26]
C. Y. Xu, J. Ye, D. L. Marks, and S. A. Boppart, "Near-infrared dyes as contrast-enhancing
agents for spectroscopic optical coherence tomography," Optics Letters, vol. 29, pp. 1647-1649,
JUL 15 2004.
[27]
A. L. Oldenburg, J. R. Gunther, and S. A. Boppart, "Imaging magnetically labeled cells with
magnetomotive optical coherence tomography," Optics Letters, vol. 30, pp. 747-9, 2005/04/01
2005.
[28]
C. Loo, A. Lin, L. Hirsch, M. H. Lee, J. Barton, N. Halas, J. West, and R. Drezek, "Nanoshellenabled photonics-based imaging and therapy of cancer," Technol Cancer Res Treat, vol. 3, pp.
33-40, Feb 2004.
[29]
H. Cang, T. Sun, Z. Y. Li, J. Y. Chen, B. J. Wiley, Y. N. Xia, and X. D. Li, "Gold nanocages as
contrast agents for spectroscopic optical coherence tomography," Optics Letters, vol. 30, pp.
3048-3050, Nov 15 2005.
[30]
J. Chen, F. Saeki, B. J. Wiley, H. Cang, M. J. Cobb, Z.-Y. Li, L. Au, H. Zhang, M. B. Kimmey,
X. Li, and Y. Xia, "Gold nanocages: bioconjugation and their potential use as optical imaging
contrast agents," Nano Letters, vol. 5, pp. 473-7, 2005/03/ 2005.
[31]
A. L. Oldenburg, F. J. J. Toublan, K. S. Suslick, A. Wei, and S. A. Boppart, "Magnetomotive
contrast for in vivo optical coherence tomography," Optics Express, vol. 13, pp. 6597-6614, Aug
22 2005.
[32]
A. Agrawal, S. Huang, A. W. H. Lin, M. H. Lee, J. K. Barton, R. A. Drezek, and T. J. Pfefer,
"Quantitative evaluation of optical coherence tomography signal enhancement with gold
nanoshells," Journalof Biomedical Optics, vol. 11, p. 8, Jul-Aug 2006.
[33]
A. L. Oldenburg, M. N. Hansen, D. A. Zweifel, A. Wei, and S. A. Boppart, "Plasmon-resonant
gold nanorods as low backscattering albedo contrast agents for optical coherence tomography,"
Optics Express, vol. 14, pp. 6724-6738, Jul 2006.
[34]
A. M. Gobin, M. H. Lee, N. J. Halas, W. D. James, R. A. Drezek, and J. L. West, "Near-infrared
resonant nanoshells for combined optical imaging and photothermal cancer therapy," Nano
Letters, vol. 7, pp. 1929-1934, Jul 2007.
[35]
J. Oh, M. D. Feldman, J. Kim, H. W. Kang, P. Sanghi, and T. E. Milner, "Magneto-motive
detection of tissue-based macrophages by differential phase optical coherence tomography,"
Lasers in Surgery and Medicine, vol. 39, pp. 266-272, Mar 2007.
[36]
T. S. Troutman, J. K. Barton, and M. Romanowski, "Optical coherence tomography with plasmon
resonant nanorods of gold," Optics Letters, vol. 32, pp. 1438-1440, Jun 2007.
[37]
C. M. Pitsillides, E. K. Joe, X. B. Wei, R. R. Anderson, and C. P. Lin, "Selective cell targeting
with light-absorbing microparticles and nanoparticles," Biophysical Journal, vol. 84, pp. 40234032, Jun 2003.
[38]
K. Sokolov, M. Follen, J. Aaron, I. Pavlova, A. Malpica, R. Lotan, and R. Richards-Kortum,
"Real-time vital optical imaging of precancer using anti-epidermal growth factor receptor
antibodies conjugated to gold nanoparticles," Cancer research, vol. 63, pp. 1999-2004, May 1
2003.
[39]
C. Loo, A. Lowery, N. Halas, J. West, and R. Drezek, "Immunotargeted nanoshells for integrated
cancer imaging and therapy," Nano Letters, vol. 5, pp. 709-711, Apr 2005.
[40]
M. A. Choma, A. K. Ellerbee, C. H. Yang, T. L. Creazzo, and J. A. Izatt, "Spectral-domain phase
microscopy," Optics Letters, vol. 30, pp. 1162-1164, May 15 2005.
[41]
C. Joo, T. Akkin, B. Cense, B. H. Park, and J. E. de Boer, "Spectral-domain optical coherence
phase microscopy for quantitative phase-contrast imaging," Optics Letters, vol. 30, pp. 21312133, Aug 15 2005.
[42]
B. J. Vakoc, S. H. Yun, J. F. de Boer, G. J. Tearney, and B. E. Bouma, "Phase-resolved optical
frequency domain imaging," Optics Express, vol. 13, pp. 5483-5493, Jul 11 2005.
[43]
D. C. Adler, R. Huber, and J. G. Fujimoto, "Phase-sensitive optical coherence tomography at up
to 370,000 lines per second using buffered Fourier domain mode-locked lasers," Optics Letters,
vol. 32, pp. 626-628, Mar 2007.
[44]
M. J. C. van Gemert, G. W. Lucassen, and A. J. Welch, "Time constants in thermal laser
medicine .2. Distributions of time constants and thermal relaxation of tissue," Physics in
Medicine and Biology, vol. 41, pp. 1381-1399, Aug 1996.
[45]
T. L. Troy and S. N. Thennadil, "Optical properties of human skin in the near infrared wavelength
range of 1000 to 2200 nm," JournalofBiomedical Optics, vol. 6, pp. 167-176, Apr 2001.
[46]
M. A. Choma, A. K. Ellerbee, S. Yazdanfar, and J. A. Izatt, "Doppler flow imaging of
cytoplasmic streaming using spectral domain phase microscopy," Journal of Biomedical Optics,
vol. 11, pp. -, Mar-Apr 2006.
[47]
A. K. Ellerbee, T. L. Creazzo, and J. A. Izatt, "Investigating nanoscale cellular dynamics with
cross-sectional spectral domain phase microscopy," Optics Express, vol. 15, pp. 8115-8124, Jun
2007.
[48]
E. J. McDowell, A. K. Ellerbee, M. A. Choma, B. E. Applegate, and J. A. Izatt, "Spectral domain
phase microscopy for local measurements of cytoskeletal rheology in single cells," Journal of
Biomedical Optics, vol. 12, pp. -, Jul-Aug 2007.
[49]
B. J. Vakoc, G. J. Tearney, and B. E. Bouma, "Real-time microscopic visualization of tissue
response to laser thermal therapy," JournalofBiomedical Optics, vol. 12, p. 3, Mar-Apr 2007.
[50]
S. Makita, Y. Hong, M. Yamanari, T. Yatagai, and Y. Yasuno, "Optical coherence angiography,"
Optics Express, vol. 14, pp. 7821-7840, Aug 2006.
[51]
D. C. Adler, S. W. Huang, R. Huber, and J. G. Fujimoto, "Photothermal detection of gold
nanoparticles using phase-sensitive optical coherence tomography," Optics Express, vol. 16, pp.
4376-4393, Mar 31 2008.
CHAPTER 5
5.0
3D-OCT Platform for Clinical Gastroenterology
5.1 Motivation
Clinical environments require a degree of reliability and robustness in imaging technology that
presents challenges to translating research from the laboratory bench to the patient's bedside. For OCT
imaging systems, real-time data acquisition and display are needed to provide continuous feedback to the
operator. Reliability, ease of use, and rapid data acquisition are also required to maintain short procedure
times and efficient patient flow. For these reasons, typical benchtop or commercial 3D-OCT systems are
not suitable for clinical use and must be heavily modified prior to conducting human studies in a clinical
environment.
Customized OCT systems have previously been applied to numerous biomedical fields including
ophthalmology [1], cardiology [2], gastroenterology [3], urology [4], and gynecology [5]. One
particularly promising application for ultrahigh speed 3D-OCT imaging is gastroenterology. Previously,
two-dimensional OCT (2D-OCT) imaging of the gastrointestinal (GI) tract has been demonstrated with
high resolution but limited imaging speeds of 0.25 - 20 kHz [6-8]. 3D-OCT would provide a significant
advantage since it allows a greater region of tissue to be analyzed and reduces the risk of failing to detect
an abnormal lesion. Additionally, 3D-OCT enables a wide range of image analysis and visualization
techniques to be applied and results in a more complete characterization of tissue microstructure. 3DOCT imaging of the esophagus has been previously demonstrated using a Fourier domain OCT system
based on a conventional wavelength-swept laser source [9]. The imaging rate was limited to 10 kHz,
however, and the resolvable transverse feature size was restricted to 64 ipm due to coarse spatial sampling.
Since the earliest GI lesions typically appear as structures that are -100 [tm in diameter [10], this
performance is not suitable for detecting early cancers. Furthermore, the slow imaging rate resulted in a
long acquisition time (5.8 minutes) and the presence of significant motion artifacts.
In this section of the thesis work an FDML laser was interfaced to a prototype high-speed data
acquisition and data processing system built by Joseph Schmitt at LightLab Imaging Inc. This system
enabled 3D-OCT clinical studies in the GI tract at the unprecedented imaging speeds supported by FDML
technology. The laser and data processing system were modified to provide an optimal combination of
imaging speed, ranging depth, axial resolution, and sensitivity. This technology development effort
produced a 3D-OCT endomicroscopy platform that was used to support clinical pilot studies of a range of
human pathologies. The system will continue to serve this purpose for the foreseeable future.
5.2 System Description
The system constructed here was a swept-source OCT configuration using an FDML laser as the light
source. Data acquisition, real-time processing, and display was carried out by a computer system designed
by LightLab Imaging Inc. Post-processing and image visualization was performed using a combination of
purpose-built Matlab software and a commercial 3D rendering package (ResolveRT, Mercury Computer
Systems). The imaging probes used to conduct clinical studies are a critical aspect of the system and are
described fully in Chapter 6. Preliminary 3D-OCT data used to guide development of the system, signal
processing, and visualization methods was acquired using a modified commercial cardiovascular probe,
also described in Chapter 6.
To be practical in a clinical setting a 3D-OCT endomicroscopy system must provide continuous realtime data capture and display with dense spatial sampling. Extremely large datasets must be digitized,
transferred to computer memory, processed, and displayed in milliseconds. The OCT imaging engine
developed here operated at sustained speeds compatible with FDML lasers by using a novel optical
frequency clock (OFC) integrated with a 200 Msample/s 12 bit analog-to-digital converter (ADC). The
OFC circuitry was designed and constructed by LightLab Imaging Inc. OFC optimization for use with an
FDML laser was carried out as part of this thesis work as described in Section 5.4.
The formation of OCT images from frequency domain interference fringes requires the interference
signal to be evenly spaced in optical frequency v prior to Fourier transformation[11, 12]. Since the
frequency sweeps of both conventional and FDML lasers are not linear in time, it is necessary to correct
for this effect. One common approach is to acquire a reference interference fringe simultaneously with the
OCT signal. The v evolution of the reference signal can be analyzed and used to later correct the v spacing
of the OCT signal[l 1, 13]. This has the disadvantages of doubling the required DAQ bandwidth since two
signals must be captured, requiring signal post-processing to correct the OCT data and limiting real-time
operation.
The OCT imaging engine developed here used an optical triggering technique to automatically correct
for nonlinear v spacing during data acquisition without storing a separate reference signal. Figure 5.1
shows a schematic of the system. 5% of the laser output was routed to an asymmetric Mach-Zehnder
interferometer (MZI) that produces interference fringes with zero crossings evenly spaced in v. This
concept is shown in the left inset of Figure 5.1. The MZI fringes are detected by a dual balanced
photoreceiver and the zero crossings are identified by an analog voltage comparator in the clock
generator, creating a digital pulse train that functions as an OFC. The OFC is synchronized to the start of
each laser sweep by a trigger signal from the FDML laser.
A dual-balanced Michelson interferometer consisting of a pair of optical circulators and a 50/50
fiberoptic splitter is used to generate the OCT interference fringes. Dual-balanced detection cancels
excess noise, reducing the dynamic range requirements of the DAQ system and improving sensitivity.
The same frequency sweep from the laser generated both the MZI and OCT fringes, so the spacing of the
OFC pulses corresponded to evenly spaced v intervals in the OCT signal. The OFC triggers the 12 bit,
200 MSample/s, circularly buffered ADC that samples the OCT fringes. Since the resulting signal is
evenly spaced in v, only one signal needed to be digitized and stored and no post-processing was required
for v correction. The sample arm of the Michelson interferometer included a Patient Interface Unit (PIU),
supplied by LightLab Imaging Inc., that produced high-speed rotational motion at up to 80 Hz and linear
pullback motion at 0.5 - 2.0 mm/s. The PIU attached to the proximal end of the imaging probe and
produced a cylindrical spiral scan at the distal tip of the probe.
After digitization the OCT signal was continuously streamed to the computer RAM over a PCI-X link
at an average of 46 MB/s and a peak of 150 MB/s. Hamming windowing and a fast Fourier transform
were then performed in software to synthesize the image. Each frame was interpolated into polar
coordinates and displayed as a radial image in real-time at > 20 frames/s. During 3D acquisition,
sustained acquisition rates of 100 kHz were achieved while maintaining real-time display.
5.3 Laser Design and Optimization
The FDML laser source was a key component of the 3D-OCT endomicroscopy system. Laser
performance directly affects imaging performance and therefore careful selection of each FDML design
parameter was necessary to ensure optimal clinical utility. Figure 5.2 shows a schematic diagram of the
non-buffered FDML laser used in the 3D-OCT endomicroscopy system. The cavity length was 3.4 km,
giving an FFP-TF drive frequency of 59 kHz and a bi-directional sweep rate of 118 kHz. Only the
backward sweep was used to acquire data, however, giving an effective sweep rate of 59 kHz. It is
possible to operate the system at effective sweep rates of up to 100 kHz but, as described in Section 5.3.1,
a lower sweep rate resulted in optimal imaging performance for the human GI tract. It is also possible to
use a buffered FDML laser to obtain unidirectional sweeps at a high duty cycle, but due to the data
acquisition limitations described in Section 5.3.2 a non-buffered laser was the more optimal design.
The fiber Fabry-Perot tuneable filter (FFP-TF) used in the FDML laser was acquired from
LambdaQuest. The FFP-TF had a finesse of -600 and a free spectral range (FSR) of -190 nm, giving a
linewidth of -0.3 nm. The semiconductor optical amplifier (SOA) chips were high-gain, broadband,
polarization-sensitive devices providing a gain bandwidth of > 100 nm (full width at half maximum,
FWHM) and amplified spontaneous emission (ASE) power of > 5 mW. The output coupling ratio was
chosen to provide maximum tuning range while maintaining a post-booster output power that was
sufficiently high to reach the tissue exposure limit of 20 mW after transmission through the OCT
interferometers and imaging probe. With a 50% output coupler, the total tuning range was 180 nm and the
average output power after booster amplification was - 50 mW. Power output directly from the cavity
was -5 mW.
Figure 5.3 shows a series of point spread functions (PSF) measured by connecting the FDML laser to
an unbalanced MZI and varying the path imbalance. Data was collected using a digital oscilloscope and
only the low-noise backward sweep was used for analysis. Sweep recalibration and Fourier
transformation was performed offline using Matlab software with the method described in Chapter 3. A
reference fringe at an MZI delay of -250 Rm was used to recalibrate all of the other fringes in the dataset.
It is generally not advisable to use a long MZI delay (> 1 mm) to generate reference fringes for
recalibration due to decreased coherence length and fringe contrast at larger path imbalances. Use of
extremely short delays (< 100 ptm) is also not advisable due to an insufficient number of fringe cycles to
calculate an accurate phase evolution curve. This measurement gives the axial resolution and sensitivity
rolloff supported by the laser without taking into account the effects of the OFC and real-time data
processing algorithms used in the complete 3D-OCT endomicroscopy system. Sensitivity rolls off by 6
dB at an imaging depth of 1.6 mm in air and by 10 dB at 2.8 mm in air. Figure 5.4 shows an enlarged
view of a single PSF at an imaging depth of 65 gim. The FWHM of the PSF is 6.4 tm on a linear scale,
corresponding to an axial resolution of 4.6 gm in tissue. It should be noted that the sensitivity rolloff and
axial resolution specifications for the entire system are slightly degraded compared to the values
supported by the laser itself, as described in Section 5.5, due to the effects of the OFC and real-time data
processing methods.
5.3.1 Sweep Rate
The sweep rate of the FDML laser directly determines the imaging speed of the 3D-OCT
endomicroscopy system. Although higher imaging speeds are desirable to maximize the field of view and
minimize motion artifacts, imaging speed also trades off against ranging depth and sensitivity. As the
sweep rate is increased, the frequency associated with the interference fringe generated at a fixed imaging
depth increases in direct proportion. Since the 3D-OCT endomicroscopy system has a fixed ADC with a
maximum sampling rate of 200 MSamples/s, the maximum interference fringe frequency that can be
sampled without aliasing is 100 MHz.
From Equation (1.1) in Chapter 1, the interference fringe signal idet (t) generated at the photodetector
for a fixed point reflector at zo is:
det
(t)
cos(2k(t)z +0)
(5.1)
If the FDML sweep is linear in k between values ko and k, ,bidirectional, and has a backward sweep
duration of Tback , then the variation k (t) for the backward sweep only is:
(k, -k)(5.2)
k(t) = ko +
T
(5.2)
t
back
The photocurrent can then be rewritten as:
idet (t)
S 2z (k,-k
Tback
cos 2koz o +
This gives the interference fringe frequency fdet =
fdet
=
de
)
(5.3)
(5.3)
+
/ 2r as:
(5.4)
z(k, - ko )
CTback
The FDML laser used in the 3D-OCT endomicroscopy system tuned between 1225 nm and 1405 nm.
6 This gives values of 4.47x10 6 m -' and 5.13x10 m ' for ko and k, respectively. The maximum allowable
fringe frequency of 100 MHz therefore occurs at z o =476[m/s]-Tback. If the backward sweep is
assumed to occupy 50% of the total FFP-TF drive period, a drive frequency of 59 kHz gives a fringe
frequency of 100 MHz at a range of 4 mm in air. Since most applications require a ranging depth of at
least 1.5 mm in tissue in addition to the 1.25 mm radius of the imaging probe (total range of 3.8 mm in
air), an effective sweep frequency of 59 kHz is a good operating point. The number of k samples acquired
during each sweep can be estimated as N = 2 Tback fclk where flk
= fet
(z 0 = Zmax
) is the OFC fringe
frequency at an MZI delay of Zmax equal to the maximum ranging depth. Note that this is a rough
approximation and assumes a linear sweep with no wavelength dependence in coherence length. As
discussed in Section 5.4, there is increased loss of fringe contrast at the edges of the sweep which
significantly reduces the number of acquired points at large MZI delays.
Sensitivity is also affected by sweep rate, as shown in Equation (1.6) in Chapter 1. Assuming a
detector efficiency of 50%, incident power on the sample of 20 mW, and centre wavelength of 1315 nm,
the theoretical sensitivity for a 59 kHz filter drive frequency and a 50% backward-sweep duty cycle is 118 dB in the shot noise limit. In reality, excess loss in the sample arm and Michelson interferometer can
be expected to reduce the sensitivity by -5 dB. Imaging in highly-scattering biological tissue such as GI
mucosa requires higher sensitivities than are typically necessary in ophthalmic OCT applications.
Sensitivities of -100 dB or better are necessary to obtain high-quality images, so 59 kHz was also found to
be a good operating point from a sensitivity perspective.
5.3.2 Buffered versus Non-Buffered Cavity Selection
As described in Chapter 2, buffered FDML lasers are capable of generating unidirectional wavelength
sweeps with overall duty cycles near 100%. Buffered FDML lasers are excellent choices for data
acquisition systems where little or no "dead time" is required between consecutive wavelength sweeps to
perform data acquisition or signal processing steps. Many non-real-time imaging systems would be
examples of this situation, since they can acquire one long record and perform sweep segmentation offline following conclusion of the data acquisition.
The 3D-OCT endomicroscopy system uses real-time data processing and image display in order to
provide maximum utility to the clinical user. The system requires -4.5 [ts between the end of one sweep
and the beginning of the next sweep to perform data offload between the ADC and the computer, conduct
signal processing, and display the axial line. Because of this requirement for significant dead time, a
buffered FDML design is sup-optimal since the high duty cycle could not be fully utilized. A nonbuffered FDML design was more efficient since the noisy forward sweep could be removed via SOA
modulation and the resulting time window could be used to perform the ADC and processing tasks. The
choice of a non-buffered FDML laser also increased the output power by -2x compared to a buffered
laser due to use of a single output coupler and removal of the external 50/50 splitter prior to the booster
SOA.
The 59 kHz non-buffered FDML laser was initially operated at an overall duty cycle of 90% (ie, laser
output was produced during 90% of the FFP-TF drive period) split evenly between the forward and
backward sweeps. This provided 7.5 jis forward and backward sweeps. After removal of the forward
sweep, the dead time between consecutive backward sweeps was 9.2 [ts. Application of a non-sinusoidal
drive waveform, as described in Section 5.3.3, was used to reduce the dead time to 6.9 [ts by temporally
compressing the forward sweep. This provided enough dead time for ADC and signal processing tasks
and allowed for a reasonable safety margin.
5.3.3 Sweep Linearization
The FFP-TF of an FDML laser can be driven with a non-sinusoidal waveform in order to temporally
shape the output sweep. This method can be used to generate a sweep that is linear in optical frequency v
as described by Eigenwillig et al [14]. Additionally, the drive wave can be skewed to generate a backward
sweep that is longer in duration than the forward sweep. This has the advantage of increasing exposure
time during the low-noise backward sweep, increasing sensitivity. Sweep skewing also enables the intersweep dead time to be optimized for use with the 3D-OCT endomicroscopy system. Finally, increasing
the duration of the backward sweep decreases the interference signal frequency associated with a given
ranging depth. This effectively increases the maximum imaging depth that can be interrogated with a
given ADC sampling rate.
The FDML laser used in this section of the thesis work was linearized and skewed to favour the
backward sweep. First, an ideal FFP-TF displacement waveform was generated by choosing a skew ratio
and linearity region. A 2:1 skew ratio was chosen (ie, backward sweep occupying twice as much time as
the forward sweep) to reduce the inter-sweep dead time to approximately 7 gs for an FFP-TF drive
frequency of 59 kHz. For applications in the 3D-OCT endomicroscopy system, sweep linearity was only
required during the backward sweep. The forward sweep was not used and therefore the exact shape of
the forward section of the FFP-TF drive wave was not important.
Next the frequency response of the FFP-TF was measured by applying a low-amplitude sine wave to
the filter, scanning the frequency, and recording the amplitude and phase of the FFP-TF displacement. An
SOA generating broadband ASE was connected to the FFP-TF input. A portion of the FFP-TF output was
directed to an optical spectrum analyzer (OSA). The width of the transmitted spectrum was used to
measure the amplitude of the FFP-TF response at each drive frequency. The phase of the response was
measured by directing the rest of the FFP-TF output to an unbalanced MZI and observing shifts in the
start time of observed interference relative to the start of the driving sine wave. Figure 5.5 shows the
measured amplitude response of the FFP-TF used in the FDML laser. Clear harmonic peaks are seen near
56 kHz and 163 kHz.
