Download Biomedical Imaging and Applied Optics Laboratory

Document related concepts

Atmospheric optics wikipedia , lookup

Microscopy wikipedia , lookup

Silicon photonics wikipedia , lookup

Optical amplifier wikipedia , lookup

Confocal microscopy wikipedia , lookup

Two-dimensional nuclear magnetic resonance spectroscopy wikipedia , lookup

Fourier optics wikipedia , lookup

Nonimaging optics wikipedia , lookup

Preclinical imaging wikipedia , lookup

Super-resolution microscopy wikipedia , lookup

Spectrum analyzer wikipedia , lookup

Hyperspectral imaging wikipedia , lookup

Phase-contrast X-ray imaging wikipedia , lookup

Retroreflector wikipedia , lookup

Imagery analysis wikipedia , lookup

Astronomical spectroscopy wikipedia , lookup

Ellipsometry wikipedia , lookup

Ultrafast laser spectroscopy wikipedia , lookup

Optical rogue waves wikipedia , lookup

Vibrational analysis with scanning probe microscopy wikipedia , lookup

Optical tweezers wikipedia , lookup

X-ray fluorescence wikipedia , lookup

Nonlinear optics wikipedia , lookup

Photon scanning microscopy wikipedia , lookup

Magnetic circular dichroism wikipedia , lookup

3D optical data storage wikipedia , lookup

Harold Hopkins (physicist) wikipedia , lookup

Rutherford backscattering spectrometry wikipedia , lookup

Spectral density wikipedia , lookup

Ultraviolet–visible spectroscopy wikipedia , lookup

Chemical imaging wikipedia , lookup

Optical coherence tomography wikipedia , lookup

Transcript
Biomedical Imaging and
Applied Optics Laboratory
Department of Electrical and Computer Engineering - University of Cyprus
Spectral Analysis of Optical Coherence
Tomography Signals
by
Andreas Kartakoullis
Submitted to the University of Cyprus in partial fulfillment
of the requirements for the degree of
Master of Science in Electrical Engineering
Department of Electrical and Computer Engineering
May 2008
SPECTRAL ANALYSIS OF
OPTICAL COHERENCE TOMOGRAPHY SIGNALS
By
Andreas Kartakoullis
Submitted to the University of Cyprus in partial fulfillment
of the requirements for the degree of Master of Science in Electrical Engineering
Department of Electrical and Computer Engineering
May 2008
SPECTRAL ANALYSIS OF
OPTICAL COHERENCE TOMOGRAPHY SIGNALS
By
Andreas Kartakoullis
Examination committee:
Dr. Constantinos Pitris
Assistant Professor, Department of ECE, Research Supervisor
Dr. Stavros Iezekiel
Associate Professor, Department of ECE, Committee Member
Dr. Michalis Averkiou
Assistant Professor, Department of MME, Committee Member
Abstract
More than 85% of all cancers originate in the epithelium that lines the internal
surfaces of organs throughout the body. Before they become invasive, at stages
known as dysplasia and carcinoma in situ, early cancer cells alter the epithelial-cell
architecture. More specifically, the number of cells, and therefore the number of
nuclei, increases. The nuclei themselves become enlarged and hyperchromatic, i.e.
their chromatin changes and colours more darkly when stained during histopathology. Nuclear sizes change from the normal 5 − 10 µm to 1.5 − −2 x the size, i.e.
10 − 20 µm. Currently, these early changes are only detectable by histopathology
or, non-invasively, by optical imaging techniques such as confocal or multi-photon
microscopy. Unfortunately, neither of the two techniques has been clinically implemented due their complexity and their limited penetration in tissue.
Optical Coherence Tomography (OCT) is a prominent non-invasive biomedical tissue imaging technique that generates in vivo cross-sectional images of tissue microstructure with micron-scale spatial resolution. OCT can overcome the
limitations of the other techniques; but the dysplastic alterations are not clearly
discernible even for the resolution of OCT. However, it is well known that changes
in scatterer size induce changes in the spectral content of scattered light. These
changes have been found to be diagnostically useful in scattering spectroscopy
in the visible range. Cellular organelles in epithelial tissue can be modelled as
spheroidal scatterers whose interactions with light are governed by Mie theory. Mie
theory provides an analytical closed-form description of optical scattering from single spheroidal particles as a function of particle size, refractive index, wavelength,
observation angle, and optical polarization. This type of analysis can be applied to
OCT to enable the detection of the early changes of dysplasia. Spectroscopic OCT
(SOCT) is the spectroscopic extension of OCT and it can enhance the contrast of
OCT systems by differentiating tissue based on properties other than the intensity of the backscattered light, such as different spectral features of the detected
spectrum at each pixel in the image.
Currently two approaches for sizing the scatterers based on the measured spectra, the first approach is based on curve fitting the normalized experimental measurement to the theoretical Mie prediction. The second approach is based on pitch
detection such as using the Fourier transform or determining the autocorrelation.
We have exploited the stochastic nature of the OCT signal due to speckle, since
multiple scatterers coexists in the OCT imaging volume causing the backscattered
fields to interfere coherently with one another in a stochastic process. The detected
spectrum therefore is an outcome of that stochastic process, where the statistical
properties are determined by the scatterers size and concentration. By analyzing
the statistical properties of the backscattered spectra, information about the underii
lying distribution of scattering particles can be extracted and speckle can be used
as a diagnostic tool.
This thesis introduces a novel method for scatterer size estimation and classification using SOCT based on autoregressive spectral estimation techniques and
statistical analysis. Principal Component Analysis (PCA) based algorithm was
used to extract diagnostic features and to reduce the dimension of the spectra. The
extracted features of each sample are then input into a classification process using Multivariate Analysis of Variance (MANOVA), and to give solution to the size
equations. The results are very encouraging and indicate that the spectral content
of OCT signals can be used to estimate scatterer size and classified with high accuracy. This technique can result in an extremely valuable tool for the investigation
of disease tissue features which now remain below the resolution of OCT.
iii
PerÐlhyh
Ta
j lio
perissìtera
pou
eÐdh
kalÔptei
tic
karkÐnou,
perissìtero
eswterÐkec
epifˆniec
apo
twn
85%
,
proèrqontai
orgˆnwn
se
ìlo
apì
to
to
epi-
s¸ma.
Ta
prwÐma karkÐnika stˆdia, eÐnai gn¸sta wc dusplˆsia kai karkÐnwma in-situ, ìpou sta
stˆdia
autˆ
sunteloÔntai
sugkekrimèna,
oi
pur nec.
pur na
ta
qarakthrÐs tikec
epijhlÐaka
EpÐshc
oi
alloиnetai
kÔttara
pur nec
apì
alloÐwseic
auxˆnontai
diogkìnwntai
kai
me
sta
epijhliˆka
apotèlesma
na
10 − 20 µm
se
.
Pio
auxˆnontai
uperqrwmatÐzontai,
5 − 10 µm 1.5 − −2 x
kÔttara.
to
mègejoc
Mèqri
kai
tou
s tigm c
h
diˆgnwsh twn neoplasi¸n basÐzetai se is tologikèc exetˆseic twn is t¸n apì bioyÐec