To avoid applying excess electrical power to the FFP-TF, the ideal drive wave was approximated by a
summation of three harmonic sine waves at 59 kHz, 118 kHz, and 177 kHz. These frequencies are near
the resonant peaks of the filter but are not directly on resonance. Driving directly on resonance can lead to
unstable behaviour and filter damage. The sinusoidal fit was carried out using Matlab curve fitting
utilities. The forward portion of the drive wave was weighted at 0% to allow maximum accuracy for the
backward portion. The resulting ideal drive wave is shown in Figure 5.6. This wave represents the desired
response of the FFP-TF to give a linear backward sweep with a 2:1 skew ratio. This desired response is
modified according to the measured amplitude and phase responses at the three harmonic drive
frequencies as shown in Figure 5.5. This correction results in the actual drive waveform shown in Figure
5.7. This drive wave is applied to the filter with minor manual adjustments used to minimize the PSF
width.
The results of the drive linearization are shown in Figure 5.8 and Figure 5.9. Figure 5.8 shows the
amplitude and phase output of an unbalanced MZI using an FDML laser with a linearized, 2:1 skewed
waveform driving the FFP-TF. This data was acquired using a digital oscilloscope. The overall duty cycle
is 88% with the remaining 12% used as a buffer to prevent sweep overlap during DC bias setpoint drift.
The backward sweep, which is used to generate images in the 3D-OCT endomicroscopy system, is > 2x
longer than the forward sweep. The phase evolution of the backward sweep is much more linear than the
forward sweep, although the backward sweep deviates from a linear phase evolution near the end of the
drive period. Figure 5.9 shows Fourier transformed PSF's generated at the same imaging depth using
different FFP-TF drive waves. The blue curve is an "ideal" PSF generated by recalibrating a sinusoidal
wavelength sweep with post-processing software to give a maximally linear v evolution. The red curve is
a non-recalibrated PSF generated with a linear drive wave but no skewing. The PSF width is similar to the
ideal case with a low-frequency tail extending down to 15 MHz. The black curve is a non-recalibrated
PSF generated with a sinusoidal drive wave. The peak RF frequency is 22% higher than the PSF from the
linearized waveform and the bandwidth is -4x larger, indicating inefficient use of ADC time and
unnecessarily high peak RF frequencies. The magenta curve is a non-recalibrated PSF generated with a
linearized and skewed drive wave. The RF peak is shifted down by -2x, although a high frequency tail
extends up to 25 MHz. FFP-TF linearization and skewing gives an optimal combination of low
interference fringe frequency, efficient use of ADC time, and increased sensitivity through increased
sample exposure time.
5.4 Optical Frequency Clock Optimization
The OFC used in the 3D-OCT endomicroscopy system is a unique way to enable real-time imaging at
speeds up to 100,000 axial lines per second by removing the requirement to sample a second interference
signal and recalibrate the OCT data. Since FDML sweep linearization does not result in optimal PSF
shapes as shown in Figure 5.9, the OFC must be used to remove residual v nonlinearity in the sweep to
provide optimal axial resolution and minimal sidelobe levels. Although the OFC is extremely useful it
presents additional performance tradeoffs not typically found in OCT imaging systems. These tradeoffs
are a result of the link between the path imbalance in the OFC MZI used to generate the ADC sample
clock and the imaging depth. First, the OFC introduces a tradeoff between maximum imaging depth and
system noise. Second, it introduces an additional tradeoff between maximum imaging depth and axial
resolution. These tradeoffs are described below.
5.4.1 Tradeoffs Between Imaging Depth and System Noise
As described in Section 5.2, the OFC uses an unbalanced MZI and analog electronics to generate one
ADC clock pulse every time the MZI interference signal crosses 0 V. These zero crossings are evenly
spaced in optical frequency v so the OFC ensures that the OCT signal is sampled at evenly-spaced v
intervals. One side effect of this setup is that the path mismatch in the MZI determines the maximum
imaging depth of the system. As described in Chapter 1, the maximum imaging depth of a swept source
OCT system is determined by the optical frequency spacing between consecutive samples in the
interference signal. The OFC generates exactly two sample clock pulses for every MZI interference fringe
period. Therefore Nyquist-limited sampling is achieved for OCT signal frequencies equal to the MZI
fringe frequency and the MZI path imbalance is equal to the maximum imaging range. For sample ranges
higher than the MZI path imbalance, the OCT signal would be under-sampled and aliasing would occur.
This is prevented by placing a 100 MHz lowpass filter in the ADC circuitry.
Since the system ADC operates at 200 MSamples/s, the MZI would ideally be set to produce 100 MHz
interference fringes to obtain the largest possible imaging depth (4.7 mm in air for the 59 kHz linearized
and skewed FDML laser). However, as shown in Figure 5.2, the SNR of the interference signal decreases
at increased path imbalances. At 4.7 mm the fringe contrast decreases by > 18 dB compared to the
optimal value at short imbalances. This results in a substantial increase in OFC sample clock jitter caused
by detection of false zero crossings at the lower-amplitude start and end of the sweep. Clock jitter
translates into erroneous assignment of v values prior to Fourier transformation, dramatically elevating the
noise floor of the image. Loss of MZI SNR can also cause temporal walk-off of the OCT signal. The
system requires a fixed number of sample points to be acquired during each axial line. If an insufficient
number of clock events are generated due to OFC noise, a portion of the next sweep is stored as the
previous sweep. This also causes artifacts in the resulting OCT image.
In practice, the MZI imbalance was set to a less aggressive value to ensure the generation of low-noise
images. MZI imbalances of 2.4 - 2.5 mm were typically used for human imaging studies. At this range,
the MZI fringe contrast is decreased by less than 10 dB compared to the optimal value. No signal walkoff or noise floor fluctuations occurred at this setpoint. An imaging range of 2.4 mm in air corresponds to
1.7 mm in tissue, which is sufficient to ensure that penetration depth is limited by scattering and
absorption in the tissue rather than the system parameters.
5.4.2 Tradeoffs Between Imaging Depth andAxial Resolution
The MZI path imbalance, and therefore the maximum imaging depth, also affects the axial resolution
of the 3D-OCT endomicroscopy system. This arises fundamentally from the finite coherence length of the
FDML laser and the presence of chromatic dispersion in the long fiber cavity. As described in Chapter 2,
the drive period of the FFP-TF in an FDML laser can be precisely synchronized to the optical roundtrip
time of light in the cavity at only one wavelength. This wavelength is typically selected to be in the
middle of the sweep. Chromatic dispersion causes a slight desynchronization between the FFP-TF and the
light at the edges of the sweep. In addition to increasing cavity losses at those wavelengths, this also
reduces the number of times the FFP-TF is effectively propagated through by the photons at the sweep
edges. Reducing the number of FFP-TF transmission cycles results in a broadened instantaneous
linewidth at the blue and red edges of the sweep, reducing the edge coherence lengths and thereby
reducing the fringe contrast at large delays for these wavelengths.
The tradeoff between imaging depth and axial resolution caused by these effects is illustrated in Figure
5.10 and Figure 5.10. When the MZI imbalance is set to 1.0 mm, the OFC MZI generates a fringe that
covers all of the wavelengths contained in the FDML sweep. ADC clock events are therefore triggered
over a time duration spanning the entire OCT signal (shown in magenta for a short imaging depth) and the
entire tuning range of the laser is utilized. When the MZI imbalance is set to 4.5 mm, however, the
imbalance exceeds the coherence length of the blue edge of the sweep. The MZI fringe duration decreases
and no clock events are generated during the last 25% of the sweep. Therefore the OCT signal is not
entirely sampled and 25% of the tuning range is unused. This decreases the spectral content of the
sampled OCT signal and reduces axial resolution by a corresponding 25%.
In practice, as discussed in Section 5.4.1, the MZI path imbalance was set to 2.4 - 2.5 mm. At this
distance the axial resolution decreases by only -10% due to MZI fringe dropout. Further degradation of
the axial resolution is caused by jitter in the OFC that tends to broaden the PSFs by sampling the OCT
signal at unevenly spaced optical frequency positions. The axial resolution of the system was measured as
described in Section 5.5.1.
5.5 System Characterization
The specialized OFC, ADC, and signal processing software used in the 3D-OCT endomicroscopy
system significantly affect the overall performance. Axial resolution, sensitivity and sensitivity rolloff of
the overall system are significantly different than the values supported by the FDML laser itself. The true
resolvable feature size or "true resolution" of the system is also a function of the spatial sampling density
in addition to the optical resolution determined by the laser tuning range and imaging probe spot size.
These performance metrics were characterized for the overall system as described below.
5.5.1 Axial Resolution
Axial resolution was measured by replacing the imaging probe with an equivalent length of SMF-28
fiber, bulk collimating lens, neutral density filter, and metallic mirror. The mirror provided a point
reflection at an adjustable position away from the fiber tip while the filter provided 20 dB of single-pass
attenuation to prevent detector saturation. The OFC MZI was set to a path mismatch of 2.45 mm
corresponding to the maximum imaging range typically used for clinical studies. The resulting OCT
interference signal was captured by the ADC and converted to a linear PSF in real time using the system's
signal processing software. The linear PSF at an imaging depth of 70 gm is shown in Figure 5.12. The
FWHM of the linear PSF is 8.0 tm in air compared to 6.4 pm in air as measured with the FDML laser
alone. The corresponding resolution in tissue is 5.8 pm for the complete system compared to 4.6 [pm for
the laser alone. As discussed above, the degradation is caused partially by OFC clock dropout at the blue
edge of the MZI fringe and partially by OFC jitter.
5.5.2 Sensitivity Rolloff
System sensitivity was measured using the same sample arm setup as described in Section 5.5.1.
Sensitivity is obtained by comparing the peak signal value at a short imaging depth to the system noise
floor with the sample arm blocked, correcting for the attenuation of the neutral density filter. With 15 mW
of power (backward sweep only) incident on the sample a sensitivity of -113 dB was recorded. The
theoretical shot-noise limited sensitivity for a backward sweep duration of 9.9 ls, 15 mW of power on the
sample, and 50% detector sensitivity is -117 dB from Equation 1.8. Single-pass losses in the sample arm
and Michelson interferometer of -2 dB account for the variation in sensitivity away from the shot-noise
limited value.
Sensitivity rolloff for the complete system was measured by translating the mirror away from the fiber
tip and recording a PSF at each new position. The results of this measurement are shown in Figure 5.13.
Compared to the rolloff induced by the laser itself, the system rolloff is slightly more severe. The 6 db
rolloff position is at 1.2 mm for the complete system compared to 1.6 mm for the laser only. This increase
in rolloff is also due to OFC clock dropout and jitter.
5.5.3 True Spatial Resolution
When characterizing the resolution of a 3D-OCT system it is important to differentiate between optical
resolution, spatial sampling density, and true spatial resolution. Optical resolution, determined by the
focal spot size of the sample optics and the width of the axial point spread function, defines the best
resolution that the system can theoretically obtain. Spatial sampling density is the distance between
consecutive axial scans in X and between consecutive frames in Y. Spatial sampling density in Z can be
set arbitrarily by selecting the Fourier transform length in a swept source setup. True resolution is
determined by a combination of optical resolution and spatial sampling density, and defines the smallest
feature size that can actually be visualized. It should be noted that optical resolution is not constant with
imaging depth, since the beam diverges away from the focal position. In addition, the ability to resolve
structures depends on feature contrast.
The Nyquist criterion requires at least two spatial samples in each dimension for every optical
resolution element. For the OCT system reported here the optical resolution in tissue is 12 x 12 x 6 lIm, so
spatial samples should be acquired every 6 x 6 x 3 htm. The XY spatial sampling density is fundamentally
limited by imaging speed, tissue surface area, and maximum imaging duration. Since surface area and
imaging duration are determined by anatomy and physiology, the only parameter available for increasing
spatial sampling density is imaging speed. Therefore high imaging speeds, as determined by the laser
sweep rate and data acquisition capacity, are required to obtain true microscopic 3D-OCT resolution.
The XY spatial sampling density of the 3D-OCT endomicroscopy system constructed here could be
adjusted by varying the probe rotation rate and pullback speed. To obtain a reasonable tradeoff between
spatial sampling density and field of view, the rotational speed was typically set to 60 Hz and the pullback
speed to 1.0 mm/s. With axial image lines acquired at 59 kHz, this gave 983 lines per frame with 16.7 jim
between consecutive frames. The transverse pixel spacing was approximately 8 jim, given an average
imaging depth of 1240 jLm in tissue (70% of the maximum imaging depth) and a corresponding
circumferential scan length of 7.8 mm. The actual transverse pixel spacing varies with depth due to the
rotational nature of the probe. Axial pixel spacing was 3.5 jm with a 1024 point Fourier transform
performed on 864 k-space samples per line. These parameters gave a true spatial resolution of 16 x 33 x 7
3
jtm (3696 jim3) compared to an optical resolution of 12 x 12 x 6 jm (864 jim ).
In the future, faster imaging speeds and increased probe rotational rates will be necessary to make full
use of the optical resolution supported by the FDML laser and imaging probes. The current performance
nevertheless represents a significant improvement over previous endoscopic swept source OCT systems,
which used polygon mirror based swept lasers and achieved optical resolutions of 15 x 15 x 7 jtm with
XY spatial sampling densities of only 25 x 33 jim due to restricted imaging speeds of 10 kHz [9, 15].
3
Consequently, the true resolution of these previous systems was 50 x 66 x 7 gm (23,100 gim ) at imaging
speeds of 10 kHz in the esophagus and 54 kHz in the coronary artery.
5.6 Data Post-Processing
Several data post-processing algorithms were developed in order to achieve optimal visualization of
tissue microstructure. Since the mucosal structure of the GI tract is difficult to evaluate in a cylindrical
form, the individual 3D-OCT frames must be converted from radial images to rectangular image. To
facilitate the formation of enface images at arbitrary depths, the frames must then be flattened to prevent
curvature artifacts. Finally, the frames must be stored as compressed JPEG files to prevent memory
overflow in the commercial 3D rendering software used to analyze the datasets. These steps are described
in Sections 5.6.1 and 5.6.2 below.
5.6.1 FrameFlattening
The frame flattening algorithm is shown in Figure 5.14. Rectangular blocks indicate automated steps
while trapezoidal blocks indicate steps requiring user input. Diamond-shaped blocks are decision steps
and cylindrical blocks are data operations. The flattening algorithm works by detecting the imaging probe
sheath's irregular surface and then shifting each axial line of each frame such that the sheath becomes flat.
Since the sheath is nominally in direct contact with the tissue surface, this simultaneously flattens the
tissue as well.
100
The algorithm consists of three main sections as shown by the three main columns in the algorithm
flowchart. First, the software obtains user input for several processing parameters. Since each frame is
stored as a series of axial lines in the LightLab software (as opposed to a radial image), no processing is
required to convert the radial frame to a rectangular frame. Each frame is stored as a series of Fourier
transformed logarithmic image lines with 12 bit resolution. Three representative test frames are parsed
from the beginning, middle, and end of the 3D dataset in order to obtain the required user input. The
frames are displayed as rectangular Matlab figures. The. user is first asked to set an axial exclusion range
to remove fixed line artifacts from the image. These artifacts are typically the result of a small reflection
from an optical surface in the proximal section of the probe and appear as moderately bright lines inside
the probe lumen. All pixels located inside the exclusion range are set to zero value to avoid erroneous
edge detection.
Next the user is asked to input a horizontal wrapping point for the image. The GI tissue is not typically
in contact with the probe over the entire 360' rotation of a single frame. The horizontal wrapping point
sets the position for the first axial line of the rectangular dataset and allows the user to centre the useful
portion of the rotation in the middle of the frame. Next the user is given the option to set up small
rectangular masks to remove any residual artifacts not covered by the axial exclusion range that may
corrupt the sheath detection. At this point the algorithm applies a 6 x 3 pixel Gaussian smoothing filter to
remove speckle noise in the three representative test frames. The user is shown each filtered frame and
asked to provide a noise threshold level such that the sheath edge reflection remains above the threshold
but all structures at more shallow axial pixels are set to zero. This is the final piece of manual input
required for the algorithm.
Next the post-processing software sequentially reads in each frame and attempts to detect the sheath
boundary. This is done by removing the axial exclusion zone and masks, horizontally shifting the frame
according to the user-defined wrapping point, applying the Gaussian filter, and thresholding out pixel
values below the user-defined noise level. Then a Sobel edge detector is applied to the frame to generate a
vector of edge points for each axial line. Sobel edge detection is a rapid, robust method for locating edges
based on 2D image information. The sheath is detected by isolating the first edge point in each line. For
lines where no edge is detected, the missing sheath location is estimated by interpolating between
neighbouring lines. Finally a
12 th
order polynomial fit is applied to the sheath edge in order to smooth out
abrupt intra-frame jumps caused by the Sobel algorithm. After the sheath location has been stored for
each frame, a 10-frame rolling average is applied to smooth out inter-frame jumps.
101
5.6.2 JPEG Conversion
The final section of the algorithm shifts each axial line in each frame and saves the output as a stack of
JPEG files. Each frame is sequentially read in a second time. The stored sheath location for the frame is
recalled and each axial line is axially shifted such that the sheath, and therefore the top surface of the
tissue, becomes horizontally oriented. No thresholding or Gaussian filtering is applied at this stage in
order to preserve the original image data. A Matlab figure with the correct number of pixels is created and
saved as a high-quality 3-channel 24 bit JPEG with each channel encoding an 8-bit colour level. Once this
process is completed for each frame the 3D-OCT data file is closed and the program terminates.
In order to visualize the 3D-OCT data using a commercial rendering package, the high-quality 24-bit
JPEGs created by the flattening algorithm must be compressed to lower-quality versions. The rendering
package used throughout this thesis work (ResolveRT, Mercury Computer Systems) is not capable of
manipulating - 1000 x 512 x 1200 datasets when each pixel is a 24-bit value. A second JPEG stack is
therefore created that consists of compressed 8-bit data. The greyscale image quality is very similar to the
original 24-bit data since the human eye has difficulty distinguishing more than 256 levels of grey.
5.7 Image Visualization
3D-OCT imaging enables the employment of a variety of powerful visualization methods in order to
comprehensively assess tissue microstructure. Full 3D renderings of the dataset can be created to inspect
the tissue at any spatial orientation. Individual 2D image planes can also be defined, enabling arbitrary
cross-sectional analysis of the tissue. For example, en face sections can be produced to provide a virtual
microscopic view of the tissue at any depth. Cross-sectional images can then be generated with exact
registration to en face features and anatomic landmarks. Additionally, consecutive 2D images from a
densely sampled 3D dataset can be averaged to reduce speckle noise and improve the visualization of
subtle tissue structures. Quantitative 3D measurements can also be performed on tissue structures in order
to assess pathologic state. In general, 3D-OCT imaging provides a more complete understanding of the
tissue architectural morphology compared to 2D imaging. The following sections describe several
visualization methods developed as part of this thesis work. Preliminary data was acquired in the rabbit
colon using a modified commercial cardiovascular imaging probe.
5.7.1 3D Rendering
3D rendering enables inspection of the imaged tissue from arbitrary orientations and is also helpful for
orienting individual 2D image planes within the dataset. Figure 5.15 illustrates the generation of 3D-OCT
renderings from a stack of 885 radial frames. Figure 5.15(A) shows a single radial frame acquired in the
rabbit colon, while Figure 5.15 (B) shows a cutaway view of the complete rendered dataset. No unfolding
102
or frame flattening was applied while generating the rendering shown in Figure 5.15 (B) so the data retains
the cylindrical shape of the original acquisition. The cutaway view enables visualization of the luminal
surface, with crypts appearing as dark features surrounded by bright white bands of lamina propria. The
enlarged view of the epithelium shown in Figure 5.15 (C) illustrates crypt structure in a single frame.
Figure 5.15 (D) shows a 3D rendering formed after unfolding and flattening the dataset as described in
Section 5.6. The unfolded visualization enables clearer appreciation of tissue morphology and is more
amenable to comparisons with endoscopic images and biopsy-based histology.
5.7.2 OrthoplaneSectioning
After application of the unfolding and flattening algorithm, cross-sectional OCT images with arbitrary
orientations can be generated that are precisely registered to the surface of the tissue. These orthoplanes
can be scanned through the tissue volume at any location, facilitating detailed inspection of tissue
microstructure. Figure 5.16 illustrates the concept of orthoplane analysis for the rabbit colon dataset shown
in Figure 5.15(D). Figure 5.16(A) shows an en face XZ orthoplane image located in the epithelial tissue
layer of the colonic mucosa. The mottled appearance of the image is a result of the colonic crypts in the
epithelium. Orthogonal cross-sectional images showing depth-resolved structure, such as the YZ image
shown in Figure 5.16(B), can be aligned to surface features for enhanced analysis of tissue microstructure.
The layered nature of the mucosa can be clearly appreciated in the YZ cross-sectional orthoplane.
5.7.3 Projection Viewing
Since the 3D-OCT data set is sampled with a high spatial density, consecutive cross-sectional or en
face slices can be averaged or "projected" to reduce speckle noise without significantly blurring image
features. This concept is illustrated in Figure 5.17. The image in Figure 5.17(A) was formed by calculating
the mean of 7 consecutive YZ slices, equivalent to averaging the data over a 21 jtm thick section. A 21
[tm section is smaller than the dimension of two epithelial cells in the rabbit colon, so tissue
microstructure is largely constant and minimal image blurring is observed compared to the single image
in Figure 5.16(B). Since the speckle size is approximately equal to the 9 jtm focal spot size of the probe
used to collect this data, the speckle pattern is decorrelated over the 21 jtm section. Averaging a
decorrelated speckle pattern provides significant rejection of speckle noise and enhances tissue contrast.
Figure 5.17(B) shows a 3x enlarged view of a region of Figure 5.17(A). The 3D-OCT image correlates
well with representative histology of colonic tissue from the same animal (Figure 5.17(C)), although the
mucosal surface is flattened in the 3D-OCT images due to compression by the imaging probe. Variations
in apparent layer thickness are due to probe compression as well as tissue shrinkage that occurs during
103
histology processing. The dark vertical bands in the 3D-OCT image are caused by fecal material on the
luminal surface.
5.7.4 Linear En FaceImages
3D-OCT data is typically stored and visualized using a logarithmic scale in order to display the full
dynamic range (> 50 dB) of the data acquired over a typical tissue depth of -1000 [m. Logarithmic
compression of the data, however, has the unwanted side-effect of reducing the effective transverse
resolution in enface images. This occurs because the system's transverse point spread function, defined
by the width of the incident OCT beam, is broadened when a logarithmic transform is applied. This
makes small transverse features appear wider than their true size.
A comparison between linear and logarithmically scaled en face images is shown in Figure 5.18. Both
images were generated using the same dataset by averaging the data axially over a 20 tm range centered
at a tissue depth of 400 tm. In the linear image shown in Figure 5.18(A) the transverse crypt features
appear sharper and more defined than in the logarithmic image shown in Figure 5.18(B). Due to the lower
dynamic range of the linear image, however, colour map saturation artifacts are clearly visible at the right
hand side of the linear image (red arrows). Conversely, the logarithmic image is able to display the wider
dynamic range of the 3D-OCT data and does not produce equivalent artifacts. For most GI imaging
applications a logarithmic pixel compression is desirable to allow visualization of the complete range of
pixel intensities found in highly scattering biological media.
104
5.8 Figures
Sample
RM
1C 2
3
E
N
time
Av
Av
TRG
+DA-
+DA-
OCT
OFC
OFC
OCT
Custom 200 MHz
DAQ / DSP System
Personal
Computer
Figure 5.1. 3D-OCT endomicroscopy system schematic. C, circulator. PIU, patient interface unit. RM,
reference mirror. MZI, Mach-Zehnder interferometer. P, photodetector. DA, differential amplifier. MZI,
Mach-Zehnder input. TRIG, sweep trigger input. OFC, optical frequency clock input. OCT, OCT signal
input. RAM, random access memory. FFT, fast Fourier transform. Left inset shows the principle of OFC
generation using the MZI output. Zero crossings are unevenly spaced in time, but evenly spaced in optical
frequency v.Originally published in [16].
PSF Rolloff: 60 kHz FDML Laser
0
-FDML laser used in 3D-OCT
Figure 5.2.
endomicroscopy system. FFP-TF, fiber Fabry-Perot
tunable filter. ISO, isolator. SOA, semiconductor
optical amplifier.
105
3
2
Imaging Range [mm]
1
4
Figure 5.3. PSF rolloff measured using a 60
kHz FDML laser and an unbalanced MZI. 6
dB rolloff point is at 1.6 mm. Originally
published in [16].
FFP-TF Amplitude Response
Linear PSF: 60 kHz FDML Laser
20
E
*--15-
,j
ir
,.10
0.04
0.1
0.08
0.06
Imaging Depth [mm]
04
0.12
60
Applied FFP Drive Wave
FFP Drive Waveform for Linear Sweep
1.5
1
Time
05
180
160
Figure 5.5. Amplitude response of an FFP-TF as
the drive frequency is varied over three regions.
Clear harmonic peaks are observed near 56 kHz
and 163 kHz.
Figure 5.4. Linear PSF measured with 60 kHz
FDML laser at a short imaging depth. Axial
resolution is 6.4 um in air or 4.6 um in tissue.
-1'
0
80 100 120 140
Frequency [kHz]
I
x 10
Time
x104
Figure 5.7. Actual drive wave applied to FFP-TF
after correcting for amplitude and phase
response of filter.
Figure 5.6. Desired FFP-TF response for linear,
skewed sweep in the backward direction.
FFP-TF Amplitude Response
FDML Output Using Skewed Linear Drive Wave
2-
240
60% Backward
0% Forward
Sweep
I %4 ~
Sweep
E
C
2M6
1200
MI
C
q
.0
E
00
no
-14
-12
-10
-8
-6
-4
-2
0
2
Time [us]
Figure 5.8. Amplitude and phase of interference signals
generated by FDML laser and unbalanced MZI after
application of linearized and skewed FFP-TF drive
wave.
106
C
5
10
30
25
20
15
Frequency [MHz]
35
40
Figure 5.9. PSFs generated at the same
imaging depth for various FFP-TF drive
waveforms. Ideal PSF is obtained with postprocessing recalibration. Other PSF's were
not recalibrated.
Figure 5.10. U-;u
-igure 5.11.
signal durations wnen MLI
t-L; signal aurations wnen MLI
imbalance is set to 4.5 mm. Clock events are
generated for only a portion of the sweep. 25% of
the OCT signal is not sampled, reducing the axial
resolution.
imbalance is set to 1.0 mm. Clock events are
generated for the entire sweep duration (red
bars) and the entire OCT signal can be sampled.
PSF Rolloff: Complete System
Linear PSF: Complete System
1
,0.8
EA 0.6
0.4
E __
.4
80
60
Imaging Depth [um]
1.5
1
0.5
Imaging Range [mm]
120
Figure 5.13. PSF rolloff measured with complete
3D-OCT endomicroscopy system. 6 dB rolloff
point is at 1.2 mm.
Figure 5.12. Linear PSF measured with complete
3D-OCT endomicroscopy system at a short
imaging depth. Axial resolution is 8.0 um in air or
5.8 um in tissue.
107
Figure 5.14. Flowchart of 3D-OCT data post-processing algorithm for frame flattening and JPEG
conversion. Frames are flattened by detecting the probe sheath and shifting each axial line such that the
sheath and tissue surface become horizontally oriented.
108
Figure 5.15. Construction of 3D-OCT renderings by stacking a series of radial frames. A: Single radial
frame of rabbit colon. B: Cutaway view of rendering formed by stacking 885 frames. C: Enlarged view of
epithelium in A. D: 3D rendering of unfolded dataset after frame flattening. Originally published in [16].
109
Figure 5.16. Single XZ and YZ orthoplanes near the middle of the dataset shown in Figure 4.15. A: En
face XZ orthoplane showing details of colonic epithelium in the rabbit. Crypt structures are visible
between bright bands of lamina propria. Red arrow pairs indicate location of cross-sectional image in B.
B: Longitudinal YZ orthoplane near the centre of the tissue volume. The layered mucosal structure can be
appreciated over a range > 1000 um in tissue. Red arrow pairs indicate location of en face image in A.
AC
LI
M
SM
MC
MP
ML
SE
AC: Absorbing Cells
CM: Colonic Mucosa
LP: Lamina Propria
MM: Muscularis Mucosa
SM: Submucosa
MC: Circular Muscle
MP: Myenteric Plexus
ML: Longitudinal Muscle
SE: Serosa
FT: Fatty Tissue
LP
MM
SM
MP
ML
Figure 5.17. Averaging or "projecting" images over a thin section reduces speckle noise and improves
contrast. A: Longitudinal YZ image formed by averaging 7 consecutive cross-sectional images over a
span of 21 um. Speckle is reduced compared to the single image shown in Figure 5.16(B). B: Enlarged
view of A showing all colonic layers. C: comparison to histology from excisional biopsy specimen.
Originally published in [16].
110
Figure 5.18. Comparison of linear and logarithmically scaled en face images of the human colon. A:
Linear scaled en face image at a tissue depth of 400 um averaged over a range of 20 um. Saturation
artifacts are visible (red arrows). B: Logarithmically scaled en face image at the same location, also
averaged over a range of 20 um. Features appear sharper in the linear image but saturation artifacts
appear due to insufficient dynamic range.
111
5.9 References
[1]
W. Drexler, U. Morgner, R. K. Ghanta, F. X. Kirtner, J. S. Schuman, and J. G. Fujimoto,
"Ultrahigh-resolution ophthalmic optical coherence tomography," Nature Medicine, vol. 7, pp.
502-507, Apr 2001.
[2]
G. J. Tearney, M. E. Brezinski, S. A. Boppart, B. E. Bouma, N. Weissman, J. F. Southern, E. A.
Swanson, and J. G. Fujimoto, "Catheter-based optical imaging of a human coronary artery,"
Circulation,vol. 94, p. 3013, 1996.
[3]
B. E. Bouma, G. J. Tearney, C. C. Compton, and N. S. Nishioka, "High-resolution imaging of the
human esophagus and stomach in vivo using optical coherence tomography," Gastrointestinal
endoscopy, vol. 51(4) Pt 1, pp. 467-74, Apr 2000.
[4]
G. J. Tearney, M. E. Brezinski, J. F. Southern, B. E. Bouma, S. A. Boppart, and J. G. Fujimoto,
"Optical biopsy in human urologic tissue using optical coherence tomography," The Journalof
urology, vol. 157, pp. 1915-9, May 1997.
[5]
P. F. Escobar, J. L. Belinson, A. White, N. M. Shakhova, F. I. Feldchtein, M. V. Kareta, and N.
D. Gladkova, "Diagnostic efficacy of optical coherence tomography in the management of
preinvasive and invasive cancer of uterine cervix and vulva," International Journal of
Gynecological Cancer,vol. 14, pp. 470-474, MAY-JUN 2004.
[6]
A. R. Tumlinson, B. Povazay, L. P. Hariri, J. McNally, A. Unterhuber, B. Hermann, H. Sattmann,
W. Drexler, and J. K. Barton, "In vivo ultrahigh-resolution optical coherence tomography of
mouse colon with an achromatized endoscope," Journal of Biomedical Optics, vol. 11, pp. -,
Nov-Dec 2006.
[7]
P. R. Herz, Y. Chen, A. D. Aguirre, J. G. Fujimoto, H. Mashimo, J. Schmitt, A. Koski, J.
Goodnow, and C. Petersen, "Ultrahigh resolution optical biopsy with endoscopic optical
coherence tomography," Optics Express, vol. 12, pp. 3532-3542, JUL 26 2004.
[8]
A. R. Tumlinson, J. K. Barton, B. Povazay, H. Sattman, A. Unterhuber, R. A. Leitgeb, and W.
Drexler, "Endoscope-tip interferometer for ultrahigh resolution frequency domain optical
coherence tomography in mouse colon," Optics Express, vol. 14, pp. 1878-1887, Mar 6 2006.
[9]
S. H. Yun, G. J. Tearney, B. J. Vakoc, M. Shishkov, W. Y. Oh, A. E. Desjardins, M. J. Suter, R.
C. Chan, J. A. Evans, I. K. Jang, N. S. Nishioka, J. F. de Boer, and B. E. Bouma, "Comprehensive
volumetric optical microscopy in vivo," Nature Medicine, vol. 12, pp. 1429-1433, Dec 2006.
[10]
T. Takayama, S. Katsuki, Y. Takahashi, M. Ohi, S. Nojiri, S. Sakamaki, J. Kato, K. Kogawa, H.
Miyake, and Y. Niitsu, "Aberrant crypt foci of the colon as precursors of adenoma and cancer,"
New EnglandJournalof Medicine, vol. 339, pp. 1277-1284, Oct 29 1998.
112
[1 l]
E. Brinkmeyer and R. Ulrich, "High-Resolution Ocdr in Dispersive Wave-Guides," Electronics
Letters, vol. 26, pp. 413-414, Mar 15 1990.
[12]
M. Wojtkowski, R. Leitgeb, A. Kowalczyk, T. Bajraszewski, and A. F. Fercher, "In vivo human
retinal imaging by Fourier domain optical coherence tomography," Journalof Biomedical Optics,
vol. 7, pp. 457-463, 2002/07/ 2002.
[13]
S. H. Yun, G. J. Tearney, J. F. de Boer, N. Iftimia, and B. E. Bouma, "High-speed optical
frequency-domain imaging," Optics Express, vol. 11, pp. 2953-2963, Nov 3 2003.
[14]
C. M. Eigenwillig, B. R. Biedermann, G. Palte, and R. Huber, "K-space linear Fourier domain
mode locked laser and applications for optical coherence tomography," Optics Express, vol. 16,
pp. 8916-8937, Jun 9 2008.
[15]
B. J. Vakoc, M. Shishko, S. Yun, W. Y. Oh, M. J. Suter, A. E. Desjardins, J. A. Evans, N. S.
Nishioka, G. J. Tearney, and B. E. Bouma, "Comprehensive esophageal microscopy by using
optical frequency-domain imaging," GastrointestinalEndoscopy, vol. 65, pp. 898-905, May 2007
2007.
[16]
D. C. Adler, Y. Chen, R. Huber, J. Schmitt, J. Connolly, and J. G. Fujimoto, "Three-dimensional
endomicroscopy using optical coherence tomography," Nature Photonics, vol. 1, pp. 709-716,
Dec 2007.
113
CHAPTER 6
6.0
High Speed Imaging Probes for Clinical 3D-OCT
114
6.1 Motivation
Imaging probes are a critical component of 3D-OCT systems, with probe performance directly
impacting overall image quality and clinical utility. Gastrointestinal (GI) imaging with 3D-OCT requires
beam delivery through a probe that is flexible, robust, and small enough to fit through the working
channel of a conventional endoscope. Probe design can be optimized for different imaging requirements
and entails tradeoffs between the scan pattern, scan speed, transverse resolution, working distance, size,
robustness, and cost. Initial animal studies and validation of the 3D-OCT endomicroscopy system
described in Chapter 5 were performed using a modified commercial cardiovascular probe supplied by
LightLab Imaging Inc. This commercial probe was a proximally-actuated radial imaging device that
combined rapid rotary motion (up to 80 Hz) with a linear pullback to image a cylindrical volume.
The commercial probe was helpful during development of the 3D-OCT system and image processing
methods, but its modified cardiovascular design was not robust or stable enough for GI applications. The
diameter of the commercial probe was minimized to allow access to narrow blood vessels, which reduced
mechanical strength and rendered the probe vulnerable to damage during GI endoscopy. Furthermore, the
small diameter torque coil used to transfer rotational motion to the distal tip was somewhat unstable when
the probe was passed through anatomic bends in the colon. Finally, the maximum insertion length of the
probe was insufficient for studying the transverse colon and right colon. For these reasons an improved
spiral-scanning probe that was optimized for GI imaging was developed in this portion of the thesis work.
Two different optical lens systems were developed in parallel to reduce the risk of failure. One lens
system used a series of fused optical fibers to create an integrated lens with minimal diameter and backreflected power. The second lens system used a bulk graded index (GRIN) lens to achieve high
transmission, efficient back-coupling, and good beam quality. Both lens systems were designed to be
compatible with a common torque coil for translation of rotational motion, plastic sheath for electrical and
biological isolation, and proximal fiber connector for attachment to the patient interface unit (PIU)
developed by LightLab Imaging Inc. The bulk micro-optic GRIN design was found to provide better
performance and much higher assembly yields than the fuser fiber design, although the fused fiber design
could be used in the future for applications such as needle-based imaging requiring the smallest possible
probe diameter.
6.2 Fused Fiber Lens Systems
Fused fiber lenses consisting of single mode fiber, non-waveguiding "coreless" fiber, and graded index
multimode fiber have been described previously for OCT imaging probes [1-4]. In a fused fiber probe the
light travels from the proximal end to the distal tip through an arbitrary length (typically several meters)
115
of single mode fiber such as SMF-28. A short section (typically several hundred microns) of coreless
fiber is used to expand the 9.2 jtm beam diameter at the SMF-28 core. The expanded beam then enters a
graded index multimode fiber section that focuses the beam to a desired radius and working distance.
Each fiber section is joined together using fusion splicing to ensure excellent optical contact between the
elements, thereby minimizing backreflected light that can corrupt OCT images. A variety of polishes can
be applied to the lens tip in order to change the direction of the beam or further reduce backreflection.
For this section of the thesis work two different fused fiber lens systems were developed based on an
initial design by LightLab Imaging Inc. One design used standard commercially-available multimode
fiber for the focusing element, while the other used custom multimode fiber with a low index gradient.
The low gradient fiber produces a weaker focusing effect, allowing the working distance to be extended
compared to the standard multimode fiber. In both cases, standard commercially-available coreless fiber
was used for beam expansion. Both lenses were designed using ZEMAX modeling software as described
in Section 6.2.1. A performance comparison between the two designs is provided in Section 0
6.2.1 ZEMAX Modeling - StandardMultimode Fiber
The ZEMAX model of the fused fiber lens system using commercial multimode fiber is shown in
Figure 6.1. The lens is formed by fusing a 145 [tm length of multimode fiber between 493 tm and 230 gim
lengths of coreless fiber. The second length of coreless fiber is used to provide space for end tip polishing
as described in Section 6.2.3. The working distance of the lens was set to 1540 im in water. This working
distance was required in order to propagate the beam outside of the probe body (-1040 jtm radius, filled
with water) and 500 gm into the tissue. The refractive index properties of the multimode fiber was
obtained from the supplier (Newport Inc.) and the coreless fiber was modeled as pure silica with a
diameter of 125 jim. The multimode section had a core diameter of 100 jtm and was used as the field stop
in the ZEMAX model.
Optimization of the design was performed by setting the three fiber lengths as independent variables.
The optimization function was set up to place a beam waist at the output plane, effectively setting the
focus at the final surface in the model. Equal weight was placed on the X and Y components of the beam
for optimization. A Gaussian skew beam with a diameter of 9.2 gm was launched from the first surface of
the model to emulate the mode field of SMF-28 fiber, and Gaussian modeling was used for the
optimization loop. Optimization resulted in the component lengths shown in Figure 6.1(A). Figure 6.1(B)
shows the spot diagram at the output surface, giving an estimated beam radius of 15 jm. This spot size is
- 2x larger than ideally desired for imaging small features such as crypts or Barrett's esophagus glands in
the human GI tract, but would provide reasonable performance for many applications.
116
6.2.2 ZEMAX Modeling - Custom Multimode Fiber
The ZEMAX model of the fused fiber lens system using custom multimode fiber is shown in Figure
6.2. The lens is formed by fusing a 559 jtm length of multimode fiber between 573 [tm and 200 jim
lengths of coreless fiber. The second length of coreless fiber is used to provide space for end tip polishing
as described in Section 6.2.3. As with the previous design, the working distance of the lens was set to
1540 jim in water. The refractive index properties of the multimode fiber was obtained from the supplier
(LightLab Inc.) and the coreless fiber was modeled as pure silica with a diameter of 125 jtm. The
multimode section had a core diameter of 62.5 jtm and was used as the field stop in the ZEMAX model.
Some vignetting of the expanded beam is visible at the first surface of the multimode fiber section in
Figure 6.2. This resulted in a slight decrease in transmission but gave an improved beam radius at the
output surface. Optimization of this design was performed in the same manner as the previous design.
Figure 6.2(B) shows the spot diagram at the output surface, giving an estimated beam radius of 8 Rm.
This spot size and working distance is close to ideal for GI imaging applications.
6.2.3 Polishing
In order to produce a radial scan pattern when the lens is rotated, the beam must exit the lens
perpendicular to the long axis of the fiber. This is accomplished by placing a 400 angle polish on the front
face of the lens tip to tilt the beam by 80' as shown in Figure 6.3. A small air pocket at the tip of the lens
is formed by adding an ultrathin plastic sheath to the lens and filling the end with epoxy. This air pocket
creates total internal reflection when the light strikes the interface between the coreless fiber and the air,
tilting the beam by 80' relative to it's original direction as shown by the red arrow in Figure 6.3. A 900 tilt
would result in the beam striking the probe sheath at normal incidence, giving a high backreflection that
can corrupt OCT images. A second 50 polish is also required on the longitudinal face of the fiber lens as
shown in Figure 6.3. This polish removes the radius of the fiber and prevents strong negative lensing as
the beam exits the coreless fiber and enters the water inside the probe lumen. Polishing was conducted
using a mechanical disc polisher and a purpose-built jig to hold the fiber in place. The fiber was placed in
a standard single mode fiber chuck and the chuck was clamped into the jig. The 500 face angle polish was
performed first, then the entire jig was rotated 450 and the longitudinal polish was applied. Polishing was
done in stages down to a final grit size of 0.3 jim.
6.2.4 FabricationTolerances
One significant drawback of the fused fiber lens designs is that they are highly intolerant to length
errors in the multimode fiber section. Table 6.1 shows ZEMAX model results of tolerance testing for both
the commercial and custom multimode fiber lenses. In these simulations the length of the multimode
117
section was varied by -30 tim to +50 jtm in steps of 10 jtm and the resulting change in working distance
was recorded. With a desired focal depth of 500 jtm in tissue the maximum variation in working distance
that can be tolerated is -400 jtm / +500 jim. At negative variations greater than -400 jtm the focus would
be set inside of the probe's outer sheath, and at positive variations greater than +500 jtm the focus would
be too deep to generate sufficiently high resolution of epithelial tissue layers.
Table 6.1 shows that the commercial multimode fiber reaches these tolerance limits at length
variations of +/- 20 jm. When the length of the commercial multimode fiber was reduced by more than
20 jam the focus shifted to inside of the last coreless fiber section. The custom multimode fiber reaches
the defined tolerance limits at even smaller length variations of < +/- 10 jam. Since each fiber section must
be cleaved and fusion spliced manually, this resulted in a very low assembly yield for both types of fused
fiber lenses. Poor assembly yield was the main reason that the fused fiber probes were not pursued
further, although they do provide exceptionally small diameters and may be useful for certain imaging
applications.
Commercial:
Multimode Length
Change in WD [umrn
Error [um]
N/A
-30
-278
-20
-397
-10
0
0
274
10
473
20
614
30
719
40
799
50
Custom: Change
in WD [um]
2344
3830
2886
0
-578
-799
-920
-996
-1049
Table 6.1. Variation in lens working distance (WD) resulting from length errors in commercial
and custom multimode fiber lengths in fused fiber probes.
6.2.5 MeasuredPerformance
The transmission, back-coupling, and working distance were measured for one of each type of fused
fiber lens representing the best fabricated units. Spot sizes were estimated by entering the measured
working distance back into the ZEMAX models and allowing the fiber lengths to vary to produce a focal
spot at that location. The spot size was then recorded from the model output. This was necessary since
spot size could not be measured directly using diagnostic equipment at hand in the laboratory.
The optimal fused fiber lens incorporating commercial multimode fiber produced a transmission of
71%, back-coupling efficiency of 11%, and working distance of 203 jm giving an estimated spot size
(diameter) of 4.3 jim. The optimal lens incorporating custom multimode fiber produced a transmission of
118
96%, back-coupling efficiency of 75%, and working distance of 508 [tm giving an estimated spot size
(diameter) of 5.4 Rm. Several other fused fiber lenses produced vastly different working distances well
outside the desired range for GI applications. While the small size of the fiber probes is highly desirable,
automated splicing and cleaving equipment would be required to reproducibly construct the devices.
6.3 Micro-Optic Lens Systems
Bulk GRIN lenses are used for a wide variety of applications in photonics and biomedical optics due
to their small size and high optical quality. A bulk micro-optic lens system consisting of a GRIN lens,
fiber pigtailed glass ferrule, metallized angle prism and epoxy spacer was designed as an alternative to the
fused fiber lenses described in Section 6.2. The micro-optic lens was found to provide good performance
and, importantly, much higher assembly yields compared to the fused fiber lenses.
Figure 6.4 shows a schematic diagram of the micro-optic lens system and the complete distal probe
tip. A length of SMF-28 fiber running to the proximal end of the probe is terminated in a glass ferrule.
The beam exiting the fiber expands in an epoxy gap that is transmissive at 1310 nm. This epoxy gap
serves the same function as the coreless fiber in the fused lens design, allowing the input aperture of the
bulk GRIN lens to be filled more effectively. The GRIN lens is a commercially available element with a
numerical aperture (NA) of 0.46 and a pitch of 0.25 at 1310 nm. An angle prism with a metallized
hypotenuse is glued to the GRIN lens to direct the beam out the side of the probe, serving the same
function as the angle polish in the fused lens design. The second surface of the GRIN lens is polished to
50
to ensure that the beam exits the probe 100 off of normal incidence to the sheath. All of the micro-optic
components are glued to a stainless steel hypo tube using rigid epoxy. The lens system is joined to the
flexible torque coil using an outer hypo tube that is epoxied to both the lens tube and the torque coil. A
window is ground in the outer hypo tube to allow light to exit the lens. A transparent plastic sheath
encapsulated the entire probe, providing electrical and biological isolation. Water is flushed down the
probe from the proximal end to improve index matching between the angle prism, sheath, and tissue and
also to provide lubrication for rotational motion.