kai
ektomèc.
Oi
monèc
mh-epembatikèc
teqnikèc
pou
èqoun
aut thn
dunatìthta
di-
ˆgnwshc den èqoun brei klinik efarmog lìgw problhmˆtwn me thn axiopis tÐa touc
kai me
H
Ð
na
thn praktik efarmog touc.
Optik SÔmfwnh
xeperˆsei
autèc
(∼ 2 − 3 mm)
kai
TomografÐa
tic
(Optical
adunamÐec.
eukrÐneia
thc
çherence
MporeÐ
tˆxewc
na
Tomography
petÔqei
OCT),
ikanopoihtik (1 − 15 µm)
mikromètrwn
,
mpore-
dieÐsdush
ìmwc
oi
du-
splas tikèc alloi¸seic den eÐnai pl rwc eudiˆkritec akìma kai me thn eukrÐneia aut .
Oi
dusplas tikèc
toskopik¸n
allagèc
(Light
teqnik¸n,
sto
Ektetamènh
alloi¸seic
mègejoc
ereunˆ
Scattering
gnws tikˆ
tìte
afoÔ
tou
èqei
ìpwc
gÐnei
Ean
to
ta
ìmwc
eÐnai
skedas t sthn
Spectroscopy
qr simh.
skedas tèc,
mporoÔn
na
diagnws toÔn
gnws tì
apì
dhmiourgoÔn
perioq LSS)
epijhliakˆ
thc
pou
opisjoskedasmèno
fˆsma
thn
sto
aut h
Optik c
mia
jewrÐa,
san
sunis t¸sa
fˆsma.
Skèdashc
prosèggish
montelopoihjoÔn
perièqei
fasma-
skedasmèno
FasmatoskopÐac
ìti
qr sh
hlektromagnhtik metabolèc
èdeixe
kÔttara
thn
me
eÐnai
dia-
sfairikoÐ
Mie
pou
exartˆtai
apì to mègejoc, thn sugkèntrwsh kai to deÐkth diˆjlashc twn skedas t¸n.
Ta qara-
kthris tikˆ autˆ mporeÐ na prosdioris toÔn me duo trìpouc, me thn qr sh thc jewrÐac
Mie opoÔ ta peiramatikˆ apotelèsmata prosarmìzontai sta jewrhtikˆ [5].
O deÔte-
roc trìpoc eÐnai me ton entopismì apotìmwn ektinˆxewn, panw apì èna kat¸fli, thc
autosusqètishc
kurÐwc
thc
tou
qr sewc
fˆsmatoc.
fak¸n
me
'Omwc
qamhlì
h
LSS
èqei
ˆnoigma
arketèc
èmfutec
diafrˆgmatoc
(NA),
adunamÐec
pou
èqei
lìgw
san
a-
potèlesma qamhl egkˆrsia eukrÐneia kai periorismènh dieÐsdush lìgw exasjènhshc.
An kai èqei prosarmos teÐ sthn LSS sumbolometrÐa qamhl c sumfwnÐac gia skopoÔc
beltÐwshc thc
'Ena
fwc
gia
rioqèc
anˆlushc den
sumbatikì
na
me
sÔs thma
dhmiourg sei
parìmoia
eÐnai teqnik OCT
qrhsimopoieÐ
dusdiˆs tatec
èntash
den
apeikìnishc allˆ leitourgik teqnik .
mporoÔn
thn
èntash
eikìnec
diatom c
na
pl rwc
eÐnai
tou
tou
opisjoskedasmènou
deÐgmatoc.
eudiˆkritec
stic
'Omwc
pe-
sumbatikèc
eikìnec OCT epÐshc swreÐa plhrofori¸n pou perièqetai se ˆllec paramètrouc ektìc
apì thn èntash den axiopoieÐtai.
Mia prìsfath fasmatoskopik proèktash thc OCT
eÐnai
fasmatoskopik h
SOCT
pou
opisjoskedasmèno
epitrèpei
fwc,
se
thn
epÐshc
dusdiˆs tatec
v
exagwg qarakthris tik¸n
apeikonÐseic.
Ston
apì
to
apeikonis tikì
q¸ro tou OCT sunupˆrqoun tautìqrona pollaploÐ skedas tèc pou lìgw thc optik c
sumfwnÐac
s tigm c
san
thn
h
tou
s matoc
kìkkwsh
s toqas tik Autˆ
èntash
mia
tou
OCT
ta
idiìthtec
kai
s matoc
tou
pou mporoÔn
eÐnai
tìso
dhmiourgoÔn
wc
pou
phg eÐnai
katanom s toqas tikˆ
eikìna
OCT
jewreÐtai
diergasÐa
sugkèntrwsh
OCT.
thc
twn
motÐba
kai
apotèlesma
jorÔbou
èqei
skedas t¸n
mèsa
sthn
thc
ston
SOCT
mporoÔme
na qrhsimopoihjoÔn gia upologismì
kìkkwshc.
Mèqri
apeikìnish.
'Omwc
exart¸ntai
apì
to
apeikonis tikì
parousiˆzontai
fˆsma.
s toqas tik c
fˆsmatoc
motÐba
bioiatrik pou
opisjoskedasmèno
aut c
opisjoskedasmènou
sthn
idiìthtec
kìkkwshc
sto
s toqas tikˆ
'Etsi,
diadikasÐac,
na
exˆgoume
mègejoc,
q¸ro
tìso
thc
sthn
jewr¸ntac
ìti
melet¸ntac
tic
qarakthris tikˆ
megèjouc tou skedas t .
AnalÔontac tic s tatis tikèc idiìthtec tou opisjoskedasmènou optikoÔ fˆsmatoc
me
thn
qr sh
plhroforÐec
prohgmènwn
sqetikˆ
me
mejìdwn
thn
trìpo se pragmatikì qrìno.
anˆlushc
morfologÐa
kai
mporoÔme
sÔnjesh
na
tou
anakt soume
is toÔ
me
mh-
polÔtimec
epembatikì
Pio sugkekrimèna qrhsimopoÐhjhke èna autopalindromi-
kì montèlo fasmatoskopik c ektÐmhshc basismèno sthn mèjodo Burg gia thn anˆkth-
sh
tou
opisjoskedasmènou
(Principal
jhke
gia
çmponentc
exagwg kai
thn
upˆrqoun
twn
apotelesmˆtwn
Ta
tou
'Epeita
anˆlush
twn
opisjoskedasmènou
pou
ja
mac
prwteÔwn
fˆsmatoc
odhg soun
ston
sunis tws¸n
pragmatopoÐh-
upologismì
tou
Me thn qr sh PCA petuqaÐnoume thn meÐwsh tou megèjouc twn
epanadom sh
pou
mejìdou.
na
Analysic)
qarakthris tik¸n
megèjouc skedas t .
dedomènwn
fˆsmatoc.
touc
anagnwrÐzontai
katadeÐxe
apotelèsmata
thn
thc
me
arketˆ
tètoio
pio
eÔkola.
euaisjhsÐa
mejìdou
pou
trìpo
kai
thn
¸ste
ta
diaforetikˆ
AkoloÔjwc
eidikìthta
anaptÔqjhke
eÐnai
prìtupa
kathgoriopoÐhsh
thc
proteinomènhc
arkèta
enjaruntikˆ
gia peraitèrw melèth ètsi ¸ste na axiopoÐhjei san èna qr simo ergaleÐo gia diˆgnwsh
pajologÐwn pou paramènoun
kˆtw
apì thn eukrÐneia twn sus t matwn OCT.
vi
Contents
1 Introduction
12
1-1 Optical Coherence Tomography as a Biomedical Imaging Technology 12
1-2 Spectroscopic Optical Coherence Tomography . . . . . . . . . . . . 13
1-3 Scope of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2 Optical Coherence Tomography
2-1 Introduction . . . . . . . . . . . . . . . . . . . .
2-2 Time-Domain OCT (TD-OCT) . . . . . . . . .
2-2.1 Interferometer with Coherent Light . . .
2-2.2 Interferometer with Low Coherence Light
2-2.3 Nondispersive Medium . . . . . . . . . .
2-2.4 OCT Resolution . . . . . . . . . . . . . .
2-2.5 Group Velocity Dispersion . . . . . . . .
2-3 Optical Sources . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3 Mie Theory
3-1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . .
3-2 General Formulation of the Problem . . . . . . . . . . .
3-2.1 Boundary Conditions . . . . . . . . . . . . . . . .
3-2.2 Solutions to the vector wave equations . . . . . .
3-3 Expansion of a plane wave in vector spherical harmonics
3-4 The internal and scattered fields . . . . . . . . . . . . . .
3-4.1 Angle-Dependent Functions . . . . . . . . . . . .
3-4.2 Scattering Coefficients . . . . . . . . . . . . . . .
3-5 Scattering Matrix . . . . . . . . . . . . . . . . . . . . . .
3-6 Computation of Scattering Coefficients . . . . . . . . . .
3-7 Mie Theory for Biomedical Optics . . . . . . . . . . . . .
4 Spectroscopic OCT
4-1 Introduction . . . . . . . . . . . . . . .
4-2 Mie theory in SOCT . . . . . . . . . .
4-3 Speckle . . . . . . . . . . . . . . . . . .
4-4 SOCT Instrumentation . . . . . . . . .
4-4.1 Superluminescent diodes (SLDs)
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
15
15
16
16
19
20
21
22
23
.
.
.
.
.
.
.
.
.
.
.
27
27
27
29
30
36
37
39
40
42
44
45
.
.
.
.
.
46
46
47
47
49
49
CONTENTS
8
4-5 Tissue Phantoms . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5 Spectral Analysis of OCT Signals
5-1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
5-2 Spectral Analysis Based on Fourier Transformation .
5-2.1 Energy Spectrum Density (ESD) . . . . . . .
5-2.2 Power Spectrum Density (PSD) . . . . . . . .
5-2.3 Computation using the FFT . . . . . . . . . .
5-3 Spectral Analysis Based on Burg’s Method . . . . . .
5-3.1 Rational Power Spectra . . . . . . . . . . . .
5-3.2 Burg’s Method . . . . . . . . . . . . . . . . .
5-4 Statistical Multivariate Analysis . . . . . . . . . . . .
5-4.1 Principal Component Analysis (PCA) . . . . .
5-4.2 Multivariate Analysis of Variance (MANOVA)
5-4.3 Classification . . . . . . . . . . . . . . . . . .
50
.
.
.
.
.
.
.
.
.
.
.
.
53
53
54
54
55
56
57
57
58
60
60
62
62
.
.
.
.
63
63
63
66
71
7 Summary and Future Work
7-1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7-2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
75
75
75
Bibliography
77
6 Results
6-1 Introduction . . . .
6-2 Spectra Calculation