6.3.1 ZEMAX Modeling
A complete ZEMAX model of the micro-optic probe, including the sheath, water and tissue, was
constructed to optimize the epoxy gap size. The ZEMAX model is shown in Figure 6.5. The model is
constructed as a folded mirror image to allow estimation of the back-coupling efficiency. The sheath is
modeled as two surfaces with one-dimensional curves corresponding to the inner and outer radius of the
sheath. An 8' angle is introduced at the second surface to account for the angle polish on the glass ferrule.
Note that the ray trace used to generate Figure 6.5(A) begins inside the ferrule, whereas the Gaussian
119
beams used for optimization and spot size estimation were launched from the output surface of the
ferrule.
After selection of an appropriate GRIN lens from the catalogue of a commercial supplier, the only free
parameter in this model was the length of the epoxy spacer. This was optimized by constructing a merit
function that placed a beam waist at the desired focal plane in tissue and a second waist at the fiber
"output" of the mirror image probe. The merit function equally weighted both constraints to ensure a
balance between spot size in tissue and back-coupling efficiency. After optimization the spot size
(diameter) in tissue was estimated to be 13 tm in X and 14.5 plm in Y. The spot diagram at the focal plane
in tissue is shown in Figure 6.5(B). Minimal astigmatism is observed due to the relatively low NA of the
beam and relatively large curvature of the sheath. The estimated back-coupling into the fiber was
estimated to be 47% after accounting for all losses and phase front curvature of the beam.
6.3.2 Minimization ofBackreflection
For 3D-OCT imaging applications it is critical to minimize backreflected light from each surface in the
probe. Backreflected light creates significant amounts of image noise by flooding the photodetector with
incoherent photons. Backreflected intensities should generally be several orders of magnitude less than
the expected reference arm power to minimize this effect. With typical power levels of -20 mW
transmitted into the probe, backreflection levels of lower than -50 dB are required. Standard antireflection
coatings provide -10 to -20 dB of reflection suppression and are not suitable for this application.
Backreflection levels were reduced to -50 to -65 dB in the micro-optic lens system by angle polishing the
end face of the glass ferrule and the first surface of the GRIN lens, and by using an epoxy spacer instead
of an air spacer. An angle of +80 was introduced to the glass ferrule to reduce backreflection between the
fiber/epoxy interface. The epoxy used in the spacer was chosen to have a refractive index of 1.524, which
is very close to the refractive index of the fiber core (1.467). The first surface of the GRIN lens was
polished at an angle of -80 in order to direct reflected light away from the fiber in the centre of the ferrule.
The dominant source of remaining backreflection was the second surface of the angle prism, which could
not be easily polished, since the beam exits the prism at close to normal incidence. In the future a custom
angle prism with a slightly tilted output surface could be used to further minimize backreflection if
necessary.
6.3.3 MeasuredPerformance
The micro-optic lens design provided good performance and high assembly yields of better than 75%.
Measurements of transmission, back coupling, working distance and backreflection levels are shown in
Table 6.2. Working distance was measured in air from the output surface of the angle prism, and so does
120
not correspond to the desired working distance through water and tissue. A working distance of 940 Lm in
air is desired for an optimal final working distance. Transmission levels were high (75%) and showed a
very low standard deviation across the ten probes measured (+/- 3%). Back coupling was slightly higher
than predicted by the ZEMAX model (59% vs. 47%), possibly due to the measurement being performed
in air instead of water and tissue. The back coupling deviation was +/- 10%, which is slightly high but
understandable given the sensitivity of back coupling to wavefront aberration. The working distance was
941 pm +/- 60.6 pm compared to an ideal value of 940 ptm. The working distance was precisely set for
each individual probe by placing a mirror 940 pm from the lens tip and adjusting the axial position of the
glass ferrule relative to the GRIN lens to maximize the back-coupled power. Backreflection levels were
very low (-54 dB +/- 4.3 dB). Only one probe had to be discarded due to high backreflection (-44 dB).
One additional probe was broken during connectorization, giving a total assembly yield of 80%. The
combination of good performance and high yield made the bulk micro-optic lens system much more
attractive than the fused fiber lenses for GI applications.
Probe
Number
Transmission
1
2
3
4
5
6
7
8
9
10
MEAN
STD
76%
81%
74%
77%
76%
74%
70%
74%
73%
78%
75%
3.0%
Back
Coupling
70%
74%
54%
66%
55%
64%
65%
44%
46%
55%
59%
10.0%
Working
Backreflection
Distance [um]
[dB]
914
-51
940
-57
-54
965
902
-54
-53
940
890
-44
-59
953
1041
-56
-58
840
1026
-52
-54
941
60.6
4.3
Table 6.2. Measured performance of ten micro-optic lens probes. One probe was rejected
due to high backreflection levels (-44 dB) and one was broken during connectorization. 80%
of the probes were useable.
6.4 Mechanical Design
The mechanical design of the probe is equally important to the optical design, since stable
transmission of the rotational and pullback motion from the proximal PIU to the distal tip is required to
generate a high-quality spiral scan pattern of the imaging beam. In addition the probe must be designed to
be flexible, robust, and small enough to fit down the working channel of a conventional endoscope (2.8
mm diameter). The following sections describe the mechanical design of the torque coil, proximal joint
121
between the torque coil and rigid hypo tube, and fiber connector / water flush port for the high-speed
imaging probe.
6.4.1 Torque Coil Selection
A flexible torque coil comprised of multiple layers of counter-wound wire is used to transmit
rotational and pullback motion from the PIU to the distal tip of the probe. Torque coil designs inherently
trade off stiffness against torque transmission; a very stiff coil accurately transmits rotation and
translation, whereas a flexible coil is more compressible and therefore transmits motion less accurately.
Two torque coil designs with different stiffness characteristics were tested for GI endoscopy applications.
The more rigid design was found to translate push/pull and rotary motion very effectively, but the lack of
flexibility caused binding when the endoscope was actuated. The more flexible design had much less
severe binding problems and still produced highly stable rotational motion at up to 60 Hz. Push/pull
translation was degraded compared to the rigid torque coil but overall performance was significantly
better. The final torque coil used in the probe was a 3-layer design with an outer diameter of 1.45 mm and
an inner diameter of 0.43 mm. The outer diameter of SMF-28 fiber with a standard coating is 0.25 mm so
the fiber could be readily passed down the torque coil. The overall length of the torque coil was 2 m or 3
m, making the probes compatible with conventional GI endoscopes.
6.4.2 ProximalJoint
A potential failure point is present at every location where two probe components are joined together.
It is therefore important to design each joint for high strength while still allowing for high-speed
translational and rotational motion. As described in Section 6.3, the joint between the micro-optic lens
and torque coil is covered in a rigid hypo tube to provide strain relief and prevent separation of the lens
from the coil. A similar method is employed at the proximal end of the probe. Here the torque coil joins a
long, rigid hypo tube that is epoxied to a standard FC/APC fiber connector on the other end. The
connector and hypo tube allow the probe to be attached to the PIU. The joint between the hypo tube and
torque coil is shown in Figure 6.6. A protective hypo tube is epoxied over top of the torque coil and outer
hypo tube that is attached to the FC/APC connector. A freely floating inner hypo tube is positioned over
top of the SMF-28 fiber to prevent epoxy from adhering to the fiber and causing stress when the probe is
flexed. A layer of heat shrink is attached to the outer hypo tube to improve centration within the
protective hypo tube. This joint structure provides strain relief and high strength while still permitting the
fiber to flex within the torque coil.
122
6.4.3 Fiber Connector and Flush Port
An FC/APC connector and water flush port are attached to the proximal probe section as shown in
Figure 6.7. A thin piece of heat shrink is placed on the outer hypo tube to ensure centration inside the
FC/APC connector. Centration is critical to prevent off-axis motion of the probe when the PIU motor is
rotating. A plastic adapter to secure the probe in the PIU is fastened over a metal sleeve, and a standard
Tuohy-Borst flush port is used to allow water flushing of the probe during imaging. The plastic sheath
containing the torque coil and fiber is joined to the flush port to form a tight seal.
6.4.4 Future ProbeDesign
While initial imaging studies using the micro-optic probe were encouraging, further optimization of
the design is possible. The overall outer diameter of the probe was 2.34 mm - 2.54 mm depending on the
type of plastic sheath used, which is only slightly smaller than the nominal endoscope working channel
inner diameter of 2.8 mm. Some probe binding and unstable push/pull actuation was therefore observed
during some procedures when a tight radius bend was introduced into the endoscope. This binding was
likely caused by deformation of the working channel cross-section into an ovular shape, which put
pressure on the probe sheath and prevented uniform rotation and pullback. This problem will be addressed
in a revised version of the probe by reducing the overall outer diameter to 1.80 mm (30% - 40%
reduction).
The revised probe is shown in Figure 6.8. Further miniaturization of the probe will be possible by
using a thinner torque coil with an outer diameter of 1.00 mm, allowing the hypo tube holding the lens
components to be extended back and used for strain relief. This allows removal of the second hypo tube
previously used for strain relief between the torque coil and lens, and significantly reduces the overall
diameter. A smaller angle prism is required to ensure that it will fit into the hypo tube. Additionally, the
entire lens assembly can be sealed with IR transmissive epoxy to reduce the buildup of debris on the angle
prism surface. The small change in the torque coil thickness is not expected to significantly impact the
rotational stability of the probe, and push/pull stability should be improved by reducing pressure from the
endoscope.
6.5 Imaging Performance Comparison
The bulk micro-optic probe was tested in several normal human subjects during screening
colonoscopy and the resulting 3D-OCT data was compared to earlier images acquired with a modified
commercial fused fiber probe originally designed for cardiovascular applications. The micro-optic probe
was found to provide significantly reduced rotational jitter due to the more robust torque coil. The
pullback uniformity of the two probes was similar. The micro-optic probe also produced a higher quality
123
beam, tighter focus, higher back-coupling efficiency than the cardiovascular probe, resulting in improved
signal to noise ratios and reduced transverse speckle noise. These features translated into significantly
improved enface and cross-sectional 3D-OCT image quality in human subjects.
A side-by-side comparison of enface image quality for data acquired using the bulk micro-optic probe
and modified cardiovascular probe is shown in Figure 6.9. Both images were acquired in the sigmoid
colon of normal patients and show transverse features consistent with colonic crypts. The significantly
improved rotational stability of the micro-optic probe results in enhanced visualization of crypts. Both
probes show some non-uniformity in pullback motion that cause crypts to appear elongated.
A side-by-side comparison of radial image quality for data acquired using the bulk micro-optic and
modified cardiovascular probe is shown in Figure 6.10. Both images were acquired in the human finger
pad, although the image in Figure 6.10(A) also shows an infrared viewing card in the bottom-left
quadrant. Both images were acquired at the same axial line rate and probe rotational speed, ensuring that
the number of lines per image is the same for both cases. The better optical performance of the microoptic probe results in significantly improved transverse resolution, higher signal to noise ratio, and
reduced speckle size compared to the cardiovascular probe. Overall, the micro-optic probe provides
excellent optical performance and can be reliably used for GI endoscopy procedures.
124
6.6 Figures
1540 um
145 um 230 um
I
-
'1
,,
30 LAYOUT
FUSED FIBER LENS WITH COMMERCIAL MULTIMODE SECTION
THU TAN 15 2009
in
IMagluI1*Auruml
CONFIGURATION
50
lu-
1 OF
um
1
Figure 6.1. ZEMAX model of a fused fiber lens system using commercial multimode fiber as the focusing
element. A: Schematic diagram showing optimized layout. Order of elements is coreless fiber, multimode
fiber, coreless fiber, and water. B: Spot diagram at focal plane. Estimated beam radius is 15 um at a
working distance of 1540 um.
573 um
559 um
FUSED FIBER LENS WITH
THU TRN 15 2009
1540 um
200 um
3D LRYOUT
CUSTOM MULTIMODE SECTION
40 um
Figure 6.2. ZEMAX model of a fused fiber lens system using custom multimode fiber as the focusing
element. A: Schematic diagram showing optimized layout. Order of elements is coreless fiber, multimode
fiber, coreless fiber, and water. B: Spot diagram at focal plane. Estimated beam radius is 8 um at a
working distance of 1540 um.
CONFIGURATION
1 OF
1
To Proximal End
Figure 6.3. Schematic diagram of fused fiber lens showing polish orientations on distal tip. An air gap
formed by an ultrathin sheath and epoxy bead creates total internal reflection, directing the beam out of
the lens at an 800 angle. C, coreless fiber. MM, multimode fiber. E, epoxy bead.
125
02.34
01.93
01.83
-01.27
GRIN Lens
Epoxy
IR Transmissive
Epoxy
Glass Ferrule
Angle Prism
Torque Coil
Glass Fele
Hypo Tube
Plastic Sheath
18.00
Hypo Tube SMF-28 Fiber
Water
Desired Focal
Plane
.
.
..
Figure 6.4. Schematic diagram of bulk micro-optic lens system and distal tip of complete probe. All
dimensions are in mm. Focal plane is set at 500 um from the sheath surface. Inset shows an end view of
the probe.
SEpoxy
GRIN Lens
Tissue
Prism Water Sheath
43D
Mirror Image
-
LAYOUT
BULK MICRO-OPTIC LENS SYSTEM RNO PROXIMRL PROBE TIP
THU TRN 15 2009
30 um
CONFIGURATION
1 OF
1
Figure 6.5. ZEMAX model of a bulk micro-optic lens system. A: Schematic diagram showing optimized
layout. Model is constructed as a mirror image to allow estimation of fiber back-coupling efficency. B: Spot
diagram at focal plane. Estimated beam radius is 13 pm x 14.5 pm (XY) at a working distance of 500 pm
in tissue.
Epoxy
Heat Shrink -
Outer Hypo
Tube
Inner Hypo Tube
Protective Hypo
Tube
SMF-28 Fiber
Torque Coil
Figure 6.6. Schematic diagram of the proximal joint between the torque coil and rigid hypo tube for the
bulk micro-optic probe. A freely floating inner hypo tube is used to prevent epoxy from stressing the fiber.
126
Figure 6.7. Schematic diagram of the FC/APC connector and water flush port for the bulk micro-optic
probe.
0
01.80
01.50
-01.27
01.07
Angle Prism
01.00
Torque Coil
Epoxy
IR Transmissive
Epoxy
Glass Ferrule
Plastic Sheath
17.00
Hypo Tube
GRIN
Lens
Water
..
...
km..
Desired Focal
Plane
hypo tube is
Figure 6.8. Schematic diagram of the future reduced-diameter micro-optic probe. A single
torque coil.
used to hold the lens elements and provide strain relief for the joint between the lens and the
probe.
the
of
view
Overall outer diameter can be reduced to 1.80 mm. Inset shows an end face
127
Figure 6.9. Comparison of en face image quality for normal human colon acquired with a bulk micro-optic
probe a modified cardiovascular probe. A: En face image obtained using a bulk micro-optic probe. B: En
face image obtained using a modified cardiovascular probe. The micro-optic probe provides improved
rotational stability, resulting in improved visualization of transverse tissue features such as colonic crypts.
Nonuniform pullback motion in both probes causes occasional image artifacts (arrows).
Figure 6.10. Comparison of radial image quality for human finger pad acquired with a bulk micro-optic
probe a modified cardiovascular probe. A: Image obtained using a bulk micro-optic probe. B: Image
obtained using a modified cardiovascular probe. The micro-optic probe provides improved beam quality,
spot size, and back-coupling efficiency, resulting in improved contrast and resolution.
128
6.7 References
[1]
W. A. Reed, M. F. Yan, and M. J. Schnitzer, "Gradient-index fiber-optic microprobes for
minimally invasive in vivo low-coherence interferometry," Optics Letters, vol. 27, pp. 17941796, Oct 15 2002.
[2]
M. S. Jafri, S. Farhang, R. S. Tang, N. Desai, P. S. Fishman, R. G. Rohwer, C. M. Tang, and J.
M. Schmitt, "Optical coherence tomography in the diagnosis and treatment of neurological
disorders," Journalof Biomedical Optics, vol. 10, pp. -, Sep-Oct 2005.
[3]
H. Li, B. A. Standish, A. Mariampillai, N. R. Munce, Y. X. Mao, S. Chiu, N. E. Alarcon, B. C.
Wilson, A. Vitkin, and V. X. D. Yang, "Feasibility of interstitial Doppler optical coherence
tomography for in vivo detection of microvascular changes during photodynamic therapy,"
Lasers in Surgery and Medicine, vol. 38, pp. 754-761, Sep 2006.
[4]
Y. X. Mao, S. Chang, S. Sherif, and C. Flueraru, "Graded-index fiber lens proposed for ultrasmall
probes used in biomedical imaging," Applied Optics, vol. 46, pp. 5887-5894, Aug 10 2007.
129
CHAPTER 7
7.0
Clinical 3D-OCT in the Upper Gastrointestinal Tract
130
7.1 Motivation
Endoscopic therapies are becoming increasingly common in gastroenterology due to their less invasive
nature, lower morbidity, and faster recovery time compared to most surgical interventions. Endoscopic
therapies, however, are not indicated for every patient and multiple treatments may be required to obtain
adequate control of a disease. Gastroenterologists can therefore benefit from imaging tools that can aid in
pre-interventional planning and follow-up assessment of endoscopic therapies. Follow-up assessment can
include both determining the completeness of treatment and checking for recurrence of disease. 3D-OCT
is desirable for these applications since it is compatible with existing commercial endoscopes and enables
comprehensive depth-resolved tissue imaging. Advanced image processing techniques enable improved
visualization of tissue microstructure for assessing features beneath the superficial layers typically
apparent under endoscopy. These techniques include the generation of cross-sectional images with
arbitrary orientations, generation of projection views similar to en face microscopy images, improved
quantitative measurements of morphology, 3D image processing methods for speckle noise reduction, and
virtual manipulation of tissue geometry for visualizing structural morphology.
Barrett's esophagus (BE) is a pathology of the upper gastrointestinal (GI) tract that is associated with
an approximately 40-fold increases in risk of progression to dysplasia and adenocarcinoma compared to
the general population. Recently BE has become amenable to new endoscopic therapies to treat the
disease prior to progression to invasive cancer. BE results from chronic mucosal injury and is a precursor
condition to esophageal adenocarcinoma [1]. Esophageal cancer has a five-year survival rate of only 16%
[2], but early detection and treatment achieves a high percentage of regression in patients with dysplasia
[3]. BE is characterized by the replacement of squamous epithelium with specialized intestinal columnar
epithelium [4]. Neoplastic changes in BE develop in stages from non-dysplastic metaplasia to increasing
grades of dysplasia and eventually to adenocarcinoma [5].
Differentiation of normal mucosa, BE, dysplasia, and carcinoma are necessary for an imaging method
to be used to assist pre- or post-therapy assessment. Despite initial enthusiasm over chromoendoscopy
(using chemical dyes to increase contrast), narrow band imaging, autofluorescence, and other imageenhancing methods, larger studies have shown poor discrimination between BE and early-stage cancers.
A technology such as 3D-OCT that can be used for guiding excisional biopsy, providing subsurface tissue
imaging, planning endoscopic therapies, and assessing patients at follow-up would significantly improve
overall outcomes. Currently, random four-quadrant biopsies taken every 1-2 cm along the region of BE
(Seattle protocol) is the clinical standard for detecting dysplasia and adenocarcinoma during BE
surveillance [6, 7]. This procedure suffers from high false negative rates due to sampling errors, since
high-grade dysplasia and early invasive cancer are typically not visible or distinguishable from
131
surrounding BE by standard endoscopy. Moreover, it is known from surgically resected specimens that
both dysplastic and cancerous lesions can be multiple, scattered, and small, thereby making detection of
these lesions especially prone to sampling errors. The prevalence of undetected adenocarcinoma from the
random biopsy protocol ranges from 33-57% [8-11].
Radiofrequency ablation (RFA) has recently been introduced to treat diffuse conditions such as BE
[12]. RFA produces a broad, superficial ablation pattern compared to the deep, focal ablation obtained
with argon plasma coagulation, electrocautery, or heater probes. Endoscopic 2D-OCT analysis has shown
that some patients can experience recurrence of BE following RFA treatment [13] [14, 15], which could
be associated with a future risk of adenocarcinoma under the neosquamous epithelium [16, 17].
Endoscopic mucosal resection (EMR) is a common method for treating visible dysplastic nodules arising
in a setting of BE. In EMR, fluid is typically injected into the submucosa to raise the nodule. The nodule
is then removed with an electrocautery snare with or without the aid of a resection cap or lift-banding.
EMR is also associated with some disease recurrence due to incomplete extraction of dysplastic lesions.
This motivates the need for 3D-OCT endomicroscopy over larger fields of view with complete
visualization of 3D tissue microstructure in order to assess endoscopic therapies in the planning stage and
at follow-up.
In this section of the thesis work, a 3D-OCT endomicroscopy system and spiral-scanning imaging
catheter were used to assess five tissue categories: 1) no pathology ("Normal"); 2) BE prior to undergoing
radiofrequency ablation therapy ("BE pre-RFA"); 3) sites previously treated with radiofrequency ablation
therapy ("BE post-RFA"); 4) visible esophageal nodules prior to undergoing endoscopic mucosal
resection ("Nodule pre-EMR"); and 5) sites previously treated with endoscopic mucosal resection
("Nodule post-EMR"). 3D-OCT was used to establish characteristic features of esophageal mucosa for
each tissue group. 3D-OCT was found to be a valuable adjunct for endoscopic therapies such as
radiofrequency ablation (RFA) and endoscopic mucosal resection (EMR) for treatment of BE and
esophageal nodules, enabling tissue characterization over surface areas many times larger than the typical
size of excisional biopsy specimens.
7.2 Clinical Protocol
All subjects in this study were imaged at the Veterans Affairs Boston Healthcare System (VABHS)
Jamaica Plain campus in collaboration with Dr. Hiroshi Mashimo, MD PhD, and Dr. Qin Huang, MD.
Subjects were recruited from the pool of patients undergoing elective endoscopy for screening,
surveillance, or treatment of BE or nodules. Subjects followed a standard preparation procedure for upper
endoscopy, including cessation of blood thinners and fasting prior to the appointment. 3D-OCT imaging
was performed in tandem with standard video endoscopy.