6-3 Spectral Analysis .
6-4 Depth Analysis . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
List of Figures
1.1
Growth of OCT scientific and medical research publications. The
graph shows the number of publications per year with optical coherence tomography in the title as indexed in the PubMed and
Citation databases [8]. . . . . . . . . . . . . . . . . . . . . . . . .
13
Schematic of Time Domain OCT system with a moving mirror in
the reference arm. . . . . . . . . . . . . . . . . . . . . . . . . . . .
16
2.2
The Michelson interferometer. . . . . . . . . . . . . . . . . . . . .
17
2.3
Interferometric pattern resulting from moving the reference arm
and using a coherent source. Interferometric pattern resulting from
moving the reference arm and using a low coherence source. . . . .
19
Spectral absorption of some tissue molecules, and of aortic tissue.
The brightened wavelength ranges have already been used in OCT
[30]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
24
The incident field (Ei ,Hi ) gives rise to a field (E1 ,H1 ) inside the
particle and a scattered field (Es ,Hs ) in the medium surrounding
the particle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
28
3.2
Closed surface separating regions 1 and 2.
. . . . . . . . . . . . .
30
3.3
Scattering by a spherical particle . . . . . . . . . . . . . . . . . .
32
3.4
Polar plots of the first five angle-dependent functions πn (a) and τn
(b) (n = 1, . . . , 5). All functions are plotted to the same scale. . .
41
Example of back-scattered intensity. (a) Back-scattered intensity
for 1 µm scatterer. (b) Back-scattered intensity for 2 µm scatterer.
(c) Back-scattered intensity for 4 µm scatterer. . . . . . . . . . . .
48
4.2
Schematic of the Time Domain OCT system that we used. . . . .
50
4.3
OCT image of a piece of the phantom placed on an unpolished
metal substrate. The two vertical bars represent z, z 0 , from top to
bottom. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
52
2.1
2.4
3.1
4.1
6.1
The autocorrelation of the backscattering spectra obtained with FFT 65
LIST OF FIGURES
6.2
6.3
6.4
6.5
10
Back-scattered normalized average power spectra obtained with
Burg’s method . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
65
Plots of mean error and sensitivity versus N Burg (a), (c) and principal components (b), (d) for Burg’s method . . . . . . . . . . . .
69
Plots of mean error and sensitivity versus N F F T (a), (c) and principal components (b), (d) for FFT/Autocorrelation . . . . . . . . .
69
Results of the MANOVA analysis of 100 training and 100 sample
spectra for the single power (a), the single power and the derivative (b), the single and square power (c). (The dark lines are the
classification area borders resulting from the discriminant analysis)
70
6.6
Spectra from different depths with the focus at the top of the sample 72
6.7
Spectra from different depths with the focus at the middle of the
sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
72
Spectra from different depths with the focus at the bottom of the
sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
73
6.9
Spectra from below the surface at different pathlengths . . . . . .
73
6.10
Spectra from the same depth from three different OCT images . .
74
6.11
In focus spectra from three different OCT images . . . . . . . . .
74
6.8
List of Tables
2.1
Short coherent sources suitable for use in OCT systems [34]
. . . .
25
4.1
Recipes for tissue phantoms (acrylamide gel) of different imaging
volume concentration . . . . . . . . . . . . . . . . . . . . . . . . .
51
6.1
6.2
6.3
6.4
6.5
6.6
Results for Burg’s Method with window 3001 x 25 pixels and p = 120
AR parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Results for FFT/Autocorrelation with window 3001 x 25 pixels and
N F F T = 212 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Results for square power of the spectrum P̂Bu
. . . . . . . . . . . .
tr
Bu
ˆ tr . . . . . . . . . . . . .
Results for derivative of the spectrum dP
. . . . . . . . . . . .
Results for single power and derivative of P̂Bu
tr
. . . . . . . . . .
Results for single power and square power of P̂Bu
tr
68
68
68
68
68
68
Chapter 1
Introduction
1-1
Optical Coherence Tomography as a Biomedical Imaging Technology
In the last 40 years, many tomographic imaging modalities have been developed
such as Computed Tomography (CT), Magnetic Resonance Imaging (MRI), Ultrasound (US), and Positron Emission Tomography (PET). Optical coherence tomography (OCT) is a relatively recent imaging technology for producing high-resolution
cross-sectional images [1]. OCT plays an important role in biomedical imaging due
to its micrometer resolution and millimeter penetration depth. If different biomedical imaging technologies are classified by their applicable size scales, which vary
from whole body imaging, organ imaging, tissue imaging, and cellular imaging down
to molecular imaging, OCT currently represents one the finest resolutions among
in vivo tomographic imaging modalities.
Optical Coherence Tomography (OCT) is analogous to ultrasound imaging but
is based on the detection of infrared light waves, instead of sound, back-scattered
(reflected) from different layers and structures within the tissue. Typically, a fiber
optic Michelson interferometer is used to measure the relative optical path difference
between a reference arm and the various layers of the tissue sample in the sample
arm. By mechanically varying the path length in the reference arm with a scanning
galvanometer an interference fringe is obtained for an arbitrary path length in the
sample. To generate a single Axial scan (A-scan) in the image, an optical beam
is incident on the sample at a single transverse location while the reference arm is
scanned.
The field of OCT has evolved rapidly in the past years years (Figure 1-1 encompassing many clinical applications. A major step in clinical implementation was
the achievement of high speed imaging capabilities by using Fourier domain detection techniques, such as Fourier-Domain OCT (FD-OCT) and Swept-Source OCT
1-2 Spectroscopic Optical Coherence Tomography
13
(SS-OCT). OCT has made its most significant clinical contribution in the field of
ophthalmology, where it has become a key diagnostic technology in the areas of retinal diseases and glaucoma [2]. Also, OCT has been extensively applied in the field
of cardiology, to examine the structural integrity of the vasculature in the coronary
arteries [3, 4]. More recently it has been applied for cellular and molecular analysis
using modalities such as Spectroscopic OCT [5, 6] and Polarization-Sensitive OCT
(PS-OCT) [7]. When fully exploited, OCT has the potential to dramatically change
the way physicians, researchers and scientists see and understand human tissues in
order to better diagnose and treat disease.
Figure 1.1: Growth of OCT scientific and medical research publications. The graph
shows the number of publications per year with optical coherence tomography in
the title as indexed in the PubMed and Citation databases [8].
1-2
Spectroscopic Optical Coherence Tomography