132
For normal subjects, the endoscopist identified regions for 3D-OCT imaging that appeared normal and
unremarkable using white light video endoscopy. Imaging was performed by inserting the 3D-OCT
catheter down the working channel of a standard endoscope. For pre-treatment subjects, the endoscopist
identified locations suspicious for BE or esophageal nodules using changes in tissue surface texture and
colour visible under white light video endoscopy and narrow band imaging (NBI). For post-treatment
subjects, the endoscopist identified regions of prior treatment using patient history notes and, in some
cases, subtle visual cues that differentiate treated tissue from untreated tissue. Where clinically indicated,
pinch biopsy samples were obtained from the 3D-OCT imaging sites following the conclusion of all
image acquisition.
Between 1-6 sites were assessed for each subject, increasing the total procedure time by an average of
10.5 minutes and by no longer than 19 minutes. This increase in procedure length is similar to the time
increase associated with chromoendoscopy [18]. After the acquisition of all 3D-OCT datasets, the
imaging catheter was withdrawn from the working channel and replaced with pinch biopsy forceps to
acquire biopsies of the imaged tissue where clinically indicated. Biopsies were generally not obtained
from normal subjects due to the slight risk associated with pinch biopsy and the lack of clinical necessity.
Histopathology from pinch biopsies was evaluated by an expert blinded pathologist in order to establish a
tissue classification for each site. The histological diagnoses were then compared to tissue features based
on the 3D-OCT data.
In all cases, the 3D-OCT endomicroscopy system described in Chapters 5 - 7 was used to obtain
volumetric image data. While imaging parameters varied slightly due to system upgrades (such as FDML
sweep linearization) and continuing probe development, typical performance is as described in Chapter 5.
To briefly reiterate, sensitivity was 105 - 107 dB with 13 - 15 mW of power incident on the tissue.
Optical resolution in tissue was 4.7 - 5.1 Rim (axial) and 10 - 15 [tm (transverse). Maximum imaging
ranges were 1600 - 1800 gpm from the centre of the probe, although the useable attenuation-limited
imaging depth was typically 1000 - 1200 [tm from the tissue surface. True 3D resolution varied with
probe rotation speed and pullback speed, but the system was typically set to provide axial line spacing of
7 - 12 [tm and frame-to-frame spacing of 8 - 12 pm. Axial pixel spacing was typically 3.0 - 3.7 tm, and
varied depending on the total imaging range.
In total, 61 sites from 18 unique patients were imaged during this study from February to December
2008. Table 7.1 shows a summary of the collected 3D-OCT image data. Note that the total number of
patients in Table 7.1 exceeds 18 since multiple tissue types were imaged from several patients, typically
consisting of a normal region and a pathological region. The "other" grouping includes pathologic tissue
that was outside the main scope of the study, including ulcers, strictures, gastric anastomosis, and
133
eosinophilic esophagitis. The following sections show representative 3D-OCT data, characteristic
features, and comparisons to histology for each tissue type.
Tissue Type
Normal Esophagus
BE Pre-RFA
BE Post-RFA
Nodule Pre-EMR
Nodule Post-EMR
Other Pathology
Number of
Patients
7
5
7
3
2
7
Number of
Imaging Sites
11
11
12
8
5
14
Table 7.1. Summary of 3D-OCT data collected during upper endoscopy.
7.3 Normal Esophagus
Obtaining normal image data was a necessary first step in this study in order to establish a baseline for
comparison to pathologic tissue. 3D-OCT images of normal esophageal mucosa were obtained at 11 sites
from 7 unique patients. All patients tolerated the procedure well and there were no known immediate or
long-term complications. Useable data was obtained for 5 out of the 7 patients. 2 patients produced
unusable data due to severe motion artifacts from breathing, heartbeat, gagging, or whole-body motion
during 3D-OCT imaging. Results of the normal esophagus imaging are described in Sections 7.3.1 to
7.3.2 below.
7.3.1 CharacteristicFeatures
Figure 7.1 shows typical 3D-OCT data obtained from the normal esophagus. Figure 7.1(A) shows a
volumetric rendering of the entire 980 x 512 x 1268 (lines x axial pixels x frames, X x Y x Z) dataset,
which was acquired in -21 s. The pullback length was 20.7 mm, the rotational circumference was 7.5
mm, and the imaging range was 1.7 mm in tissue. The images in Figure 7.1 have been axially cropped to
remove the probe sheath and non-useable points beyond the penetration depth of light in the tissue. The
useable imaging depth in tissue was approximately 1100 jim. Unless otherwise specified, all crosssectional and en face images shown below are formed by averaging the volumetric data over a 20 !imthick section to reduce speckle noise. Since 20 lim is approximately the size of a single epithelial cell,
minimal image blurring is observed from this process [19]. The red, green, and blue boxes in Figure
7.1(A) illustrate the locations of orthogonal cut planes used to extract en face and cross-sectional image
slices for closer examination. In this and all subsequent 3D-OCT image data the X axis represents the
probe's rotational axis (fast, 50 - 70 Hz), the Y axis represents the axial OCT line data, and the Z axis
represents the probe's pullback direction (slow, 500 - 1000 jim/s).
134
Figure 7.1(B) shows an en face image centered at a tissue depth of 175 tm, corresponding to the
centre of the squamous epithelial layer. This image is similar in concept to a conventional microscopy
image obtained by setting the focus to a depth of 175 jtm. Consistent with previous 2D-OCT studies [20]
the normal squamous epithelium is featureless and unremarkable. The small, scattered hypointense
regions in this image are artifacts caused by debris on the probe surface. Slight transverse jitter oriented
primarily along the X axis is due to patient heartbeat and respiration artifacts. The blue and green arrow
pairs on the border of Figure 7.1(B) indicate the positions of the cross-sectional images shown in Figure
7. 1(C,D).
One of the main benefits of 3D-OCT endomicroscopy compared to 2D-OCT is the ability to align
cross-sectional images to enface tissue features identified in XZ images. Figure 7.1(C) shows a YZ crosssectional image at the position indicated by the green arrow pair in Figure 7. I1(B). A regular, well-defined
layered architecture can be appreciated throughout the length of the pullback. A slight discontinuity is
present at -14 mm due to sudden patient motion. The red arrow pair indicates the position of the en face
image in Figure 7.1(B). Figure 7.1(D) shows an XY cross-sectional image at the position indicated by the
blue arrow pair in Figure 7.1(B). The regular layered architecture is visible in XY as well. In both crosssections, the epithelium is devoid of glands or vessels. The lamina propria and submucosa tissue layers
(first and second hyperintense layers, see Figure 7.3) contain small (< 100 tm) regular blood vessels,
identifiable by their axial shadows and interconnectivity across numerous cross-sectional images. The
submucosa also contains scattered, non-shadowing glandular structures that may represent normal
seromucous glands. The ability to definitively distinguish vascular from glandular features without the
use of Doppler processing is unique to 3D-OCT. This ability is valuable in the assessment of pathologic
tissue, as discussed below.
The dense spatial sampling density, high optical resolution, and large field of view of 3D-OCT
endomicroscopy enables high-magnification analysis of each tissue layer over arbitrary regions of interest
(ROI). This principle is demonstrated in Figure 7.2. Figure 7.2(A-D) shows en face images at the ROI
shown as a black dashed box in Figure 7.1(B). Images are shown at tissue depths of 175, 340, 420, and
520 p~m for Figure 7.2(A-D) respectively. These depths correspond to the epithelium, lamina propria,
muscularis mucosa, and submucosa respectively. Regular vascular networks are visible in both the lamina
propria and submucosal layers. Individual vessels can be identified in the en face images (red and black
triangles in Figure 7.2(B,D)) as well as the cross-sectional images shown in Figure 7.2(E-G). Nonvascular glands are small (< 100 tm), regular, and confined to the muscularis mucosa and submucosa.
135
7.3.2 Comparison to Histology
Comparison of 3D-OCT endomicroscopy images to histology slides obtained from excisional biopsy
and to white light video endoscopy examination is a good method for linking image features to known
tissue structures. Figure 7.3(A) shows an endoscopic video capture obtained prior to 3D-OCT imaging of
the same patient presented in Section 7.3.1. The esophageal mucosa appears pale-pink, smooth, and
regular, consistent with normal squamous tissue. Figure 7.3(B) shows representative histology of normal
esophageal squamous mucosa. Since there was no clinical indication to biopsy this particular patient, the
histology image in Figure 7.3(B) is taken from another subject where biopsy was required in the midesophagus but a normal diagnosis was made by the pathologist. Excellent correlation is shown between
the histology image and the 3D-OCT image. The epithelium, lamina propria, muscularis mucosa,
submucosa, and muscularis propria are visible as well-separated and distinct layers with alternating hypoand hyperintensity under 3D-OCT (Figure 8.3(C)). The pinch biopsy sample did not extend deeper than
the lamina propria, highlighting another advantage of 3D-OCT compared to conventional histology.
Small, regular vessels that produce axial shadows are visible (V) as well as non-shadowing submucosal
glands (G) under 3D-OCT.
In total, 5 out of the 5 normal patients (100%) with useable 3D-OCT data showed features consistent
with those described above. A normal diagnosis was histologically confirmed in 2 out of 2 patients
(100%) where biopsy specimens were available. The remaining 3 patients did not have biopsy samples
available due to lack of clinical indication. Moderate variability was observed in the apparent tissue layer
thicknesses from patient to patient, and also at multiple sites within the same patient. In general the
variability was uniform across each of the mucosal layers, suggesting that this is due to varying tissue
compression from endoscope suction and changes in probe contact pressure with the luminal surface.
Similar issues have also been observed during 2D-OCT endoscopy and confocal endoscopic microscopy.
7.4 Barrett'sEsophagus: Pre-Radiofrequency Ablation
3D-OCT images of BE were obtained at 11 sites from 5 different patients. This data was collected in
order to establish the characteristics of active BE and allows a comparison between untreated and RFAtreated BE. 3 of these patients had not previously been treated with RFA, EMR, or other endoscopic or
surgical therapies for BE. 1 patient was previously treated with EMR 3 months prior to 3D-OCT imaging
to remove a dysplastic nodule present within a BE region. 1 patient was treated with RFA 4 months prior
to 3D-OCT imaging to ablate a region of BE, but presented on the imaging day with residual BE. All 5
patients had been medically treated with proton pump inhibitors or other pharmaceuticals to control BE
progression. All patients tolerated the procedure well and there were no known immediate or long-term
136
complications. Useable data was obtained for all patients. Results of the BE pre-RFA imaging are
described in Sections 7.4.1 to 7.4.2 below.
7.4.1 CharacteristicFeatures
Figure 7.4 shows typical 3D-OCT data obtained from a BE region prior to any endoscopic or surgical
therapy. This patient presented with small BE islands in the distal esophagus near the gastro-esophageal
junction (GE junction). BE islands represent a clinical situation where 3D-OCT imaging could be
particularly beneficial, since small isolated pockets of BE surrounded by normal squamous mucosa may
be more difficult to discern than the more typical continuous BE "fingers" extending proximally from the
GE junction. BE may also present as short segment, ultra-short segment, or invisible lesions at the GE
junction. These cases may also represent situations where 3D-OCT can provide enhanced visualization
over endoscopic examination alone. Figure 7.4(A) shows a volumetric rendering of the entire 980 x 512 x
1268 (lines x axial pixels x frames, X x Y x Z) dataset, which was acquired in -21 s. The pullback length
was 20.7 mm, the rotational circumference was 7.5 mm, and the imaging range was 1.7 mm in tissue. The
red, orange, green, blue, and purple boxes in Figure 7.4(A) illustrate the locations of orthogonal cut
planes used to extract enface and cross-sectional image slices.
Figure 7.4(B) shows an en face image centered at a tissue depth of 175 pm, corresponding to the
centre of the squamous epithelial layer. The probe was pulled back over the GE junction during
acquisition of this dataset and a clear delineation can be appreciated between gastric mucosa on the lefthand side of the image (distal) and esophageal mucosa on the right (proximal). The gastric mucosa is
covered with a regular pit pattern whereas the esophageal tissue appears relatively featureless. Figure
7.4(C) shows a second en face image at a tissue depth of 345 jpm. Several discrete regions of the
esophageal tissue reveal striking atypical glandular structures consistent with BE. Glands are ovular or
round in cross-section and range in length from 200 - 600 [pm. The BE islands are surrounded by
comparatively normal-appearing squamous mucosa, illustrating the difficulty in analyzing these regions
with 2D-OCT or random excisional biopsy. The blue, green, and purple arrow pairs on the border of
Figure 7.4(B,C) indicate the positions of the cross-sectional images shown in Figure 7.4(D-F).
Figure 7.4(D-F) shows cross-sectional images at multiple positions within the dataset. In Figure 7.4(D)
normal gastric mucosa can be identified by regular vertical pit patterns and low penetration depths,
consistent with previous 2D-OCT studies [20]. A -3 mm length of normal squamous mucosa with regular
layered architecture is adjacent to a -5 mm length of BE tissue. In cross-section, the BE island exhibits
distortion of the normal layered architecture and the presence of discrete hypointense glands. "Distortion"
refers to localized, atypical variations in mucosal layer thicknesses as well as changes in layer-to-layer
contrast compared with normal esophageal tissue. An orthogonal cross-section through a region
137
containing transitionary gastric cardia tissue and normal gastric mucosa is shown in Figure 7.4(E). This
area appears to be free of BE. In comparison, a cross-section at a more proximal location as shown in
Figure 7.4(F) is clearly afflicted by BE. These images illustrate that the BE glands are located beneath
300 - 500 pm of overlying tissue. A large number of glands are present along with distorted layered
architecture.
Figure 7.5 shows an enlarged view of the ROI illustrated in Figure 7.4(B,C). This ROI spans a portion
of one BE island as well as an adjacent normal area. There is clear demarcation between the normal
region (left) and BE island (right). Cross-sectional images in the ROI highlight the difference in layered
appearance between the normal and BE tissue regions. The presence of glands and layer distortion are
hallmarks of BE. Previous 2D-OCT studies have not conclusively established criteria for differentiating
BE with low-grade dysplasia (LGD) from BE without LGD, although the gland density and degree of
layer disruption suggested a more severe form of BE based on the 3D-OCT data. The intact submucosa
indicates that HGD or adenocarcinoma is not present.
7.4.2 Comparison to Histology
Figure 7.6 shows a comparison of 3D-OCT data to conventional white light video endoscopy and
histology analysis of the BE island region. Figure 7.6(A) shows an endoscopic video capture obtained
prior to 3D-OCT imaging of the same patient presented in Section 7.4.1. The esophageal mucosa contains
discrete regions that appear red, inflamed, and irregular, consistent with BE islands. Figure 7.6(B) shows
histology of BE islands taken near the GE junction of the same patient following 3D-OCT imaging.
Excellent correlation is shown between the histology image and the 3D-OCT image. Numerous large BE
glands buried beneath a layer of squamous tissue are visible in both the histology and 3D-OCT images
with distortion of the regular layered architecture. This patient was diagnosed with BE without dysplasia
after histology analysis, compared to dysplasia with LGD from the 3D-OCT data. This discrepancy is not
surprising given the variations in both BE gland density and layered architecture visible in Figure 7.4(C).
Gland density varies significantly over the 3D-OCT image volume and could easily be classified as BE
without dysplasia if a single, less glandular region were used for analysis. This highlights an advantage of
3D-OCT for characterizing BE. Unlike 2D-OCT or excisional biopsy, 3D-OCT endomicroscopy enables
a large surface area of 155 mm 2 to be comprehensively assessed for pathology, potentially reducing
diagnostic sampling errors.
In total, 4 of the 5 BE pre-RFA patients (90%) with useable 3D-OCT data showed features consistent
with those described above. 1 of the 5 patients showed distortion of regular layered architecture without
the development of clear glandular structure. A significant degree of variability was observed in gland
size, gland packing density, and mucosal layer distortion from patient to patient. This was also observed
138
in previous larger-scale studies of BE with 2D-OCT [20] and is indicative of the heterogeneity of the BE
pathology. Using the 3D-OCT data, 2 patients showed evidence of BE with LGD (both diagnosed as BE
without dysplasia under histology) and 2 showed evidence of BE without dysplasia. For both tissue sites
classified as BE without dysplasia based on 3D-OCT data, no histology was available due to RFA
treatment and increased risk of bleeding on the day of imaging.
7.5 Barrett's Esophagus: Post-Radiofrequency Ablation
3D-OCT images of regions previously containing BE that had been treated with RFA were obtained at
12 sites from 7 unique patients. All 7 patients had received RFA at least once and at most five times prior
to 3D-OCT imaging. All patients tolerated the procedure well and there were no known immediate or
long-term complications. Useable data was obtained for 6 out of the 7 patients. 1 patient produced
unusable data due to severe motion artifacts from breathing, heartbeat, gagging, or whole-body motion
during 3D-OCT imaging. Results of the BE post-RFA imaging are described in Sections 7.5.1 to 7.5.2
below.
7.5.1 CharacteristicFeatures
Figure 7.7 shows typical 3D-OCT data obtained from a region previously afflicted with BE that had
been treated with RFA. This particular patient received his last RFA treatment 6 months prior to 3D-OCT
imaging. Under white light video endoscopy the esophageal mucosa appeared regular and unremarkable,
indicating that the RFA treatment was successful and that the BE had resolved. Assessment of previously
treated BE regions represent another clinical situation where 3D-OCT imaging could be beneficial, since
small regions of residual or recurrent BE surrounded by normal squamous mucosa may be extremely
challenging to detect. Figure 7.7(A) shows a volumetric rendering of the entire 736 x 512 x 1613 (lines x
axial pixels x frames, X x Y x Z) dataset, which was acquired in -21 s. The pullback length was 20.7 mm,
the rotational circumference was 8.6 mm, and the imaging range was 1.9 mm in tissue. The red, orange,
green, and blue boxes in Figure 7.7(A) illustrate the locations of orthogonal cut planes used to extract en
face and cross-sectional image slices for closer examination.
Figure 7.7(B) shows an en face image centered at a tissue depth of 105 tm, corresponding to the
centre of the squamous epithelial layer. The probe was pulled back over the GE junction during
acquisition of this dataset and a clear delineation can be appreciated between gastric mucosa on the lefthand side of the image and esophageal mucosa on the right. The distinct linear border of the GE junction
as visualized in Figure 7.7(B) may be a result of the previous RFA treatment, which uses a rectangular
electrode to produce superficial tissue ablation. The gastric mucosa is, as expected, covered with a regular
pit pattern whereas the squamous esophageal tissue appears relatively featureless. Figure 7.7(C) shows a
139
second en face image at a tissue depth of 390 jim. While the tissue appears largely normal, there are a
small number of scattered glandular structures consistent with BE as indicated in Figure 7.7(C). Some
hypointense features that appear to be glandular from this single en face plane are actually regular
submucosal vessels, which was verified by vertically scanning the en face plane and observing
interconnection of the features. The BE glands are isolated and surrounded by normal squamous mucosa,
illustrating the difficulty in analyzing these regions with 2D-OCT or random excisional biopsy. The blue
and green arrow pairs on the border of Figure 7.7(B,C) indicate the positions of the cross-sectional images
shown in Figure 7.7(D,E).
Figure 7.7(D,E) shows cross-sectional images at multiple positions within the dataset. In Figure
7.7(D), normal gastric mucosa and relatively normal squamous mucosa can be readily distinguished. An
orthogonal cross-section through a region of normal gastric mucosa and isolated BE glands is shown in
Figure 7.7(E). In cross-section, the isolated BE glands appear to be buried beneath regular epithelial and
lamina propria layers. These glands can be differentiated from normal seromucous glands by their larger
size, ovular cross-section, and position beneath the lamina propria.
Figure 7.8 shows an enlarged view of the ROI illustrated in Figure 7.7(B,C). This ROI covers a region
containing what appears to be a residual BE gland. Cross-sectional images in the ROI suggest that normal
epithelium (first hypointense layer) has regrown over top of the BE gland. The lamina propria (first
hyperintense layer) appears slightly thicker than normal, and there is not good demarcation between the
laminia propria and muscularis mucosa. The lamina propria also contains subtle horizontally-oriented
hyperintense striations. These striations could be fibrotic filaments formed as a result of healing following
RFA treatment. The BE gland is ovular and approximately 300 x 85 jim in cross section. Compared to the
untreated BE case presented in Section 7.4, the RFA-treated subject has significantly fewer BE glands.
The few scattered glands that do remain are buried beneath a thicker layer of normal-appearing tissue.
Buried glands are known to occur in some fraction of patients who have undergone ablative therapy
for BE, although the exact prevalence is not well understood and seems to vary significantly according to
the treatment type. Buried glands are thought to have reduced malignant potential compared to BE glands
exposed to the esophageal lumen, but the chance for progression to dysplasia and adenocarcinoma is
likely not reduced to zero. Buried glands are difficult to detect with endoscopic examination and pinch
biopsy due to their focal and subepithelial nature. 3D-OCT could be particularly well-suited for detection
and characterization of buried glands due to its large field of view and depth-resolved imaging
capabilities.
140
7.5.2 Comparison to Histology
Figure 7.9 shows a comparison of 3D-OCT data to conventional white light video endoscopy and
histology analysis of the region previously treated with RFA. Figure 7.9(A) shows an endoscopic video
capture obtained prior to 3D-OCT imaging of the same patient presented in Section 7.5.1. The esophageal
mucosa appears regular and is similar to the image of normal esophagus shown in Figure 7.3. Figure
7.9(B) shows histology of a region near the GE junction of the same patient following 3D-OCT imaging.
A discrepancy is observed between the histology image and the 3D-OCT image. Discrete glands buried
beneath a normal epithelium and lamina propria are visible in 3D-OCT images but not the histology
image. The size, shape, and hypointensity of the glands visible under 3D-OCT are highly suggestive of
buried BE glands beneath neo-squamous epithelium although these findings could not be confirmed by
pinch biopsy. The biopsy specimen was captured very close to the gastric mucosa and appears to show
gastric/squamous transitional tissue. This tissue region was classified as normal GE junction after analysis
of the histology data. Again, given the sparseness of the BE glands observed in the 3D-OCT image, it is
not surprising that a single biopsy specimen would randomly sample a normal area.
In total, 2 out of the 6 BE post-RFA patients (50%) with useable 3D-OCT data showed features
consistent with completely regular squamous epithelium and showed no signs of residual or recurrent BE
in regions that appeared endoscopically normal (no histology available for both patients). 4 out of 6
patients (67%) showed some signs of residual BE in regions that appeared endoscopically normal as
described above (2 BE under histology, 1 normal under histology, no histology available for 1 patient).
The situation of non-visually apparent, scattered, buried BE should be differentiated from patients
presenting for follow-up after RFA where large untreated regions of BE are visible endoscopically. This
situation is relatively common and is a result of incomplete ablation with RFA, and is rectified by further
RFA treatments to the BE site. The clinical implications of occult, sparse buried BE as implicated in
Figure 7.7 - Figure 7.9 are not entirely clear. It is possible that, if left untreated, the buried BE areas could
lead to recurrent wide-spread BE or even progress to dysplasia. The sparse gland distribution would make
detection and treatment of residual BE extremely challenging with conventional endoscopic techniques,
suggesting that 3D-OCT could play a role in follow-up assessment of RFA for BE.