An extension of OCT is Spectroscopic OCT (SOCT) for performing both crosssectional tomographic imaging of the structure and imaging of the spectral content.
SOCT is able to perform depth-resolved spectroscopy, offering the possibility of reconstructing 3-d spectral map of a sample. The first time-domain broadband SOCT
technique used the spectral-centroid shift as the indicator of spectral modification,
to detected optical absorption by melanin in an African frog tadpole [9]. Later,
depth-resolved backscattered spectra and tissue transfer functions were measured
1-3 Scope of the Thesis
14
with high precision using a Fourier-Domain OCT system, which enabled quantitative estimation of absorber concentration [10]. SOCT has the potential to become a
powerful extension of OCT with applications such as the assessment of blood oxygen saturation [6]. One promising application is scatterer size and concentration
measurement for the detection of early dysplastic changes using Mie theory [11], or
angle-resolved low-coherence interferometry (aLCI), a related technique [12].
1-3
Scope of the Thesis
The goal of this thesis is to introduce a novel method for spectral analysis of OCT
signals which can be used for scatterer size estimation and classification. Chapter 2 is a review of the fundamental operating principles and theory behind OCT.
In Chapter 3, Mie scattering theory is reviewed to understand the basis for using
spectral modulation as a diagnostic feature in SOCT. Chapter 4 discusses basic issues in SOCT, such as speckle, instrumentation and fabrication of tissue phantoms.
Chapter 5 discusses spectral analysis techniques that have be used in this study.
Chapter 6 discusses the experimental results, a comparison is given between the
methods that have been used and the results are thoroughly discussed. Finally, the
thesis is summarized and future studies are described in Chapter 7.
Chapter 2
Optical Coherence Tomography
2-1
Introduction
Optical Coherence Tomography (OCT) is a modern non-invasive biomedical imaging technique that generates in vivo cross-sectional images of tissue microstructure
with micron-scale spatial resolution [1]. OCT has its roots in white-light interferometry that led to the development of Optical Coherence-Domain Reflectometry
(OCDR), an 1-d optical ranging technique. OCDR uses a low coherence source
to perform optical ranging by detecting interference reflected by the reference and
sample arm. It was developed originally for finding faults in fiberoptic cables and
network components [13, 14]. Resolutions up to 10 µm and a dynamic range of
120 dB were demonstrated with OCDR [15]. The first biological applications were
in the field of the ophthalmology, where OCDR was used to obtain structural measurements from the eye [16, 17] and from other biological tissues [18, 19]. Later
OCDR was extended to obtain 2-dimensional images and OCT was created [1].
OCT was rapidly developed to a clinically viable instrument and the first medical
applications were reported [20, 21, 22].
OCT is analogous to ultrasound B-mode imaging except that it uses infrared
light, back-scattered from microstructures of the tissue for generating images. OCT
is using the principle of Low Coherence Interferometry (LCI) along with optical
heterodyne detection to obtain high axial resolution and high sensitivity to weakly
backscattered light from the sample. At the heart of the OCT is an interferometer
illuminated by a broadband light source. The typical OCT system (Figure 2.1),
uses a fiber based Michelson interferometer for splitting the light source beam into
a reference and a sample arm. The reference beam is retroreflected from a scanning
reference mirror at a known distance and returns to the detector, where it is recombined with the sample beam. The relative optical path difference between sample
and reference arm is contained in the interference signal, which is detected at the
photodetector. Subsequently, by signal processing electronics and computer data
2-2 Time-Domain OCT (TD-OCT)
16
acquisition, the position of reflective boundaries within the sample can be precisely
detected and generate 2-dimensional images. This section will derive the Fourier
transform relationship between the power spectrum of the light source and the axial
Point Spread Function (PSF). A discussion will be given for the effect of chromatic
dispersion on the axial resolution. The discussion follows derivations in the text
edited by Bouma and Tearney [23].
reference mirror
SLD
50/50
Beam Splitter
PC
D
photoreceiver
sample
Figure 2.1: Schematic of Time Domain OCT system with a moving mirror in the
reference arm.
2-2
Time-Domain OCT (TD-OCT)
2-2.1
Interferometer with Coherent Light
Consider the simplified schematic of the Michelson interferometer [24, 25] shown
in Figure 2.2. Light leaving from the source can been described as a plane wave
propagating in the z -direction
α1 = A exp(−jkz)
where
(2.1)
2π
(2.2)
λ
Then α1 enters in the interferometer and is divided at a beamsplitter into reference
k=
2-2 Time-Domain OCT (TD-OCT)
17
and sample arms. Beamsplitters, mirrors and in general all the mediums can be
described by a scattering matrix (ABCD Matrix):
−r jt
jt −r
!
(2.3)
where r is the reflectivity and t is the transmission. For the 50/50 beamsplitter the
ABCD matrix is:
!
√j
− √12
2
(2.4)
√j
√1
−
2
2
The two planes b1 and b2 exiting from the beamsplitter are:
b1
b2
!
=
√j
2
√
− 12
− √12
√j
2
!
a1
0
!
− √12 α1
=
√j
2
!
(2.5)
α1
The two fields are then reflected from the moving mirror (reference arm b2 ), and
backscattered from the sample (sample arm b1 ). They recombine at the detector as
⎡ − rr
⎢ 0
⎣
0 ⎤
− rr ⎥⎦
Mirror
lr
b2 = −
1
a1
2
1
a1rr exp ( − jklr )
2
ls
b1 =
a1
Source
Beam
Splitter
Es =
⎡
⎢−
⎢
⎢
⎣⎢
1
2
j
2
j ⎤
2 ⎥
⎥
1 ⎥
−
2 ⎦⎥
−
j
a1
2
j
a1rs exp ( − jkls )
2
⎡ − rs
⎢ jt
⎣ s
Sample
1 j
1 j
a1rs exp ( − jkls ) Er =
a1rr exp ( − jklr )
2 2
2 2
Detector
Figure 2.2: The Michelson interferometer.
jts ⎤
− rs ⎥⎦
2-2 Time-Domain OCT (TD-OCT)
18
one field
Ed = Er + Es =
Ed =
j
(α1 rr exp(−2jklr ) + α1 rs exp(−2jkls ))
2
j
(Ar exp(−2jklr ) + As exp(−2jkls ))
2
(2.6)
(2.7)
The detector measures only irradiance and not the electric field so the absolute of
the Equation 2.7 must be taken to obtain the photocurrent of the detector
Id = |Ed Ed∗ | =
Id =
1
[|Er Er∗ | + |Es Es∗ | + Es Er∗ + Er Es∗ ]
4
1
|As |2 + |Ar |2 + As A∗r exp (2jk(lr − ls )) + Ar A∗s exp (2jk(ls − lr ))
4
(2.8)
(2.9)
With the use of the trigonometry identity cos(θ) = 21 (exp(jθ) + exp(−jθ)) and by
taking only the real part Equation 2.9 becomes
Id =
1
|As |2 + |Ar |2 + <{As A∗r + Ar A∗s }cos(2k∆l)
4
(2.10)
where
∆l = lr − ls
(2.11)
∆l is the mismatch in distance between the reference and sample beam paths.
Equation 2.10 shows that the photocurrent of the detector has two components,
a DC and an AC component. The DC component (As , Ar ) is proportional to
the reflectance from each arm and AC component is sinusoidally modulated as a
function of the pathlength difference (∆l) between the two arms with an amplitute
proportional to the reflectivity of the sample arm. This component represents