7.6 Esophageal Nodules: Pre-Endoscopic Mucosal Resection
3D-OCT images of esophageal nodules in the presence of BE were obtained at 8 sites from 3 unique
patients. 1 patient had not previously been treated with RFA, EMR, or other endoscopic or surgical
therapies for BE or nodules. 1 patient was previously treated with RFA prior to 3D-OCT imaging to treat
the underlying BE condition. 1 patient had previously undergone EMR at a different site in the esophagus
to remove a previous dysplastic nodule. EMR was attempted on all 3 patients following 3D-OCT
141
imaging, with 2 out of 3 resections concluded successfully. I EMR attempt failed due to fibrosis at the
nodule site. All patients tolerated the 3D-OCT imaging procedure well and there were no known
immediate or long-term complications. Useable data was obtained for 2 out of the 3 patients. 1 patient
produced unusable data due to kinking of the probe sheath and subsequent severe frame-to-frame image
jitter. Results of the nodule pre-EMR imaging are described in Sections 7.6.1 to 7.6.2 below.
7.6.1 CharacteristicFeatures
Figure 7.10 shows typical 3D-OCT data obtained from an esophageal nodule prior to EMR resection.
This particular patient was not previously treated with RFA or EMR. Under white light video endoscopy
the esophageal mucosa appeared red and irregular with the nodule clearly visible. EMR to remove
esophageal nodules is not clinically indicated if the nodule is malignant and has invaded through the
submucosa, so 3D-OCT imaging could be particularly beneficial for assessment prior to EMR. Figure
7.10(A) shows a volumetric rendering of the entire 736 x 512 x 1919 (lines x axial pixels x frames, X x Y
x Z) dataset, which was acquired in -21 s. The pullback length was 20.7 mm, the rotational circumference
was 7.0 mm, and the imaging range was 1.6 mm in tissue. The red, orange, green, and blue boxes in
Figure 7.10 (A) illustrate the locations of orthogonal cut planes used to extract enface and cross-sectional
image slices.
Figure 7.10 (B) shows an en face image centered at a tissue depth of 175 jim, corresponding to the
centre of the squamous epithelial layer. The probe was pulled back over the GE junction during
acquisition of this dataset and a clear delineation can be appreciated between gastric mucosa on the lefthand side of the image and esophageal mucosa on the right. The gastric mucosa is, as expected, covered
with a regular pit pattern. The esophageal tissue at the far right appears relatively featureless. A -3 mm
section of tissue between the gastric and normal squamous regions (indicated by the black dashed box in
Figure 7.10(B)) contains highly irregular glandular structures. Figure 7.10(C) shows a second en face
image at a tissue depth of 450 gm. The glands extend down at least 275 gm and branching is observed at
deeper positions. The blue and green arrow pairs on the border of Figure 7.10(B,C) indicate the positions
of the cross-sectional images shown in Figure 7.1 0(D,E).
Figure 7.10(D,E) shows orthogonal cross-sectional images through the highly glandular region. In
Figure 7.10(D) normal gastric mucosa and relatively normal squamous mucosa are flanking the glandular
region. Based on the branching, irregular structure of the glands this region is assumed to be the
esophageal nodule. An orthogonal cross-section as shown in Figure 7.10(E) confirms the highly irregular
nature of the glands. Disruption of layered architecture is apparent. These glands can be differentiated
from the more typical BE glands by their larger size, extension from the luminal surface to > 500 im
tissue depths, and distinct branched appearance. These 3D-OCT features have, to the author's knowledge,
142
not been previously observed. Figure 7.11 shows an enlarged view of the ROI illustrated in Figure
7.10(B,C). This ROI covers a portion of the nodule region. Cross-sectional images in the ROI reveal loss
of layered architecture and concentrated branching glands. No distinct submucosal layer is present,
suggesting the possibility of invasion through to the muscularis propria consistent with malignant disease.
Previous 2D-OCT studies have not reported this type of branching glandular structure in HGD or
adenocarcinoma arising in a background of BE, however, so tissue classification based on the 3D-OCT
data is challenging.
7.6.2 Comparison to Histology
Figure 7.12 shows a comparison of 3D-OCT data to conventional white light video endoscopy and
histology analysis of the esophageal nodule. Figure 7.12(A) shows an endoscopic video capture obtained
prior to 3D-OCT imaging of the same patient presented in Section 7.6.1. The nodule is clearly visible
under white light examination. Figure 7.12(B) shows histology of the nodule following excision with
EMR. Good correlation is apparent between the histology image and the 3D-OCT image. Large, irregular
glands are visible in both the histology and 3D-OCT image with clear and widespread loss of the normal
squamous layered architecture. This patient was diagnosed with BE with squamoid metaplasia of the
mucin glands after analysis of the histology image.
In total, 1 out of the 2 nodule pre-EMR patients (100%) showed evidence of HGD under 3D-OCT
imaging (confirmed with histology). 1 patient, as described above, could not be classified using 3D-OCT
due to lack of comparable data (squamoid metaplasia under histology). While a larger study is required to
accurately determine the ability of 3D-OCT to stage esophageal nodules prior to EMR, these initial results
are encouraging.
7.7 Esophageal Nodules: Post-Endoscopic Mucosal Resection
3D-OCT images of regions previously treated with EMR to remove esophageal nodules were obtained
at 5 sites from 2 unique patients. All patients tolerated the 3D-OCT imaging procedure well and there
were no known immediate or long-term complications. Useable data was obtained for 2 out of the 2
patients. Results of the nodule post-EMR imaging are described in Sections 7.7.1 to 7.7.2 below.
7. 7.1 CharacteristicFeatures
Figure 7.13 shows typical 3D-OCT data obtained from a region previously treated with EMR to
remove an esophageal nodule. This particular patient had EMR to remove a nodule with HGD
approximately 2.5 months prior to 3D-OCT imaging. Under white light video endoscopy the treated
region appeared relatively normal except for a raised tissue ridge on the border of the EMR site. Previous
143
biopsy results from the EMR borders were positive for dysplasia although this could not be visually
confirmed under endoscopic examination. 3D-OCT analysis could be a useful tool to search the prior
EMR site for suspicious tissue, improving biopsy guidance and reducing sampling errors. Figure 7.13(A)
shows a volumetric rendering of the entire 980 x 512 x 1268 (lines x axial pixels x frames, X x Y x Z)
dataset, which was acquired in -21 s. The pullback length was 20.7 mm, the rotational circumference was
7.6 mm, and the imaging range was 1.7 mm in tissue. The red, orange, green, and blue boxes in Figure
7.13(A) illustrate the locations of orthogonal cut planes used to extract enface and cross-sectional image
slices.
Figure 7.13(B) shows an en face image centered at a tissue depth of 175 [tm, corresponding to the
centre of the squamous epithelial layer. The probe was pulled back over the GE junction during
acquisition of this dataset. The GE junction is less apparent here than in previous en face images due to
poor probe contact with the gastric mucosa. The esophageal epithelium appears relatively featureless and
regular at this tissue depth. The three dark bands running parallel to the pullback axis are polarization
artifacts caused by probe rotation and are not clinically relevant. Figure 7.13(C) shows a second enface
image at a tissue depth of 350 jtm. The border of the EMR site can be distinguished by a sharp change
from hyperintense tissue (possibly fibrosis) inside the EMR site to more hypointense tissue outside. A
small section of tissue (indicated by the dashed box in Figure 7.13(B,C)) at the distal EMR margin
contains small glandular structures consistent with BE, but no large irregularities consistent with invasive
disease or HGD are visible. The blue and green arrow pairs on the border of Figure 7.13(B,C) indicate the
positions of the cross-sectional images shown in Figure 7.10(D,E).
Figure 7.13(D,E) shows orthogonal cross-sectional images through the glandular region. In Figure
7.13(D), normal gastric mucosa and normal squamous mucosa are flanking the prior EMR site scar and
the region suspicious for BE, but no features indicative of HGD or invasive disease are visible. The scar is
characterized by hyperintense tissue layers of possible fibrosis. An orthogonal cross-section as shown in
Figure 7.13(E) confirms the presence of epithelial BE glands. Some moderate disruption of layered
architecture is present, suggesting the presence of LGD. Figure 7.14 shows an enlarged view of the ROI
illustrated in Figure 7.13(B,C). This ROI covers the BE region and the edge of the EMR scar. Crosssectional images in the ROI reveal ovular BE glands confined to the mucosa with some distortion of
layered architecture.
7.7.2 Comparisonto Histology
Figure 7.15 shows a comparison of 3D-OCT data to conventional white light video endoscopy and
histology analysis. Figure 7.15(A) shows an endoscopic video capture obtained prior to 3D-OCT imaging
of the same patient presented in Figure 7.13. The EMR treatment site presents with raised borders and
144
some irregular texture but cannot be evaluated visually for dysplasia. Figure 7.15(B) shows histology of
the distal border of the EMR treatment site. Excellent correlation is apparent between the histology image
and the 3D-OCT image in Figure 7.15(C). Small BE glands are visible in both the histology and 3D-OCT
images, along with some loss of layered structure. This patient was diagnosed with BE and LGD after
analysis of the histology image, consistent with the classification under 3D-OCT. The lateral EMR
margin was also positive for BE and LGD under histology. The proximal border was normal under
histology, consistent with the lack of BE glands shown in the enface OCT image in Figure 7.13(C).
In total, 2 out of the 2 nodule post-EMR patients (100%) showed evidence of BE and LGD under 3DOCT imaging (confirmed with histology for both patients). While a larger study is required to accurately
determine the ability of 3D-OCT to evaluate EMR treatment sites for disease recurrence or incomplete
resection, these initial results are encouraging.
145
7.8 Figures
Figure 7.1. 3D-OCT images of normal squamous mucosa in the esophagus. A: Volume rendering
showing location of orthogonal cut planes. B: En face XZ image of epithelium. Dashed box indicates
region of interest shown in Figure 3.3. C: YZ Cross-section. D: XY cross-section. Coloured arrows in B-D
indicate locations of other orthogonal cut planes.
146
Figure 7.2. Enlarged views of normal esophageal squamous mucosa in the region of interest indicated in
Figure 8.1(B). A-D: En face images at tissue depths corresponding approximately to the epithelium (A),
lamina propria (B), muscularis mucosa (C), and submucosa (D). Vascular features are visible in the
lamina propria (red triangle) and submucosa (black triangle). The same features can be identified in
cross-sectional images (E-G).
147
0
EP
o LP
MM-
SM ,
MP
Figure 7.3. Comparison of 3D-OCT data to conventional examination of normal esophageal mucosa. A:
White light video endoscopy image of mid-esophagus. B: Representative histology slide from excisional
biopsy specimen. C: Cross-sectional OCT image extracted from 3D dataset. EP, epithelium. LP, lamina
propria. MM, muscularis mucosa. SM, submucosa. MP, muscularis propria.
148
I
GA
I
L
NS
I
BE
Figure 7.4. 3D-OCT images of Barrett's esophagus islands near the GE junction. A: Volume rendering
showing location of orthogonal cut planes. B: En face XZ image at 175 um tissue depth. Dashed box
indicates region of interest shown in Figure 3.6. C: En face XZ image at 345 um tissue depth. Irregular
glandular features are apparent. D: YZ Cross-section. Gastric (GA), normal squamous (NS), and BE
regions can be distinguished. E: XZ cross-section showing gastric and transitionary tissue. F: XZ crosssection showing irregular BE glands. Coloured arrows in B-F indicate locations of other orthogonal cut
planes.
149
Figure 7.5. Enlarged views of BE island in the region of interest indicated in Figure 8.4(B,C). A: En face
image showing transition from normal esophagus to BE. B: YZ cross-sectional image illustrating transition
to BE. C: XZ cross-sectional image showing superficial glandular structure and distortion of mucosal
layers.
Figure 7.6. Comparison of 3D-OCT data to conventional examination of BE island. A: White light video
endoscopy image showing BE islands (IS) B: histology slide from excisional biopsy specimen showing BE
glands beneath superficial tissue (BE). C: Cross-sectional OCT image extracted from 3D dataset. BE
glands are visible adjacent to normal squamous epithelium.
150
Figure 7.7. 3D-OCT images of region previously treated with RFA due to BE. A: Volume rendering
showing location of orthogonal cut planes. B: En face XZ image at 100 um tissue depth. Dashed box
indicates region of interest shown in Figure 3.9. C: En face XZ image at 390 um tissue depth. Scattered,
buried glandular features are apparent. D: YZ Cross-section. Gastric (GA) and normal squamous (NS)
regions can be distinguished. Scattered glands possibly representing buried BE (?BE) are buried 350 400 um beneath tissue surface. E: XZ cross-section showing scattered glands (G). Coloured arrows in BE indicate locations of other two orthogonal cut planes.
151
Figure 7.8. Enlarged views of possible buried BE glands (?BE) in region of interest indicated in Figure
8.7(B,C). A: En face image showing ovular lumen shape. B: YZ cross-sectional image. C: XZ crosssectional image. Hyperintense striations in the lamina propria and an isolated gland are visible.
Figure 7.9. Comparison of 3D-OCT data to conventional examination of RFA-treated region. A: White
light video endoscopy image. GE junction is extended proximally at site of prior RFA (triangle). B:
histology slide from excisional biopsy specimen near GE junction. C: Cross-sectional OCT image
extracted from 3D dataset. Small glands, possibly scattered buried BE glands (?BE), are visible buried
within normal squamous epithelium.
152
I
GA
II
NO
J
I
NS
I
Figure 7.10. 3D-OCT images of a nodule near the GE junction. A: Volume rendering showing location of
orthogonal cut planes. B: En face XZ image at 100 um tissue depth. Dashed box indicates region of
interest shown in Figure 3.12. C: En face XZ image at 600 um tissue depth. Region of large, branching
glands and vessels is apparent. D: YZ Cross-section. Gastric (GA), nodule (NO), and normal squamous
(NS) regions can be distinguished. Large branching glands (BG) extend to the tissue surface. E: XZ
cross-section showing highly irregular branching glands (BG). Coloured arrows in B-E indicate locations
of other two orthogonal cut planes.
153
Figure 7.11. Enlarged views of an esophageal nodule in the region of interest indicated in Figure
8.10(B,C). A: En face image showing nodule region. B: XZ cross-sectional image showing irregular,
branching glandular structure (BG) and disruption of mucosal layers. C: YZ cross-sectional image.
Figure 7.12. Comparison of 3D-OCT data to conventional examination of an esophageal nodule. A:
White light video endoscopy image. Glandular nodule is visible (triangle). B: histology slide from
excisional biopsy specimen. C: Cross-sectional OCT image extracted from 3D dataset. Large, irregular
branching glands (BG) are visible.
154
I
GA
I
I
I
I
A:
Figure 7.13. 3D-OCT images of region previously treated with EMR to remove an esophageal nodule.
tissue
um
175
at
image
XZ
face
En
B:
planes.
cut
orthogonal
of
location
Volume rendering showing
depth. Dashed box indicates region of interest shown in Figure 8.15. C: En face XZ image at 350 um
tissue depth. Small concentration of BE (BE) glands is apparent, and outline of EMR site can be
distinguished (triangles) D: YZ Cross-section. Gastric (GA), BE region (BE), EMR scar (SC) and normal
squamous (NS) regions can be distinguished. E: XZ cross-section showing BE glands (BE). Coloured
arrows in B-E indicate locations of other orthogonal cut planes.
155
in Figure 8.13(B,C).
Figure 7.14. Enlarged views of BE zone following EMR in region of interest indicated
BE glands. B: XZ
of
group
A: En face image showing BE region. B: YZ cross-sectional image showing
cross-sectional image showing three BE glands.
site.
Figure 7.15. Comparison of 3D-OCT data to conventional examination of distal margin of prior EMR
histology
B:
(triangle).
visible
is
A: White light video endoscopy image. Prior EMR site with raised margins
BE
slide from excisional biopsy specimen. C: Cross-sectional OCT image extracted from 3D dataset. Six
glands (BE) are visible.
156
7.9 References
[1]
N. J.
Shaheen,
"Advances
in Barrett's
esophagus
and esophageal
adenocarcinoma,"
Gastroenterology,vol. 128, pp. 1554-66, May 2005.
[2]
A. Jemal, R. Siegel, E. Ward, T. Murray, J. Xu, and M. J. Thun, "Cancer statistics, 2007," CA
CancerJ Clin, vol. 57, pp. 43-66, Jan-Feb 2007.
[3]
B. F. Overholt, M. Panjehpour, and J. M. Haydek, "Photodynamic therapy for Barrett's
esophagus: follow-up in 100 patients," GastrointestEndosc, vol. 49, pp. 1-7, Jan 1999.
[4]
X. Chen and C. S. Yang, "Esophageal adenocarcinoma: a review and perspectives on the
mechanism of carcinogenesis and chemoprevention," Carcinogenesis,vol. 22, pp. 1119-29, Aug
2001.
[5]
A. J. Cameron and H. A. Carpenter, "Barrett's esophagus, high-grade dysplasia, and early
adenocarcinoma: a pathological study," Am J Gastroenterol,vol. 92, pp. 586-91, Apr 1997.
[6]
S. J. Spechler, "Screening and surveillance for complications related to gastroesophageal reflux
disease," Am J Med, vol. 111 Suppl 8A, pp. 130S-136S, Dec 3 2001.
[7]
H. Mashimo, M. S. Wagh, and R. K. Goyal, "Surveillance and screening for Barrett esophagus
and adenocarcinoma," J Clin Gastroenterol,vol. 39, pp. S33-41, Apr 2005.
[8]
J. J. Nigro, J. A. Hagen, T. R. DeMeester, S. R. DeMeester, J. Theisen, J. H. Peters, and M.
Kiyabu, "Occult esophageal adenocarcinoma: extent of disease and implications for effective
therapy," Ann Surg, vol. 230, pp. 433-8; discussion 438-40, Sep 1999.
[9]
M. S. Dar, J. R. Goldblum, T. W. Rice, and G. W. Falk, "Can extent of high grade dysplasia in
Barrett's oesophagus predict the presence of adenocarcinoma at oesophagectomy?," Gut, vol. 52,
pp. 486-9, Apr 2003.
[10]
J. M. Collard, "High-grade dysplasia in Barrett's esophagus. The case for esophagectomy," Chest
Surg Clin NAm, vol. 12, pp. 77-92, Feb 2002.
[11]
G. W. Falk, T. W. Rice, J. R. Goldblum, and J. E. Richter, "Jumbo biopsy forceps protocol still
misses unsuspected cancer in Barrett's esophagus with high-grade dysplasia," Gastrointest
Endosc, vol. 49, pp. 170-6, Feb 1999.
[12]
B. J. Dunkin, J. Martinez, P. A. Bejarano, C. D. Smith, K. Chang, A. S. Livingstone, and W. S.
Melvin, "Thin-layer ablation of human esophageal epithelium using a bipolar radiofrequency
balloon device," Surgical Endoscopy and Other Interventional Techniques, vol. 20, pp. 125-130,
Jan 2006.
157
[13]
H. Mashimo, Y. Chen, S. W. Huang, Q. Huang, A. Aguirre, J. Schmitt, and J. Fujimoto,
"Endoscopic optical coherence tomography reveals Barrett's underneath squarnous neoepithelium after radiofrequency ablation," Gastroenterology,vol. 132, pp. A96-A96, Apr 2007.
[14]
J. P. Byrne, G. R. Armstrong, and S. E. Attwood, "Restoration of the normal squamous lining in
Barrett's esophagus by argon beam plasma coagulation," Am J Gastroenterol,vol. 93, pp. 1810-5,
Oct 1998.
[15]
C. J. Kelty, R. Ackroyd, N. J. Brown, T. J. Stephenson, C. J. Stoddard, and M. W. Reed,
"Endoscopic ablation of Barrett's oesophagus: a randomized-controlled trial of photodynamic
therapy vs. argon plasma coagulation," Aliment PharmacolTher, vol. 20, pp. 1289-96, Dec 2004.
[16]
J. L. Van Laethem, M. O. Peny, I. Salmon, M. Cremer, and J. Deviere, "Intramucosal
adenocarcinoma arising under squamous re-epithelialisation of Barrett's oesophagus," Gut, vol.
46, pp. 574-7, Apr 2000.
[17]
K. Ragunath, N. Krasner, V. S. Raman, M. T. Haqqani, C. J. Phillips, and I. Cheung, "Endoscopic
ablation of dysplastic Barrett's
oesophagus comparing
argon plasma
coagulation
and
photodynamic therapy: a randomized prospective trial assessing efficacy and cost-effectiveness,"
ScandJ Gastroenterol,vol. 40, pp. 750-8, Jul 2005.
[18]
R. Kiesslich, M. von Bergh, M. Hahn, G. Hermann, and M. Jung, "Chromoendoscopy with
indigocarmine improves the detection of adenomatous and nonadenomatous lesions in the colon,"
Endoscopy, vol. 33, pp. 1001-1006, Dec 2001.
[19]
D. C. Adler, Y. Chen, R. Huber, J. Schmitt, J. Connolly, and J. G. Fujimoto, "Three-dimensional
endomicroscopy using optical coherence tomography," Nature Photonics, vol. 1, pp. 709-716,
Dec 2007.
[20]
Y. Chen, A. D. Aguirre, P. L. Hsiung, S. Desai, P. R. Herz, M. Pedrosa, Q. Huang, M.
Figueiredo, S. W. Huang, A. Koski, J. M. Schmitt, J. G. Fujimoto, and H. Mashimo, "Ultrahigh
resolution optical coherence tomography of Barrett's esophagus: preliminary descriptive clinical
study correlating images with histology," Endoscopy, vol. 39, pp. 599-605, Jul 2007.
158
CHAPTER 8
8.0
Clinical 3D-OCT in the Lower Gastrointestinal Tract
159
8.1 Motivation
Chapter 7 described initial clinical study results focusing on the use of 3D-OCT endomicroscopy for
pre- and post-treatment assessment of Barrett's esophagus (BE) and esophageal nodules arising from BE.
Similar applications for 3D-OCT can also be developed in the lower gastrointestinal (GI) tract. Colorectal
cancer is another common GI disease with high morbidity and mortality rates. CRC is the third leading
cause of cancer death, accounting for about 10% of cancer deaths overall [1]. Despite its high incidence,
colorectal cancer is one of the most detectable, and, if found early, most treatable forms of cancer.
Although most colorectal cancers arise from adenomatous polyps that are detectable using conventional
endoscopy, many flat (non-polypoid) lesions are missed during routine exams [2]. Up to 50% of these
more subtle lesions are missed by conventional endoscopy [3].