the interference between the two arms because it contains all the backscattered
information from the sample. Replacing the wavenumber k in Equation 2.10 reveals
that the cosine term has a period of λ/2 relative to the length mismatch ∆l (Figure
2.3).
Also, depending on the pathlength mismatch, constructive or destructive interference occurs at the detector
π ∆l
1
)
Iinter = <{As A∗r + Ar A∗s }cos(
4
λ/2
(2.12)
The interference signal depends on the match between sample and reference fields
(Es , Er∗ ), and can be described by a cross correlation
Iinter ∝ <{Es Er∗ }
(2.13)
With TD-OCT, the mirror in the reference arm is moving with velocity ∆l =
2-2 Time-Domain OCT (TD-OCT)
19
Figure 2.3: Interferometric pattern resulting from moving the reference arm and
using a coherent source. Interferometric pattern resulting from moving the reference
arm and using a low coherence source.
2vm t and therefore a Doppler shift fD occurs
fD =
2vm
λ
(2.14)
So Equation 2.12 becomes:
Iinter
2-2.2
1
= <{As A∗r + Ar A∗s }cos(2πfD t)
4
(2.15)
Interferometer with Low Coherence Light
Section 2-2.1 describes interferometry with a monochromatic source. Equation 2.15
gives the photocurrent at the detector with a 50/50 beamsplitter and a moving
mirror in the reference arm. Low Coherence Interferometry (LCI), using a broad
bandwidth source of a finite bandwidth of frequencies rather than just a single
frequency requires a modification of Equation 2.12. First a low coherence source
can be represented as the sum of monochromatic sources:
Z
∞
A(k)exp(−jkz)dk
α1 =
(2.16)
−∞
The electric fields from the sample Es and reference Er can be represented as
functions of frequency:
Er = Ar (ω)exp(−2jkr (ω)lr )
(2.17)
Es = As (ω)exp(−2jks (ω)ls )
(2.18)
In this case the field at the detector will be:
Z ∞
Z ∞
j
Ed =
Ar (ω)exp(−2jkr (ω)lr )dω +
As (ω)exp(−2jks (ω)ls )dω
(2.19)
2 −∞
−∞
2-2 Time-Domain OCT (TD-OCT)
20
The interference signal, as shown before (Equation 2.13), is:
Iinter
Iinter ∝ <{Es Er∗ }
Z ∞
∗
Iinter ∝ <
Es Er dω
−∞
Z ∞
∝<
S(ω)exp (−j∆φ(ω)) dω
(2.20)
(2.21)
(2.22)
−∞
where
S(ω) = As (ω)Ar (ω)
(2.23)
∆φ(ω) = 2ks (ω)ls − 2kr (ω)lr
(2.24)
and
2-2.3
Nondispersive Medium
To examine the behavior in non-despersive media the reference and sample arm are
considered uniform, linear, and non-dispersive. The spectrum of the light source
S(ω − ω0 ) is also considered to be bandlimited with a center frequency of ω0 . As
a result the propagation constants in each arm can be considered equal and can be
re-written using a first order Taylor expansion around ω0
0
kr (ω) = ks (ω) = k(ω0 ) + k (ω0 )(ω − ω0 )
(2.25)
Inserting this to the Equation 2.22 gives:
Iinter
∞
d(ω − ω0 )
S(ω − ω0 )exp(−j(ω − ω0 )∆τg )
∝ < exp(−jω0 ∆τp )
2π
−∞
Z
(2.26)
where the phase delay mismatch ∆τp and the group delay mismatch ∆τg are defined
as
2∆l
k(ω0 )
∆τp = 2∆l
=
(2.27)
ω0
vp
0
∆τg = 2∆l k (ω0 ) =
2∆l
vg
(2.28)
Assuming that the sample and reference arm fields have the same spectral components as the light source, S(ω) is related to the power spectrum of the source:
Z
∞
S(ω) =
G(τ )exp(−jωτ )dτ
(2.29)
−∞
where G(τ ) is the complex temporal coherence function of the source. Using this
relation to Equation 2.26 with the Wiener − Khinchin theorem [26] it can be
2-2 Time-Domain OCT (TD-OCT)
21
obtained that
Iinter ∝ < {exp(−jω0 ∆τp )G(τ )}
(2.30)
Iinter ∝ |G(τ )|cos(ω0 ∆τp )
(2.31)
It follows that the interference signal depends on the temporal-coherence characteristics of the source. Also the interference signal consists of a carrier and an envelope.
The carrier oscillates with increasing path length mismatch 2∆l at a spatial frequency k(ω0 ) and the envelope is described by Equation 2.31 and determines the
axial point spread function of the interferometer.
2-2.4
OCT Resolution
It follows from the Equation 2.22 that the shape and width of the emission spectrum
of the light source are important variables in OCT. Laser sources with a broad
band spectrum can produce interference patterns of short temporal and spatial
extent. The relationship between S(ω) and G(τ ) from Equation 2.29, when both
are represented by gaussian functions, can be used to verify this as follows:
s
S(ω − ω0 ) =
2π
ω − ω0
exp −
σω2
2σω2
(2.32)
The Fourier transform will give the interferometric signal according to Equation
2.31
2 2
π ∆τ
Iinter ∝ exp −
cos(ω0 ∆τp )
(2.33)
2στ2
where the width is
στ =
1
σω
(2.34)
For obtaining the coherence length of the light source or the axial resolution
the Full Width Half Maximum (FWHM) of the gaussian function can be used.
The width of the gaussian source (σω ) is related to the FWHM of the wavelength
spectrum (∆λ) by:
∆ω
2πc∆λ
σω = √
= 2√
(2.35)
2ln2
λ0 2ln2
Therefore the FWHM of the interference signal, i.e. the axial resolution, in free
space will be
lc =
2cln2 1
2ln2 λ20
=
π σω
π ∆λ
2
λ
lc ≈ 0.44 0
∆λ
(2.36)
(2.37)
2-2 Time-Domain OCT (TD-OCT)
22
The axial resolution depends on the coherence length, which depends on the light
source spectrum. The SLD sources usually have Gaussian or near-Gaussian spectral
shape. However, for ultrabroadband laser sources such as the Ti:sapphire and
nonlinear fiber sources, the spectral shapes are often not Gaussian and the coherence
length is numerically calculated:
Z
∞
c
S 2 (ω)dω
lc = Z 0∞
2
S(ω)dω
(2.38)
0
The transverse resolution of OCT depends on the optics used and is usually
defined as the diameter of a gaussian beam, i.e. 2w:
r
2w =
2bλo
π
(2.39)
where b is the confocal parameter of the system. It can be shown that if the focal
length of the sample lens is f and the beam diameter incident onto the sample lens
is D , then
4f λ2o
(2.40)
w2 =
πD2
Therefore the focal length f of the lens in the sample arm should be chosen carefully
to meet both resolution and focusing depth requirements.
2-2.5
Group Velocity Dispersion
The light propagating from the source will experience Group Velocity Dispersion
(GVD) from the fibers, lens, and the specimen. The dispersion will affect the
envelope of the interferometric pattern and the axial resolution of the OCT system.
To examine the effect of GVD, the propagation constant must be expanded to a
second order Taylor series around ω0 :
0
00
kr (ω) = ks (ω) = k(ω0 ) + k (ω0 )(ω − ω0 ) + k (ω0 )(ω − ω0 )2
(2.41)
Assuming that a GVD mismatch exists in a length L of the sample and reference
arms then the effect of GVD to phase mismatch from Equation 2.24 is:
1 00
0
∆φ(ω) = k(ω0 )2∆l + k (ω0 )(ω − ω0 )2∆l + ∆k (ω0 )(ω − ω0 )2 (2L)
2
(2.42)
where
00
00
00
∆k (ω0 ) = kr (ω) − ks (ω)
is the GVD mismatch.
(2.43)
2-3 Optical Sources
23
Inserting ∆φ(ω) into the Equation 2.22 and neglecting the real operator the
interference pattern becomes:
Z
∞
S(ω − ω0 )
Iinter ∝ exp(−jω0 ∆τp )
−∞
1 00
d(ω − ω0 )
2
exp −j
∆k (ω0 )(ω − ω0 ) (2L)
exp(−j(ω − ω0 )∆τg )
(2.44)
2
2π
The effect of the dispersive medium is to broaden the interferometric signal and
induce assymetry due to chirp in the two interferometer arms.
Assuming, again, a gaussian spectrum for comparison with Section 2-2.3 the
resulting interferometric signal will be
Iinter
∆τg2
στ
exp −
exp(−jω0 ∆τp )
=