Detection of early-stage cancer is particularly difficult in patients with inflammatory bowel diseases
(IBD) such as ulcerative colitis (UC) and Crohn's disease (CD), where neoplastic tissue is often flat rather
than polypoid in form and multifocal in distribution [4]. As many as 1.4 million individuals in the United
States have inflammatory bowel disease [5] and are at increased risk for the development of colorectal
cancer [6]. UC in particular affects up to 780,000 individuals in the United States and Canada and is
newly diagnosed in 7000 - 46,000 individuals per year [5]. UC is associated with a -5x increase in risk of
developing colorectal cancer compared to the general population, with colorectal cancer accounting for
one sixth of all deaths in UC patients [6]. Unfortunately, early-stage dysplastic lesions are often flat,
diffuse and multifocal in these individuals [4]. As a result, dysplastic lesions are easily obscured by the
gross inflammatory background of UC, making early detection extremely challenging. The standard of
care for inflammatory bowel disease is similar to BE, with random biopsies being performed throughout
the length of the affected region. Colonoscopy with biopsy is generally performed when the disease is in
remission to reduce the risk of complications. 3D-OCT, on the other hand, could be used to visualize
cross-sectional tissue microstructure without biopsy and with practically no added risk of bleeding in IBD
patients. 3D-OCT could also potentially be used to differentiate early dysplastic progression from
inflammation based on variations in architectural distortion. As in BE, image guided biopsy and therapy
for IBD patients using 3D-OCT promises to improve diagnostic and therapeutic procedures by providing
comprehensive 3D tissue visualization.
Radiation proctitis (RP) is another form of chronic inflammation and injury of the lower GI tract. This
condition is a common side effect of radiation therapy used to treat prostate cancer and occurs in some
form in approximately 10 - 15% of patients. RP causes lower GI bleeding and is generally treated with
electrocautery or argon plasma coagulation. Healing from these deeply-penetrating ablative methods can
be problematic in the area of prior RP, with a high likelihood of ulcerations, structuring, or further
160
bleeding. More recently, radiofrequency ablation (RFA) has been studied as a novel treatment for RP.
RFA produces a superficial ablation field covering several square centimeters and has been demonstrated
to be effective in treating BE as well as RP in several pilot studies. Since excisional biopsy is
contraindicated in RP patients due to an increased risk of bleeding, conventional endoscopy is the primary
modality used to assess the disease and the relevant interventional therapies. 3D-OCT can be applied in
RP patients to provide subsurface tissue views that are not available with conventional endoscopy.
Subsurface information can be used to analyze the extent of RP, plan interventions based on this data,
evaluate the healing process on follow-up, and check for recurrence over time.
In this section of the thesis work, the same 3D-OCT endomicroscopy system and spiral-scanning
imaging catheters used during upper endoscopy were used to assess four tissue categories in the lower GI
tract: 1) no pathology ("Normal"); 2) inflammatory bowel diseases ("IBD"); 3) radiation proctitis prior to
undergoing RFA therapy ("RP pre-RFA"); and 4) sites previously treated with RFA to control RP ("RP
post-RFA"). 3D-OCT was used to establish characteristic features of colonic, rectal, and anal mucosa for
each tissue group. 3D-OCT was found to be particularly valuable for pre-treatment and follow-up
assessment of RP patients due to the contraindication of excisional biopsy.
8.2 Clinical Protocol
All subjects in this study were imaged at the Veterans Affairs Boston Healthcare System (VABHS)
Jamaica Plain campus in collaboration with Dr. Hiroshi Mashimo, MD PhD, and Dr. Qin Huang, MD.
Subjects were recruited from the pool of patients undergoing elective colonoscopy for screening,
surveillance, or treatment of IBD or RP. Subjects followed a standard preparation procedure for
colonoscopy, including cessation of blood thinners, fasting, and colonic clearing prior to the appointment.
3D-OCT imaging was performed in tandem with standard video endoscopy.
The imaging protocol was very similar to the procedure described in Chapter 7, Section 7.2. For
normal subjects, the endoscopist identified regions for 3D-OCT imaging that appeared normal and
unremarkable using white light video endoscopy. Imaging was performed by inserting the 3D-OCT
catheter down the working channel of a standard endoscope. For IBD patients the endoscopist identified
regions of active or controlled disease for 3D-OCT imaging using visual markers such as inflammation,
edema, and fibrosis. For RP pre-RFA subjects the endoscopist identified locations suspicious for RP
using similar criteria in patients with a history of prostate cancer and previous radiation therapy. For RP
post-RFA subjects, the endoscopist identified regions of prior treatment using patient history notes
describing the extent and location of the treated region relative to the anal verge. In some cases subtle
visual cues could also be used to differentiate treated tissue from untreated tissue. Where clinically
161
indicated, pinch biopsy samples were obtained from the 3D-OCT imaging sites following the conclusion
of all image acquisition.
Between 3-10 sites were assessed for each subject, increasing the total procedure time by an average
of 11 minutes and by no longer than 20 minutes. This increase in procedure length is similar to the time
increase associated with chromoendoscopy [3]. After the acquisition of all 3D-OCT datasets, the imaging
catheter was withdrawn from the working channel and replaced with pinch biopsy forceps to acquire
biopsies of the imaged tissue where clinically indicated. Biopsies were not obtained from normal subjects
due to the slight risk associated with pinch biopsy and the lack of clinical necessity. Biopsies were also
not obtained from RP patients due to a significant increase in the risk of bleeding. Histopathology from
pinch biopsies was evaluated by an expert pathologist in order to establish a tissue classification for each
site. The histological diagnoses were then compared to tissue features based on the 3D-OCT data where
applicable. Imaging parameters were identical to those described in Chapter 7, Section 7.2.
In total, 75 sites from 15 unique patients were imaged during this study from February to December
2008. Table 8.1 shows a summary of the collected 3D-OCT image data. Note that the total number of
patients in Table 8.1 exceeds 15 since multiple tissue types were imaged from several patients, typically
consisting of a normal region and a pathological region. The "other" grouping includes pathologic tissue
that was outside the main scope of the study and primarily consisted of ulcers and polyps. The following
sections show representative 3D-OCT data, characteristic features, and comparisons to histology for each
main tissue type.
Tissue Type
Normal Colon
IBD
RP Pre-RFA
RP Post-RFA
Other Pathology
Number of
Patients
12
1
2
3
8
Number of
Imaging Sites
28
2
3
11
31
Table 8.1. Summary of 3D-OCT data collected during lower endoscopy.
8.3 Normal Colon
Obtaining normal image data was a necessary first step in this study in order to establish a baseline for
comparison to pathologic tissue. 3D-OCT images of normal colonic mucosa were obtained at 28 sites
from 12 unique patients. All patients tolerated the procedure well and there were no known immediate or
long-term complications. Useable data was obtained for 10 out of the 12 patients. 2 patients produced
unusable data due to nonuniform probe rotation caused by excessive pressure on the colonoscope. Results
of the normal colon imaging are described in Sections 8.3.1 to 8.3.2 below.
162
8.3.1 CharacteristicFeatures
Figure 8.1 shows typical 3D-OCT data obtained from the normal colon. Figure 8.1(A) shows a
volumetric rendering of the entire 980 x 512 x 1268 (lines x axial pixels x frames, X x Y x Z) dataset,
which was acquired in -21 s. The pullback length was 20.7 mm, the rotational circumference was 7.5
mm, and the imaging range was 1.7 mm in tissue. The images in Figure 8.1 have been axially cropped to
remove the probe sheath and non-useable points beyond the penetration depth of light in the tissue. The
useable imaging depth in tissue was approximately 1100 jtm. Unless otherwise specified, all crosssectional and en face images shown below are formed by averaging the volumetric data over a 20 jtmthick section to reduce speckle noise. Since 20 jtm is approximately the size of a single epithelial cell,
minimal image blurring is observed from this process [7]. The red, green, and blue boxes in Figure 8.1(A)
illustrate the locations of orthogonal cut planes used to extract enface and cross-sectional image slices for
closer examination. In this and all subsequent 3D-OCT image data the X axis represents the probe's
rotational axis (fast, 50 - 70 Hz), the Y axis represents the axial OCT line data, and the Z axis represents
the probe's pullback direction (slow, 500 - 1000 jtm/s).
Figure 8.1(B) shows an en face image centered at a tissue depth of 300 jim, corresponding to the
bottom of the columnar epithelial layer. This image is similar in concept to a conventional microscopy
image obtained by setting the focus to a depth of 300 jtm. The left side of the image is more proximal
colon and the ride side is more distal (anal). Consistent with previous OCT studies on excised pathology
lab samples [8] the normal columnar epithelium of the colon shows low light penetration. Colonic crypts
are visible as round or slightly oval regions of hyperintensity, thought to be caused by increased light
transmission through the crypt lumen [8]. The large hyperintense region at a pullback distance of
approximately 3 mm is scar tissue from a prior polypectomy at this site. Transverse probe jitter is reduced
in the colon compared to the esophagus due to a significant reduction in respiratory motion and a
complete lack of heartbeat motion. Longitudinal (Z-oriented) artifacts are primarily caused by variations
in probe pullback velocity due to periodic sticking and slipping within the outer sheath. The blue and
green arrow pairs on the border of Figure 8.1(B) indicate the positions of the cross-sectional images
shown in Figure 8.1(C,D).
Figure 8.1(C) shows a YZ cross-sectional image at the position indicated by the green arrow pair in
Figure 8.1 (B). In regions outside of the prior polypectomy site vertical streaks are present due to increase
light transmission through crypt lumens. The polypectomy site exhibits a hyperintense layer of fibrotic
scar tissue covered by a featureless tissue layer. It is difficult to discern a layered architecture at most
locations in the colon due to poor transmission through the columnar epithelium. Figure 8.1(D) shows an
163
XY cross-sectional image at the position indicated by the blue arrow pair in Figure 8.1(B). Similar
features to the cross-section in Figure 8.1(C) can be appreciated.
The dense spatial sampling density, high optical resolution, and large field of view of 3D-OCT
endomicroscopy enables high-magnification analysis of features such as colonic crypts over arbitrary
regions of interest (ROI). Crypts are the main glandular structures in the human colon, and changes in
crypt size and appearance are associated with the earliest forms of colorectal cancer [9] and other
diseases. The ability to assess the 3D structure of crypts, possibly using automated classification
algorithms [10], is therefore of potential value for future applications in cancer detection and treatment.
Figure 8.2(A) shows an en face images of the ROI shown as a dashed box in Figure 8.1 (B). To generate
the clearest view of the crypts, the axial data was averaged over a depth of 480 tm. The average crypt
diameter is approximately 75 - 150 tm. Variations in apparent crypt size can be attributed to variations in
tissue orientation and some changes in probe pullback velocity due to stress on the probe from the
endoscope. Figure 8.2(B,C) shows the expected regular hyperintense streaks associated with the crypt
lumens.
8.3.2 Comparison to Histology
Comparison of 3D-OCT endomicroscopy images to histology slides obtained from excisional biopsy
and to white light video endoscopy examination is a good method for linking image features to known
tissue structures. Figure 8.3(A) shows an endoscopic video capture obtained prior to 3D-OCT imaging of
the same patient presented in Section 8.3.1. The colonic mucosa appears pale, smooth, and regular,
consistent with normal columnar epithelial mucosa. Figure 8.3(B) shows representative histology of
normal colonic glandular mucosa. Since there was no clinical indication to biopsy this particular patient,
the histology image in Figure 8.3(B) is taken from another subject where biopsy was required but a
normal diagnosis was made by the pathologist. Good correlation is shown between the histology image
and the 3D-OCT image. The epithelium, muscularis mucosa, and submucosa are visible under histology
and under 3D-OCT as shown in Figure 8.3(C). The lamina propria is not generally distinguishable from
the epithelium under 3D-OCT due to its close integration with the epithelial crypts. The muscularis
mucosa and submucosa boundary is visible at this particular location but is generally not appreciable. In
total, 10 out of the 10 normal patients (100%) with useable 3D-OCT data showed features consistent with
those described above.
8.4 Inflammatory Bowel Diseases
3D-OCT images of active UC were obtained at 2 sites from 1 unique patient to demonstrate imaging in
an IBD case. The patient tolerated the procedure well and there were no known immediate or long-term
164
complications. This patient had previously undergone a partial colectomy related to his UC, with the
remaining colon pulled down and re-attached to the anus and rectum. He presented with endoscopically
apparent active UC in a small region near the squamo-columnar junction. The squamo-columnar junction
in the colon is sometimes referred to as the "dentate line". Useable data was obtained without significant
motion artifacts. Results of the UC imaging are described in Sections 8.4.1 to 8.4.2 below.
8.4.1 CharacteristicFeatures
Figure 8.4 shows 3D-OCT data obtained from a region of active UC. Figure 8.4(A) shows a
volumetric rendering of the entire 771 x 512 x 1794 (lines x axial pixels x frames, X x Y x Z) dataset,
which was acquired in -21 s. The pullback length was 20.7 mm, the rotational circumference was 8.6
mm, and the imaging range was 1.7 mm in tissue. The images in Figure 8.4 have been axially cropped to
remove the probe sheath and non-useable points beyond the penetration depth of light in the tissue. The
useable imaging depth in tissue was approximately 1100 lm. The red, green, blue and purple boxes in
Figure 8.4(A) illustrate the locations of orthogonal cut planes used to extract en face and cross-sectional
image slices for closer examination.
Figure 8.4(B) shows an en face image centered at a tissue depth of 200 jtm, corresponding to where
the columnar epithelial layer should be located in normal mucosal tissue. A complete lack of regular crypt
architecture is immediately apparent from inspection of the en face 3D-OCT image. Large subsurface
voids and bands of hyperscattering tissue, possibly fibrotic, are clearly visualized. A wedge of
comparatively normal squamous epithelial mucosa is visible at the right of the image, demarcating the
boundary between pathologic UC and the squamous epithelial mucosa of the anal verge. The blue and
green arrow pairs on the border of Figure 8.4(B) indicate the positions of the cross-sectional images
shown in Figure 8.4(C,D).
Figure 8.4(C) shows a YZ cross-sectional image at the position indicated by the green arrow pair in
Figure 8.4(B). UC and normal squamous epithelial mucosa are visible on the left and right sides of the
image respectively. The large -1.5 mm subsurface void near the centre of Figure 9.4(C) may be an
ulcerative tunnel formed by repeated epithelial stripping and regeneration during disease remission. A
large number of superficial vessels and edematous regions are present in the UC section, along with a lack
of regular squamous or columnar epithelial architecture. Figure 8.4(D) shows an XZ cross-sectional
image at the position indicated by the blue arrow pair in Figure 8.4(B). Similar ulcerative features are
visible here as well.
Figure 8.5 and Figure 8.6 show detailed views of the two ROI's indicated in Figure 8.4(B). Figure
9.5(A) shows an en face view of an ulcerative region with tortuous and irregular superficial vasculature.
The cross-sectional images in Figure 8.5(B,C) show details of the UC architecture. There is a stark
165
contrast between this region and the region shown in Figure 8.6. The en face view in Figure 8.6(A) at the
same tissue depth is featureless and unremarkable, consistent with squamous epithelial mucosa. This is
confirmed with the cross-sectional images in Figure 8.5(B,C). The tissue shown in Figure 9.5(B,C)
contains subtle hyperintense striations in the epithelium not found in the squamous epithelial mucosa of
completely healthy subjects. These horizontal bands of tissue may be fibrotic structures resulting from
healing during disease remission.
8.4.2 Comparison to Histology
Figure 8.7(A) shows an endoscopic video capture obtained prior to 3D-OCT imaging of the same
patient presented in Figure 8.4. The mucosa appears red, inflamed and ulcerative. The 3D-OCT catheter is
also shown in position prior to imaging. Figure 8.7(B) shows representative histology of UC in glandular
mucosa. Since there was an increased risk of bleeding in this particular patient, the histology image in
Figure 8.7(B) is taken from another subject with a similar case of active UC. Lymphocytic mucosal
infiltration is present along with submucosal fibrosis. Ulceration results in the formation of a pseudopolyp as the epithelium is stripped away to expose the submucosa. Good correlation is shown between the
histology image and the 3D-OCT image. Large edematous regions, loss of normal crypt patterns, and
disruption of the epithelium are apparent under 3D-OCT and histology. While more study is required to
expand on these findings, 3D-OCT could potentially be useful in assessing the severity of IBD,
monitoring response to therapy, and differentiating dysplastic lesions from the gross inflammatory
background.
8.5 Radiation Proctitis: Pre-Radiofrequency Ablation
3D-OCT images of RP prior to receiving RFA were obtained at 3 sites from 2 unique patients. This
data was collected in order to establish characteristics of active, untreated RP and forms a baseline for
comparison of treated RP and for assessing the effects of RFA. Both of these patients had not previously
been treated with RFA. 1 patient had previously undergone argon plasma coagulation to control the RP
but had experienced additional bleeding subsequent to treatment. Both patients tolerated the 3D-OCT
imaging procedure well and there were no known immediate or long-term complications. Useable data
was obtained for 2 out of the 2 patients. Results of the RP pre-RFA imaging are described in Sections
8.5.1 to 8.5.2 below.
8.5.1 CharacteristicFeatures
Figure 8.8 shows typical 3D-OCT data obtained from an ectatic RP region prior to any endoscopic or
surgical therapy. This patient presented with rectal bleeding following 8 weeks of radiation therapy for
166
prostate cancer. The patient completed radiation therapy in mid-2001 and was imaged with 3D-OCT in
2008. It is not uncommon for RP to develop many years after completion of a radiation course.
Endoscopic examination revealed irregular and ectactic blood vessels along with mucosal inflammation in
the rectum consistent with RP. In this case the goal of 3D-OCT imaging was not to assist in making a
diagnosis of RP, but rather to obtain subsurface tissue information for pathology characterization. Since
excisional biopsies cannot be performed in RP patients due to a significantly increased risk of bleeding,
3D-OCT is one of the only ways to obtain such information. Figure 8.8(A) shows a volumetric rendering
of the entire 1202 x 512 x 1240 (lines x axial pixels x frames, X x Y x Z) dataset, which was acquired in
-21 s. The pullback length was 20.7 mm, the rotational circumference was 9.0 mm, and the imaging
range was 1.7 mm in tissue. The red, green and blue boxes in Figure 8.8(A) illustrate the locations of
orthogonal cut planes used to extract enface and cross-sectional image slices for closer examination.
Figure 8.8(B) shows an en face image centered at a tissue depth of 250 pim, corresponding to a depth
slightly lower than the centre of the columnar epithelial layer. A regular crypt pattern is visible over most
of the tissue, although transverse probe jitter caused by nonuniform rotation partially obscures the
features. Several large vessels can be seen near the top right of the image (triangles). Vessels of this
diameter (-500 pm) are not normally found at superficial depths and are indicative of ectasia arising from
RP. The blue and green arrow pairs on the border of Figure 8.8(B) indicate the positions of the crosssectional images shown in Figure 8.8(C,D). Figure 8.8(C,D) shows cross-sectional images bisecting the
irregular vessels identified in Figure 8.8(B). The ability to align cross-sectional image planes to surface
features such as ectatic vessels is a significant advantage of 3D-OCT. In Figure 8.8(C,D), the vessels are
visible as shadowing and slightly hypointense structures located in the epithelial layer (triangles). The rest
of the tissue appears fairly normal and exhibits typical hyperintense streaking associated with colonic
crypts.
Figure 8.9 shows an enlarged view of the ROI illustrated in Figure 8.8(B). This ROI spans a
superficial vessel that is associated with RP. The vessel is clearly identifiable in the en face image as a
large, irregular hypointense structure (triangles). Cross-sectional images in the ROI highlight the
extremely superficial nature of the vessels. The overlying tissue is less than 150 pm thick, suggesting a
high likelihood of bleeding resulting from slight agitation such as stool passage. Vessels of this large size
and superficial nature are not seen in the squamous epithelial mucosa of normal patients. In regions
immediately adjacent to the vessel there is a lack of normal colonic crypts and presence of a quasi-regular
layered structure, indicating that the vessels are embedded under the squamous epithelial mucosa of the
anal verge.
167
8.5.2 Comparison to Histology
Figure 8.10 shows a comparison of 3D-OCT data to conventional white light video endoscopy and
histology analysis. Figure 8.10(A) shows an endoscopic video capture obtained prior to 3D-OCT imaging
of the same patient presented in Section 8.5.1. Excellent correlation is shown between the histology image
in Figure 8.10(B) and the 3D-OCT image in Figure 8.10(C). The colonic mucosa contains inflammatory
infiltrates and large superficial vessels with a disrupted crypt pattern. In total, 2 of the 2 RP pre-RFA
patients (100%) showed features consistent with those described above. A significant degree of variability
was observed in vessel size and orientation, although in both cases the vessels were superficial and
irregular. The ability to obtain enface and cross-sectional tissue architecture without excisional biopsy is
a major advantage for RP patients, enabling baseline tissue characterization for later analysis of treatment
effects.
8.6 Radiation Proctitis: Post-Radiofrequency Ablation
3D-OCT images of RP subsequent to receiving RFA were obtained at 11 sites from 2 unique patients.
This data was compared to pre-treatment images to study the effects of RFA for treating RP, which is a
less-established application than treatment of BE. One patient was imaged 12 months and 14 months after
receiving RFA therapy, and in both cases useable 3D-OCT data was obtained. The other patient was
imaged within 10 minutes of receiving RFA. This did not produce useable 3D-OCT data due to the
presence of tissue debris and blood in the imaging field resulting from the application of RFA. Both
patients tolerated the procedure well and there were no known immediate or long-term complications.
Results of the RP post-RFA imaging are described in Sections 8.6.1 to 8.6.2 below.
8.6.1 CharacteristicFeatures
Figure 8.11 shows typical 3D-OCT data obtained from a formerly ectatic RP region that was treated
with RFA 14 months prior to imaging. This patient presented with no signs or symptoms of active
bleeding or proctitis since receiving RFA. Endoscopic examination revealed regular columnar epithelial
mucosa in the colon and smooth squamous epithelial mucosa in the rectum consistent with resolution of
the prior RP. No large ectatic vessels or overt inflammation were detected under white light video
endoscopy. Figure 8.11(A) shows a volumetric rendering of the entire 771 x 512 x 1794 (lines x axial
pixels x frames, X x Y x Z) dataset, which was acquired in -21 s. The pullback length was 20.7 mm, the
rotational circumference was 7.2 mm, and the imaging range was 1.7 mm in tissue. The red, green and
blue boxes in Figure 8.11(A) illustrate the locations of orthogonal cut planes used to extract en face and
cross-sectional image slices.
168
Figure 8.11(B) shows an enface image centered at a tissue depth of 350 tpm, corresponding to a depth
slightly beneath the epithelial layer. The probe was pulled back over the dentate line and a clear
demarcation between the proximal columnar epithelial mucosa and the distal (anal) squamous epithelial
mucosa is apparent. A regular crypt pattern is visible over the region containing columnar epithelium on
the left side of the image. The large hypointense structure at the far left is a hemorrhoid as noted during
endoscopic examination of the region. The squamous epithelial mucosa at this depth, likely corresponding
to lamina propria, is hyperintense and featureless with the exception of several large vascular structures
(triangles). These structures are very similar to the ectatic vessels visualized prior to RFA treatment in
Figure 8.8, but at 14 months post-RFA they are now found at a deeper tissue location. The blue and green
arrow pairs on the border of Figure 8.11(B) indicate the positions of the cross-sectional images shown in
Figure 8.11(C,D).
Figure 8.11 (C,D) shows cross-sectional images bisecting the vessels identified in Figure 8.11(B). The
combination of en face and cross-sectional data enables rapid identification and localization of vascular
structures as well as 3D architectural analysis. In Figure 8.11(C,D), the vessels are visible as nonshadowing and slightly hypointense structures (triangles) located beneath the epithelial and lamina
propria layers of regular-appearing squamous epithelial mucosa. The rest of the tissue appears fairly
normal and exhibits typical layered architecture consistent with squamous epithelial mucosa. In
comparison to the pre-treatment images shown in Figure 8.8, the vessels exposed to RFA are deeply
buried and show a reduction in axial shadowing that may coincide with reduced blood flow from reduced
vessel diameter, although this conclusion is speculative at this point.
Figure 8.12 shows an enlarged view of the ROI illustrated in Figure 8.11(B). This ROI spans a buried
vessel that is associated with a previously ectatic vessel that had been treated with RFA. Again, the vessel
is clearly identifiable in the en face image as a large, irregular hypointense structure (triangles). Crosssectional images in the ROI highlight the difference in tissue architecture as a result of RFA. The vessel is
covered in a thick, protective layer of squamous epithelium and lamina propria approximately 300 iim
thick. The untreated vessels, by comparison, were covered in a thin tissue cap less than 150 Rm thick. The
presence of squamous epithelium over top of the vessels could be a result of neo-squamous growth in a
region previously occupied by glandular columnar epithelial mucosa. Neo-squamous re-epithelialization
has been observed using 2D-OCT in the study of RFA for BE treatment, and a similar process could also
occur in the colon.
8.6.2 Comparison to Histology
Figure 8.13 shows a comparison of 3D-OCT data to conventional white light video endoscopy. No
histology data was available for the post-RFA subjects due to the possibility of incomplete treatment and
169
therefore an elevated risk of bleeding. Additionally, the squamous epithelial mucosa of the anal verge is
innervated and therefore pinch biopsy is a painful procedure that can only be performed when there is
clinically need. Figure 8.13(A) shows an endoscopic video capture obtained prior to 3D-OCT imaging of
the same patient presented in Section 8.6.1. The dentate line is clearly visible and there are no obvious
signs of edema, ectatic vessels, or large-scale inflammation. The 3D-OCT image in Figure 8.13(B) shows
another buried vessel beneath the lamina propria. In the future, modifications to the study protocol to
allow excisional biopsy following successful RFA treatment would be valuable for confirming the
presence of deeply-buried vascular remnants in these subjects. Alternatively, if the increased risk of
bleeding is deemed to be unacceptably high, animal models could be used to obtain histological correlates
of pre- and post-RFA RP sites. In this brief pilot study, 3D-OCT has shown great potential for
understanding the architectural morphology associated with RP and for evaluating the mechanism of
action of novel RFA therapy.
170
8.7 Figures
Figure 8.1. 3D-OCT images of normal columnar epithelial mucosa in the colon. A: Volume rendering
showing location of orthogonal cut planes. B: En face XZ image of epithelium. Dashed box indicates
region of interest shown in Figure 9.2. C: YZ Cross-section. D: XY cross-section. Coloured arrows in B-D
indicate locations of other orthogonal cut planes.
Figure 8.2. Enlarged views of normal colonic crypts in the region of interest indicated in Figure 9.1(B). A:
En face image formed by projecting volumetric dataset over a 480 um range. Crypt lumens are clearly
visible due to increased light transmission. B,C: Cross-sectional images showing typical columnar
structure of crypts with increased transmission (C). Red arrows indicate projection region for forming en
face image in A.
171
4-
EP P-
4-MM --SM
-4
of normal glandular mucosa in the
Figure 8.3. Comparison of 3D-OCT data to conventional examination
histology slide from excisional biopsy
colon. A: White light video endoscopy image. B: Representative
Crypts (C) are visible but deeper
specimen. C: Cross-sectional OCT image extracted from 3D dataset.
layers cannot be distinguished.
location of orthogonal cut
Figure 8.4. 3D-OCT images of ulcerative colitis. A: Volume rendering showing
of interest shown in
regions
indicates
boxes
planes. B: En face XZ image at 300 um tissue depth. Dashed
(UC) and
Ulcerative
Cross-section.
YZ
C:
apparent.
Figure 3.5. Irregular vascular and cystic features are
and
vessels
surface
showing
cross-section
XZ
D:
normal squamous (NS) regions can be distinguished.
planes.
cut
orthogonal
other
of
locations
edema. Coloured arrows in B-D indicate
172
of interest indicated in Figure 9.4(B).
Figure 8.5. Enlarged views of ulcerative colitis region in first region
vessels and edema. B,C: Cross
A: En face image showing irregular subsurface structure including
sectional images showing region of ulcerative colitis.
of interest indicated in
Figure 8.6. Enlarged views of squamous epithelial mucosa in second region
mucosa. B,C: Cross
Figure 9.4(B). A: En face image showing regular, unremarkable squamous epithelial
possibly due to
sectional images showing regular layered structure with subtle horizontal striations,
healing.
173
A: White light
Figure 8.7. Comparison of 3D-OCT data to conventional examination of ulcerative colitis.
C: Crossspecimen.
biopsy
excisional
from
slide
histology
video endoscopy image. B: Representative
architecture
layered
of
loss
and
(ED)
regions
Edematous
dataset.
3D
from
sectional OCT image extracted
are apparent.
Figure 8.8. 3D-OCT images of region containing edematous vessels due to radiation proctitis. A: Volume
depth.
rendering showing location of orthogonal cut planes. B: En face XZ image at 250 um tissue
located
vessels
Large
Dashed box indicates region of interest shown in Figure 9.9. C: YZ Cross-section.
100 - 150 um below the tissue surface can be visualized (triangles). D: XZ cross-section showing multiple
superficial vessels.
174
in Figure 9.8(B). A: En
Figure 8.9. Enlarged views of ectatic vessels in the region of interest indicated
showing detail of
images
face image showing large edematous vessel (triangles). B,C: Cross-sectional
superficial vessels with thin, hyperintense tissue cap.
from radiation
Figure 8.10. Comparison of 3D-OCT data to conventional examination of ectatic vessels
specimen. C:
biopsy
proctitis. A: White light video endoscopy image. B: histology slide from excisional
are visible in
(triangles)
vessels
Cross-sectional OCT image extracted from 3D dataset. Large superficial
all three images.
175
I
NC
II
NS
location of
Figure 8.11. 3D-OCT images of region treated with RFA for RP. A: Volume rendering showing
region of
indicates
box
orthogonal cut planes. B: En face XZ image at 350 um tissue depth. Dashed
are now
(triangles)
interest shown in Figure 3.9. C: YZ Cross-section. Previously edematous vessels
vessels.
covered in a protective neo-squamous layer. D: XZ cross-section showing covered
Figure 8.12. Enlarged views of treated radiation proctitis in the region of interest indicated in Figure
300
3.11(B). A: En face image showing covered vessel (triangles). B,C: Cross-sectional images showing
vessels.
edematous
- 350 um thick protective layer of neo-squamous epithelium over top of previously
176
A: White light video endoscopy
Figure 8.13. Comparison of 3D-OCT data to conventional examination.
OCT image extracted from
image. B: histology slide from excisional biopsy specimen. C: Cross-sectional squamous epithelium.
appearing
3D dataset. Buried vessels (triangle) are visible underneath regular
177
8.8 References
[1]
A. Jemal, R. Siegel, E. Ward, T. Murray, J. Xu, and M. J. Thun, "Cancer statistics, 2007," CA
CancerJ Clin, vol. 57, pp. 43-66, Jan-Feb 2007.
[2]
D. K. Rex, "Maximizing detection of adenomas and cancers during colonoscopy," American
Journalof Gastroenterology,vol. 101, pp. 2866-2877, Dec 2006.
[3]
R. Kiesslich, M. von Bergh, M. Hahn, G. Hermann, and M. Jung, "Chromoendoscopy with
indigocarmine improves the detection of adenomatous and nonadenomatous lesions in the colon,"
Endoscopy, vol. 33, pp. 1001-1006, Dec 2001.
[4]
S. H. Itzkowitz and N. Harpaz, "Diagnosis and management of dysplasia in patients with
inflammatory bowel diseases," Gastroenterology,vol. 126, pp. 1634-1648, May 2004.
[5]
E. V. Loftus, "Clinical epidemiology of inflammatory bowel disease: Incidence, prevalence, and
environmental influences," Gastroenterology,vol. 126, pp. 1504-1517, May 2004.
[6]
J. A. Eaden, K. R. Abrams, and J. F. Mayberry, "The risk of colorectal cancer in ulcerative
colitis: a meta-analysis," Gut, vol. 48, pp. 526-535, Apr 2001.
[7]
D. C. Adler, Y. Chen, R. Huber, J. Schmitt, J. Connolly, and J. G. Fujimoto, "Three-dimensional
endomicroscopy using optical coherence tomography," Nature Photonics,vol. 1, pp. 709-716,
Dec 2007.
[8]
P. L. Hsiung, L. Pantanowitz, A. D. Aguirre, Y. Chen, D. Phatak, T. H. Ko, S. Bourquin, S. J.
Schnitt, S. Raza, J. L. Connolly, H. Mashimo, and J. G. Fujimoto, "Ultrahigh-resolution and 3dimensional optical coherence tomography ex vivo imaging of the large and small intestines,"
GastrointestEndosc, vol. 62, pp. 561-574, Oct 2005.
[9]
T. Takayama, S. Katsuki, Y. Takahashi, M. Ohi, S. Nojiri, S. Sakamaki, J. Kato, K. Kogawa, H.
Miyake, and Y. Niitsu, "Aberrant crypt foci of the colon as precursors of adenoma and cancer,"
New EnglandJournalof Medicine, vol. 339, pp. 1277-1284, Oct 29 1998.
[10]
X. Qi, Y. S. Pan, Z. L. Hu, M. V. Sivak, J. Willis, K. Olowe, and A. M. Rollins, "Morphological
feature quantification of colonic crypt patterns using microscope-integrated OCT,"
Gastroenterology,vol. 134, pp. A577-A577, Apr 2008.
178
CHAPTER 9
9.0
Conclusions, Future Work, and Publications
179
9.1
Summary of Thesis Work
This thesis project involved a combination of technology development, pre-clinical imaging
experiments, and human clinical studies in the field of three-dimensional optical coherence tomography
(3D-OCT). First, fundamental advances were made in high-speed wavelength swept light sources to
enable the next generation of high-speed 3D-OCT platforms. The Fourier domain mode locked (FDML)
laser was proven to be an ideal light source for 3D-OCT due to its unprecedented combination of high
sweep rate, wide tuning range, high output power, and low phase noise properties. FDML laser
technology was then integrated into a number of 3D-OCT imaging systems. One system was used to
demonstrate OCT phase profilometry with sub-nanometer axial resolution at record imaging speeds. This
technology could be used in the future for studies of cellular dynamics, industrial inspection of MEMS or
microfluidic devices, or other applications requiring measurement of extremely small and extremely rapid
transient events.
Using a similar phase-sensitive OCT system and FDML laser, a new method was developed to detect
gold nanoparticle contrast agents with 1 to 2 orders of magnitude higher signal-to-noise ratios than
previously reported OCT methods. This new method used photothermal modulation of the sample and
frequency resolved signal detection to reject noise in a manner similar to lock-in detection. This technique
could enable molecularly sensitive 3D-OCT imaging of tissue pathologic state by bioconjugation of
molecular probes to the nanoshells. This advance would, if successful, fundamentally transform the OCT
field by enabling the simultaneous analysis of 3D tissue structure and function in vivo and in real time.
A state-of-the-art 3D-OCT endomicroscopy system was constructed in collaboration with LightLab
Imaging Inc. in order to perform 3D-OCT imaging of the human gastrointestinal (GI) tract. This system
combined an FDML laser with purpose-built data acquisition hardware, signal processing algorithms, and
real-time image display. Specialized spiral-scanning imaging probes were developed to provide an
optimized
combination of flexibility,
strength, rotational and pullback uniformity,
and optical
performance. This system was used in a series of clinical pilot studies including the first demonstration of
3D-OCT imaging in the human colon. New GI applications of 3D-OCT technology focusing on the use of
high-speed high-resolution imaging as an adjunct to endoscopic therapy were developed. These studies
demonstrated the utility of 3D-OCT for pre- and post-treatment analysis of radiofrequency ablation for
both Barrett's esophagus and radiation proctitis. In addition, 3D-OCT was shown to be useful for
assessment of excision margins following endoscopic mucosal resection of dysplastic nodules in the
esophagus.
These studies have opened the door for clinical acceptance of new indications for 3D-OCT, focusing
on the targeted imaging of known pathologic regions for treatment planning and follow-up assessment.
180
The core technology developed in this thesis work can also be applied to the more traditional OCT niches
of screening and initial detection or diagnosis of pathology. As endoscopic therapies continue to gain in
sophistication, capability, and popularity within the gastroenterology community, 3D-OCT can fill an
unmet need for providing on-demand, real time 3D visualization of subsurface tissue microstructure both
pre- and post-treatment.
9.2
Future Work
Future efforts for the projects described in this thesis work should focus on four main areas: 1) In vivo
detection of molecularly targeted nanoshells; 2) 3D-OCT imaging engine development; 3) advanced
imaging probe design, and; 4) long-term clinical studies.
First, the fundamental advances in nanoshell detection described in this thesis work must be translated
into successful in vivo imaging experiments. The nanoshells should be bioconjugated to antibodies
against proteins overexpressed in common human pathologies. One good example would be an antibody
targeting epidermal growth factor receptor (EGFR), which is associated with a number of epithelial
cancers including colorectal cancer. These targeted nanoshells could be tested on in vitro cell cultures as
well as animal models of colorectal cancer to demonstrate direct detection of pathologic state using 3DOCT.
Second, 3D-OCT imaging engine development should integrate the newest advances in high-speed
data acquisition technology, hardware-based signal processing, and efficient image display and storage
methods to scale up imaging speeds to the rate supported by state-of-the-art FDML laser designs. The
first generation 3D-OCT endomicroscopy system developed here was capable at operating at 60 - 100
kHz but could only be run at 60 kHz to ensure optimal image quality for human clinical studies. With
currently developed FDML lasers already operating at sweep rates of 370 kHz and with future designs
likely to target speeds of 500 kHz or more, the imaging engine must be updated to keep pace with laser
technology developments. New data acquisition hardware operating at 400 MSamples/s with onboard
digital signal processing chips, combined with recently released multicore personal computer central
processing units, could provide a viable pathway for this effort. These increased imaging speeds will
enable corresponding increases in spatial sampling density and field of view, providing clinicians with an
unsurpassed convergence of true 3D resolution, depth-resolved imaging, and rapid acquisition speeds.
Third, new imaging probes must be developed to provide further improvements in image quality and
reliability to enable long-term clinical studies of 3D-OCT endomicroscopy. Despite the advances in probe
technology described in this thesis work, image quality can still be further improved by reducing
nonuniformities in push/pull actuation. This problem may be addressed by further miniaturization of the
current design. Entirely new probe concepts, such as forward-imaging raster scan probes, could also be
181
developed to provide extremely rapid acquisition over a smaller field of view. Probe designs must also be
scaled to the higher scanning speeds envisioned for the next generation of 3D-OCT imaging engines.
Finally, long-term clinical studies must build on the initial results demonstrated in this thesis work to
decisively convince the clinical community of the utility of 3D-OCT endomicroscopy. Studies of 3DOCT for pre- and post-therapy analysis can be conducted with long term follow-up of patients treated
with radiofrequency ablation, endoscopic mucosal resection, cryospray therapy, photodynamic therapy, or
other emerging endoscopic treatments. 3D-OCT could also be studied as an intra-therapy tool for guiding
the dosage or resection depth of theses techniques. Studies in 3D-OCT screening for dysplasia in the
setting of Barrett's esophagus or inflammatory bowel disease could also be revisited as fields of view
become larger and resolutions become higher. Successful clinical studies are the only way for 3D-OCT to
become an accepted clinical tool outside of the specialty research field, and are critical for long-term
success of the technology.
9.3
Publications Produced During Thesis Work
Publications [1-15] were produced between 2006 - 2009 during work on the PhD thesis. Publications [1619] were produced between 2002 - 2004 during work on the SM thesis.
[1]
D. C. Adler, C. Zhou, T.-H. Tsai, H.-C. Lee, L. Becker, J. M. Schmitt, J. G. Fujimoto, and H.
Mashimo, "Three-dimensional optical coherence tomography examination of non-ablated
Barrett's
esophagus
and
buried
glands
beneath
neo-squamous
epithelium
following
radiofrequency ablation," Endoscopy, In Preparation, 2009.
[2]
A. D. Aguirre, D. C. Adler, and J. G. Fujimoto, "Practical Guidelines for Biological Imaging with
Optical Coherence Tomography," Nature Protocols, In Preparation, 2009.
[3]
C. Zhou, D. C. Adler, L. Becker, Y. Chen, T.-H. Tsai, M. Figueiredo, J. Schmitt, J. G. Fujimoto,
and H. Mashimo, "Effective treatment of chronic radiation proctitis using radiofrequency
ablation," Therapeutic advances in gastroenterology,In Press, 2009.
[4]
D. C. Adler, C. Zhao, T.-H. Tsai, J. Schmitt, Q. Huang, H. Mashimo, and J. G. Fujimoto, "Threedimensional endomicroscopy of the human colon using optical coherence tomography," Optics
Express, vol. 17, pp. 784-796, January 2009 2009.
[5]
B. R. Biedermann, W. Wieser, C. M. Eigenwillig, G. Palte, D. C. Adler, V. J. Srinivasan, J. G.
Fujimoto, and R. Huber, "Real time en face Fourier-domain optical coherence tomography with
direct hardware frequency demodulation," Optics Letters, vol. 33, pp. 2556-2558, Nov 1 2008.
182
[6]
V. J. Srinivasan, D. C. Adler, Y. Chen, I. Gorczynska, R. Huber, J. Duker, J. S. Schuman, and J.
G. Fujimoto, "Ultrahigh-speed Optical Coherence Tomography for Three-Dimensional and En
Face Imaging of the Retina and Optic Nerve Head," IOVS, 2008, pp. 08-2127.
[7]
P. M. Andrews, Y. Chen, M. L. Onozato, S. W. Huang, D. C. Adler, R. A. Huber, J. Jiang, S. E.
Barry, A. E.Cable, and J. G. Fujimoto, "High-resolution optical coherence tomography imaging
of the living kidney," LaboratoryInvestigation, vol. 88, pp. 441-449, Apr 2008.
[8]
D. C. Adler, S. W. Huang, R. Huber, and J. G. Fujimoto, "Photothermal detection of gold
nanoparticles using phase-sensitive optical coherence tomography," Optics Express, vol. 16, pp.
4376-4393, Mar 31 2008.
[9]
D. C. Adler, Y. Chen, R. Huber, J. Schmitt, J. Connolly, and J. G. Fujimoto, "Three-dimensional
endomicroscopy using optical coherence tomography," Nature Photonics, vol. 1, pp. 709-716,
Dec 2007.
[10]
D. C. Adler, J. Stenger, I. Gorczynska, H. Lie, T. Hensick, R. Spronk, S. Wolohojian, N.
Khandekar, J. Y. Jiang, S. Barry, A. E.Cable, R. Huber, and J. G. Fujimoto, "Comparison of
three-dimensional optical coherence tomography and high resolution photography for art
conservation studies," Optics Express, vol. 15, pp. 15972-15986, Nov 26 2007.
[11]
R. Huber, D. C. Adler, V. J. Srinivasan, and J. G. Fujimoto, "Fourier domain mode locking at
1050 nm for ultra-high-speed optical coherence tomography of the human retina at 236,000 axial
scans per second," Optics Letters, vol. 32, pp. 2049-2051, Jul 2007.
[12]
S.W. Huang, A. D. Aguirre, R. A. Huber, D. C. Adler, and J. G. Fujimoto, "Swept source optical
coherence microscopy using a Fourier domain mode-locked laser," Optics Express, vol. 15, pp.
6210-6217, May 2007.
[13]
M. W. Jenkins, D. C. Adler, M. Gargesha, R. Huber, F. Rothenberg, J. Belding, M. Watanabe, D.
L. Wilson, J. G. Fujimoto, and A. M. Rollins, "Ultrahigh-speed optical coherence tomography
imaging and visualization of the embryonic avian heart using a buffered Fourier Domain Mode
Locked laser," Optics Express, vol. 15, pp. 6251-6267, May 2007.
[14]
D. C. Adler, R. Huber, and J. G. Fujimoto, "Phase-sensitive optical coherence tomography at up
to 370,000 lines per second using buffered Fourier domain mode-locked lasers," Optics Letters,
vol. 32, pp. 626-628, Mar 2007.
[15]
R. Huber, D. C. Adler, and J. G. Fujimoto, "Buffered Fourier domain mode locking:
unidirectional swept laser sources for optical coherence tomography imaging at 370,000 lines/s,"
Optics Letters, vol. 31, pp. 2975-2977, Oct 2006.
[16]
D. S. Adler, T. H. Ko, A. K. Konorev, D. S. Mamedov, V. V. Prokhorov, J. J. Fujimoto, and S. D.
Yakubovich, "Broadband light source based on quantum-well superluminescent diodes for high-
183
resolution optical coherence tomography," Kvantovaya Elektronika, Moskva, vol. 34, pp. 915-8,
2004/10/ 2004.
[17]
D. C. Adler, T. H. Ko, and J. G. Fujimoto, "Speckle reduction in optical coherence tomography
images by use of a spatially adaptive wavelet filter," Opt Lett, vol. 29, pp. 2878-80, Dec 15 2004.
[18]
D. C. Adler, T. H. Ko, P. R. Herz, and J. G. Fujimoto, "Optical coherence tomography contrast
enhancement using spectroscopic analysis with spectral autocorrelation," Optics Express, vol. 12,
pp. 5487-5501, Nov 1 2004.
[19]
T. H. Ko, D. C. Adler, J. G. Fujimoto, D. Mamedov, V. Prokhorov, V. Shidlovski, and S.
Yakubovich, "Ultrahigh resolution optical coherence tomography imaging with a broadband
superluminescent diode light source," Optics Express, vol. 12, pp. 2112-2119, MAY 17 2004.
184
Canadian Born
We first saw light in Canada, the land beloved of God;
We are the pulse of Canada, its marrow and its blood:
And we, the men of Canada, can face the world and brag
That we were born in Canada beneath the British flag.
Few of us have the blood of kings, few are of courtly birth,
But few are vagabonds or rogues of doubtful name and worth;
And all have one credential that entitles us to bragThat we were born in Canada beneath the British flag.
We've yet to make our money, we've yet to make our fame,
But we have gold and glory in our clean colonial name;
And every man's a millionaire if only he can brag
That he was born in Canada beneath the British flag.
No title and no coronet is half so proudly worn
As that which we inherited as men Canadian born.
We count no man so noble as the one who makes the brag
That he was born in Canada beneath the British flag.
The Dutch may have their Holland, the Spaniard have his Spain,
The Yankee to the south of us must south of us remain;
For not a man dare lift a hand against the men who brag
That they were born in Canada beneath the British flag.
-- Emily Pauline Johnson
1861-1913
185
Download