Γ(2L)
Γ(2L)2
(2.45)
where Γ(2L)2 is a complex function
00
Γ(2L)2 = στ2 + j∆k (ω0 )(2L)
(2.46)
The Gaussian function in Equation 2.45 has a real and an imaginary part defining
an envelope change as well as a chirp. Inserting Equation 2.46 and taking the real
part reveals that the gaussian envelope is broadened to the new width 2σ˜τ
"
2σ˜τ = 2στ 1 +
where
τcritical
στ
4 #1/2
00
1/2
τcritical = ∆k (ω0 )(2L)
(2.47)
(2.48)
Significant broadening will be present when τcritical > 1 which implies that in the
case of air vs. some other material even a few mm are enough to significantly
broaden the PSF.
Dispersion mismatch also decreases the amplitude of the interferometric envelope, as described by the multiplicative factor
1
στ
=
1/4
Γ(2L)
1 + (τcritical /στ )4
2-3
(2.49)
Optical Sources
The results of several theoretical and experimental studies cited in Section 2-2 show
the importance of the light source on the performance of the OCT system. The
general requirements are
2-3 Optical Sources
24
1. wavelength,
2. bandwidth,
3. single-transverse-mode power.
The first requirement rises from the need to operate in a spectral range in
which the penetration of light into tissue is adequate. The near-infrared end of the
spectrum is known as the therapeutic or diagnostic window, even if water absorption
increases there (Figure 2.4). Theoretical [22, 27] and experimental studies [21, 28,
29] in OCT are suggesting that the optimal image depth of penetration should occur
with sources emitting from 1200 nm to 1650 nm due to the fact that scattering losses
decrease in that range while water absorption is still relative low
Figure 2.4: Spectral absorption of some tissue molecules, and of aortic tissue. The
brightened wavelength ranges have already been used in OCT [30].
Axial resolution in OCT imaging is determined by the bandwidth of the light
source as described in Equation 2.37,
lc ≈ 0.44
λ20
∆λ
(2.50)
where λ0 is is the source center wavelength and ∆λ is the full width at half-maximum
(FWHM) spectral width. In general the broader the emission bandwidth of the
source the better the resolution that can be achieved. The mismatches between
the optical dispersion of the reference and sample arms [31] and the chromatic
aberration of the focused beam caused by scattering in the tissue [32]. Have to be
considered and compensate for by the selection of the source. Also the shape of the
2-3 Optical Sources
25
Table 2.1: Short coherent sources suitable for use in OCT systems [34]
Source
edge-emitting LED
Center
Bandwidth
Wavelength
(nm)
(nm)
1300, 1550
50-100
Emission
Power
Reference
20-300µW
Derickson et al. [35]
superluminescent
diodes (SLD)
800
1300
20-30
40-50
1-10mW
1-5mW
......
......
Multiple QW
LED/SLD
800
‘1400
90
90
15mW max
5mW max
Lin and Lee [36]
Poole et al. [37]
Laser pumbed fluorescent organic dye
590
40
9mW
Liu et al. [38]
Mode-Locked
T i : Al2 O3 laser
820
50-145
400mW
Bouma et al. [39]
Mode-Locked
Cr : forsterite laser
1280
75
30mW
Tearney et al. [40]
1550
1800
1060
40-80
80
65
10-100mW
7mW
108mW
Bouma et al. [28]
......
Paschotta et al. [41]
4+
Superfluorescent
optical fibers:
Er-doped
Tm-doped
Nd/Yb-doped
spectrum of the source is an important variable since real sources are not perfectly
Gaussian and may result significant sidelobes which, in turn, limit the dynamic
range of the scanner near strong reflections [33].
The power required from an optical source is constrained by the relation [23,
chap. 2] :
SN R vs
= const.
(2.51)
Ps ∆L
where SN R is the signal-to-noise ratio (or sensitivity), vs is proportional to the
axial scanning acquisition rate, and Ps is the maximum source optical power that
can reach the detector. This relation implies that for obtaining higher acquisition
rate and higher sensitivity requires an increased source power. Usually singlemode optical sources are used and the source power concentrated is limited by the
requirement that the diameter D of the source and its half-angle of emission α be
2-3 Optical Sources
26
related to the center wavelength [42] as follow
Dα = λ0
(2.52)
Therefore, the most effective sources for OCT imaging emit light from a small spot
over a wide angle or from a large spot over a narrow angle.
In Table 2.1 sources suitable for use in OCT systems are listed [34]. The most
commonly used sources are edge-emitting light-emitting diodes (ELED) and superluminescent diodes (SLD), with center wavelengths in either the 800or1300 nm.
SLD are the most commonly used because of their high power and relatively low
cost, but the coherence lengths of SLD (typically 15 − 30 µm) are not short enough
to achieve the resolution required for many medical applications. Efforts have
been made to overcome this disadvantage, one example is synthesizing a broadband
source by combining the outputs of several SLD with different center wavelengths
[43]. Also multiple quantum-well devices achieve this synthesis by coupling the
output of several sources on a single substrate [36, 37].
Chapter 3
Mie Theory
3-1
Introduction
In 1908, Gustav Mie [44] developed a theory in an effort to understand the varied
colors in absorption and scattering exhibited by small colloidal particles of gold
suspended in water. Mie theory provides insight into the biological origin of OCT
signals, and justifies the choice of spectral modulation as an analysis technique.
Measuring spectral modulation to determine the scatterer size and distribution had
been used extensively to the area of Light Scattering Spectroscopy (LSS). Significant
research in LSS has been carried out for measuring the size of scatterers in tissue
structures [45, 46] and to detect malignancies in-vivo [47] and in-situ [48, 49] and
to characterize particle size using tissue phantoms [50, 51]. The central idea behind
this research is that cellular organelles of epithelial tissue can be considered as
spherical particles whose interactions with light are described by Mie theory [52, 53].
The discussion begins with a general description of the scattering problem, and
a solution of Maxwell’s equations on spherical coordinates. Next a set of harmonics
vector solutions for the vector wave equation is presented. These functions provide a complete orthogonal basis for expressing a plane wave as an infinite series.
Subsequently the expansion coefficients for the infinite series and solutions for the
scattered field of a spherical particle are derived. The most interesting part for OCT
is the relation between backscattered intensity and illuminating wavelength. The
discussion of Mie theory follows derivations in the texts by Bohren and Huffman
[54] and de Hulst [55].
3-2
General Formulation of the Problem
The problem is defined as having a particle of specified size, shape and optical properties that is illuminated by an arbitrarily polarized monochromatic wave. The electromagnetic field at all points in the particle and at all points of the homogeneous
3-2 General Formulation of the Problem
28
medium in which the particle is embedded must be determined.
The field inside the particle is defined as (E1 ,H1 ) and the medium surrounding
field as (E2 ,H2 ), which is the superposition of the incident field (Ei ,Hi ) and the
scattered field (Es ,Hs ) (Figure 3.1).
E2 = Ei + Es
H2 = Hi + Hs
(3.1)
Hi = H0 exp(ik · x − iωt)
(3.2)
where
Ei = E0 exp(ik · x − iωt)
and k is the wave number of the surrounding medium.
Figure 3.1: The incident field (Ei ,Hi ) gives rise to a field (E1 ,H1 ) inside the particle
and a scattered field (Es ,Hs ) in the medium surrounding the particle.
The fields must satisfy the Maxwell equations
∇·E=0
(3.3)
∇·H=0
(3.4)
∇ × E = jωµH
(3.5)
∇ × H = −jωE
(3.6)
at all points where and µ are continuous. The curl of Equation 3.5 and 3.6 is
∇ × (∇ × E) = jωµ∇ × H = ω 2 µE∇ × (∇ × H) = −jω∇ × E = ω 2 µH (3.7)
3-2 General Formulation of the Problem
29
and using the vector identity
∇ × (∇ × A = ∇(∇ · A) − ∇ · (∇A))
(3.8)
results in
∇2 E + k 2 E = 0
∇2 H + k 2 H = 0
(3.9)
where k 2 = ω 2 µ and ∇2 A = ∇(∇ · A). Thus, E and H satisfy the vector wave
equation in a linear, isotropic, homogeneous medium, are also be divergence-free
∇·E=0
∇·H=0
(3.10)
In addition E and H are not independent as evident, from Equations 3.5 and 3.6.
3-2.1
Boundary Conditions
The electromagnetic field must satisfy Maxwell’s equations at points where and
µ are continuous. But on the boundary between particle and medium there is a
sudden change in these properties, i.e. a discontinuity. At such boundary points
the following conditions on the fields are imposed:
b=0
[E2 (x) − E1 (x)] × n
(3.11)
b=0
[H2 (x) − H1 (x)] × n
(3.12)
b is the outward directed normal to the surface S of the particle. The boundwhere n
ary conditions require that the tangential components of E and H are continuous
across a boundary separating media with different properties.
b , which is the boundary
Consider a closed surface A, with outward normal n
between regions 1 and 2 (Figure 3.2). The rate at which electromagnetic energy is
transferred across a closed surface arbitrarily near A in region 1 is
Z
A
Z
b dA =
S1 · n
A
b · (E1 × H1 ) dA
n
(3.13)
Similarly, the rate of electromagnetic energy transfer across a closed surface arbitrarily near A in region 2 is
Z
A
Z
b dA =
S2 · n
A
b · (E2 × H2 ) dA
n
(3.14)
3-2 General Formulation of the Problem
30
Figure 3.2: Closed surface separating regions 1 and 2.
b = E1 × n
b , H2 × n
b = H1 × n
b
If the boundary conditions are imposed, then E2 × n
and the integrals 3.13 and 3.14 may be written
ZA
A
Z
Z
Z
b dA =
S1 · n
b dA =
S2 · n
H1 · (b
n × E1 )dA =
H1 · (b
n × E2 )dA
(3.15)
H1 · (b
n × E2 )dA
(3.16)
ZA
ZA
E2 · (b
n × H2 )dA =
A
A
where we have used the permutation rule for the triple scalar product
A · (B × C) = B · (C × A) = C · (A × B)
3-2.2
(3.17)
Solutions to the vector wave equations
Suppose that, given a scalar function ψ and an arbitrary constant vector c, a vector
function M is constructed
M = ∇ × (cψ)
(3.18)
and
∇·M=0
(3.19)
If the vector identities are used
∇ × (A × B) = A(∇ · B) − B(∇ · A) + (B · ∇)A − (A · ∇)B
(3.20)
3-2 General Formulation of the Problem
∇(A · B) = A × (∇ × B) + B × (∇ × A) + (B · ∇)A + (A · ∇)B
31
(3.21)
result in
∇2 M + k 2 M = ∇ × [c(∇2 ψ + k 2 ψ)]
(3.22)
If ψ is a solution to the scalar wave equation, then M satisfies the vector wave
equation. From M we can construct N
N=
∇×M
k
(3.23)
with zero divergence, which also satisfies the vector wave equation
∇2 N + k 2 N = 0
(3.24)
∇ × N = kM
(3.25)
results in
So M and N have all the required properties of an electromagnetic field:
1. they satisfy the vector wave equation,
2. they are divergence-free,
3. the curl of M is proportional to N,
4. the curl of N is proportional to M.
Thus, the problem of finding solutions to the field equations reduces to the comparatively simpler problem of finding solutions to the scalar wave equation. We shall
call the scalar function ψ a generating function for the vector harmonics M and N,
the vector c is sometimes called the guiding or pilot vector.
Scattering by a sphere required functions ψ that satisfy the wave equation in
spherical polar coordinates, r, θ and φ (Figure 3.3). For pilot function radius vector
r is defined which
M = ∇ × (rψ)
(3.26)
Thus M becomes a solution to the vector wave equation in spherical polar coordinates and so is the associated N. The scalar wave equation in spherical polar
coordinates is
1 ∂
1
∂
∂ψ
1 ∂ 2ψ
2 ∂ψ
r
+
sinθ
+
+ k2ψ = 0
(3.27)
r2 ∂r
∂r
r2 sinθ ∂θ
∂θ
r2 sinθ ∂φ2
A separation of variables is first performed
ψ(r, θ, φ) = R(r)Θ(θ)Φ(φ)
(3.28)
3-2 General Formulation of the Problem
32
Figure 3.3: Scattering by a spherical particle
When substituted into 3.27 and by dividing by ψ(r, θ, φ) gives
1 d
Rr2 dr
1
d
dΘ
1
d2 Φ
2 dR
r
+
sinθ
+
+ k2 = 0
dr
Θr2 sinθ dθ
dθ
Φr2 sin2 θ dφ2
(3.29)
By multiplying by r2 sin2 θ the azimuthal part can be isolated
1 d
1
d
dΘ
1 d2 Φ
2
2
2
2 dR
= r sin θ −k −
r
−
sinθ
Φ dφ2
Rr2 dr
dr
Θr2 sinθ dθ
dθ
(3.30)
The variable φ can separated with a separation constant m
1 d2 Φ
= −m2
Φ dφ2
thus
1 d
R dr
r
2 dR
dr
1
d
+r k =−
Θsinθ dθ
2 2
(3.31)
dΘ
m2
sinθ
+
dθ
sin2 θ
(3.32)
The variables are separated. Each side is equated to a constant n(n+1), and finally
obtain
1 d
dΘ
m2
sinθ
−
Θ + n(n + 1)Θ = 0
(3.33)
sinθ dθ
dθ
sin2 θ
1 d 2 dR
(r
) + k 2 r2 R − n(n + 1)R = 0
(3.34)
R dr
dr
3-2 General Formulation of the Problem
33
A partial differential equation of three variables has been replaced by three ordinary
differential equations. The azimuthal solution Φ(φ) is consided first, beginning with
the requirement that is single-valued
Φ(φ + 2π) = Φ(φ)
(3.35)
Because of the periodicity of the azimuthal solution, m is a constant. The linearly
independent solutions are
Φe = cosmφ,
Φo = sinmφ,
(3.36)
where subscripts e and o denote even and odd. For the other two equations, the
solution is going to be a little more complex. Equation 3.33 is actually the associated
Legendre equation and Equation 3.34 after a substitution becomes the spherical
Bessel equation. We shall begin with the associated Legendre:
dP m (cosθ)
m2
1 d
(sinθ n
) + [n(n + 1) −
]P m (cosθ) = 0
sinθ dθ
dθ
sin2 θ n
(3.37)
One way of developing the solution of the associated Legendre equation is to start
with the regular Legendre equation and convert it to the associated by using multiple differentiation. Inserting x = cosθ to the Equation 3.33 transforms to
(1 − x2 )
d m
m2
d m
P
(x)
−
2x
P
(x)
+
[n(n
+
1)
−
]P m (x) = 0
dx2 n
dx n
1 − x2 n
(3.38)
The Legendre equation is
00
0
(1 − x2 )Pn − 2xPn + n(n + 1)Pn = 0
(3.39)
with solutions the Legendre polynomials Pn (cosθ) and after differentiate m times
the result is
00
0
(1 − x2 )u − 2x(m + 1)u + (n − m)(n + m + 1)u = 0
(3.40)
where
dm
Pn (x)
(3.41)
dxm
Equation 3.40 is not self-adjoint [56]. To put it into self-adjoint form and convert
the weighting function to unity, u(x) must be replaced by
u≡
v(x) = (1 − x2 )m/2 u(x) = (1 − x2 )m/2
dm
Pn (x)
dxm
(3.42)
3-2 General Formulation of the Problem
34
Solving for u and differentiating:
mxv
)(1 − x2 )−m/2 ,
2
1−x
0
2mxv
mv
m(m + 2)x2 v
00
u = [v +
+
+
](1 − x2 )−m/2
2
2
2
2
1−x
1−x
(1 − x )
0
u = (v +
(3.43)
(3.44)
Substituting into Equation 3.40 we find that the new function v satisfies the selfadjoint ordinary differential equation
00
0
(1 − x2 )v − 2xV + [n(n + 1) −
m2
]v = 0
1 − x2
(3.45)
which is the associated Legendre with solutions the m times differentiate Legendre
polynomials, the so called associated Legendre polynomials.
Pnm (cosθ) = (1 − cos2 θ)m/2
dm
Pn (cosθ)
dxm
(3.46)
For Equation 3.34, to transform it into a Bessel equation, substitute
R(kr) =
which results
r2
Z(kr)
(kr)1/2
d2 Z
dZ
1
+r
+ [k 2 r2 − (n + )2 ]Z = 0
2
dr
dr
2
(3.47)
(3.48)
which is a Bessel equation, Z is a Bessel function of order n + 12 . The linearly
independent solutions to Equation 3.48 are the Bessel functions of first, second and
third kind Jn , Nn Hn . It is convenient to label these functions spherical Bessel
functions
r
π
jn (x) =
Jn+1/2 (x)
(3.49)
2x
r
r
π
π
n+1
nn (x) =
Nn+1/2 (x) = (−1)
J−n−1/2 (x)
(3.50)
2x
2x
r
π (1)
(1)
hn (x) =
H
(x) = jn (x) + inn (x)
(3.51)
2x n+1/2
r
π (2)
(2)
hn (x) =
H
(x) = jn (x) − inn (x)
(3.52)
2x n+1/2
The spherical Bessel functions satisfy the recurrence relations
d n+1
[x fn (x)] =xn+1 fn−1 (x)
dx
d −n
[x fn (x)] = − x−n fn+1 (x)
dx
(3.53)
(3.54)
3-2 General Formulation of the Problem
(1)
35
(2)
Here fn may represent jn , nn hn or hn . From the first two orders
sin r
sin r cos r
, j1 (r) = 2 −
r
r
r
cos r sin r
cos r
, n1 (r) = − 2 −
n0 (r) = −
r
r
r
j
j
j
(1)
(1)
h0 (r) = − exp(jr), h1 (r) = exp(jr) − − 2
r
r r
j0 (r) =
(3.55)
(3.56)
(3.57)
higher-order functions can be generated by recurrence. Now generating functions
can be constructed that satisfy the scalar wave equation in spherical polar coordinates:
ψemn = cos mφ Pnm (cos θ)zn (kr)
(3.58)
ψomn = sin mφ Pnm (cos θ)zn (kr)
(3.59)
(1)
(2)
where zn is any of the four spherical Bessel functions jn , nn hn or hn . The vector
spherical harmonics generated by ψemn and ψomn
Memn = ∇ × (r ψemn ),
∇ × Memn
,
k
which in component form are
Nemn =
Momn = ∇ × (r ψomn ),
Nomn =
∇ × Momn
,
k
(3.60)
(3.61)
Memn =
−m
dP m (cos θ)
sin mφ Pnm (cos θ)zn (kr)b
eθ − cos mφ n
zn (kr)b
eφ ,
sin θ
dθ
(3.62)
Momn =
−m
dP m (cos θ)
cos mφ Pnm (cos θ)zn (kr)b
eθ − sin mφ n
zn (kr)b
eφ ,
sin θ
dθ
(3.63)
Nemn =
Nomn =
−zn (kr)
cos mφ n(n + 1)Pnm (cos θ)b
er
kr
dPnm (cos θ) 1 d
+ cos mφ
[krzn (kr)]b
eθ
dθ
kr d(kr)
Pnm (cos θ) 1 d
− m sin mφ
[krzn (kr)]b
eφ , (3.64)
sin θ kr d(kr)
zn (kr)
sin mφ n(n + 1)Pnm (cos θ)b
er
kr
dP m (cos θ) 1 d
+ sin mφ n
[krzn (kr)]b
eθ
dθ
kr d(kr)
3-3 Expansion of a plane wave in vector spherical harmonics
+ m cos mφ
36
Pnm (cos θ) 1 d
[krzn (kr)]b
eφ , (3.65)
sin θ kr d(kr)
With these vector harmonics can be solved the problem of scattering by an arbitrary
sphere.
3-3
Expansion of a plane wave in vector spherical
harmonics
First, consider a plane x- polarized wave, written in spherical polar coordinates as
Ei = E0 exp(jkr cos θ)b
ex
(3.66)
b
ex = sin θcosφb
er + cos θ sin φb
eθ − sin φb
eφ
(3.67)
where
by an arbitrary sphere. The first step toward the solution to this problem is expanding equation 3.66 in vector spherical harmonics:
Ei =
∞ X
∞
X
(Bemn Memn + Bomn Momn + Aemn Nemn + Aomn Nomn )
(3.68)
m=0 n=m
Because sin mφ is orthogonal to cos m0 φ for all m and m0 it follows that Memn and
Momn are orthogonal.
Z
2π
Z
π
Mem0 n0 Momn sin θdθdφ = 0
0
(all m, m’, n, n’)
(3.69)
0
Similarly, (Nomn , Nemn ), (Momn , Nomn ) and (Memn , Nemn ) are mutually orthogonal
sets of functions.
The orthogonality of all the vector spherical harmonics implies that the coefficients in the expansion are of the form
2π
Z
π
Z
Ei · Memn sin θdθdφ
0
0
Bemn = Z
2π
(3.70)
π
Z
2
|Memn | sin θdθdφ
0
0
2π
Z
π
Z
Ei · Momn sin θdθdφ
0
Bomn = Z
0
2π
Z
(3.71)
π
2
|Momn | sin θdθdφ
0
0
3-4 The internal and scattered fields
2π
Z
37
π
Z
Ei · Nemn sin θdθdφ
0
0
Aemn = Z
2π
(3.72)
π
Z
2
|Nemn | sin θdθdφ
0
0
2π
Z
π
Z
Ei · Nomn sin θdθdφ
0
0
Aomn = Z
2π
Z
(3.73)
π
2
|Nomn | sin θdθdφ
0
0
It follows from Equations 3.62-3.65 together with the the orthogonality of the sine
and cosine that
Bemn = Aomn = 0,for all m, n
(3.74)
Bomn = Aemn = 0,for all m = 1
(3.75)
The incident field is finite at the origin, which requires that the spherical Bessel jn
and nn will be also finite at origin. That is the reason why nn is rejected, because
of its misbehavior at the origin. Following the notation of [54, 55] a superscript (1)
is appended to the vector spherical harmonics for specifying the radial dependence
by jn . Thus, the expansion for Ei , has the form
Ei =
∞
X
(1)
(1)
(Boln Moln + Aeln Neln )
(3.76)
n=1
By intergrating Equations 3.70 and 3.72 the coefficients Boln and Aeln are obtained
2n + 1
n(n + 1)
2n + 1
= −jE0 j n
n(n + 1)
Boln = j n E0
(3.77)
Aeln
(3.78)
This leads to the expansion of plane wave in spherical harmonics
Ei = E 0
∞
X
n=1
3-4
jn
2n + 1
(1)
(1)
(Moln − jNeln )
n(n + 1)
(3.79)
The internal and scattered fields
Suppose that a plane x-polarized wave is incident on a homogeneous, isotropic
sphere of radius a (Figure 3.3). As shown, the incident electric field may be expanded in an infinite series of vector spherical harmonics. The corresponding inci-
3-4 The internal and scattered fields
38
dent magnetic field is obtained from the curl of Equation 3.79
∞
−k X n 2n + 1
(1)
(1)
E0
j
(Meln + jNoln )
Hi =
ωµ n=1 n(n + 1)
(3.80)
The scattered electromagnetic field (Es , Hs ) and the field (E1 , H1 ) inside the sphere
in vector spherical harmonics may also be expanded. At the boundary between the
sphere and the surrounding medium the following conditions are imposed
(Ei + Es − E1 ) × b
er = (Hi + Hs − H1 ) × b
er = 0
(3.81)
The expansion of the field (E1 , H1 ) can be calculated by using the orthogonality of
the vector harmonics, the boundary conditions 3.81 and the form of expansion of
the incident field. The field must be finite at the origin so we take jn (k1 r), where
k1 is the wavenumber inside the sphere.
E1 =
∞
X
(1)
(1)
En (cn Moln − jdn Neln )
n=1
∞
−k1 X
(1)
(1)
En (dn Meln − jcn Noln )
H1 =
ωµ1 n=1
(3.82)
where µ1 is the permeability of the sphere and
En =
j n E0 (2n + 1)
n(n + 1)
(3.83)
The scattered field does not exist at the origin, so both functions jn and nn are
involved in the expansion of the scattered field. It is convenient now to switch to
(1)
(2)
spherical Hankel functions (spherical Bessel of third kind), hn and hn for values
of kr2 n2
(−j)n exp(jkr)
h(1)
(kr)
∼
n
jkr
(3.84)
n
(j) exp(jkr)
(2)
hn (kr) ∼ −
jkr
These asymptotic expressions describe a spherical wave travelling away from the
(1)
(2)
origin hn and a spherical wave travelling towards the origin hn . For the scatterred
(1)
field one can use only hn as a generating function, therefore the expansion of the
field is
∞
X
(3)
(3)
Es =
En (jan Neln − bn Moln )
n=1
∞
X
k
Hs =
ωµ
n=1
(3.85)
(3)
En (jbn Noln
+
(3)
an Meln )
3-4 The internal and scattered fields
39
(1)
The superscript (3) specifies the radial dependence by the spherical Hankel hn .
3-4.1
Angle-Dependent Functions
It is now convenient to define the angle dependent functions for deriving more
compactly the expressions
πn =
Pn1
,
sinθ
τ=
dPn1
dθ
(3.86)
Upward recurrence relations for these functions can be developed
n
2n − 1
cos θπn−1 −
πn−2
n−1
n−1
τn = n cos θπn − (n + 1)πn−1
(3.87)
πn =
(3.88)
Also the linear combinations of πn and τn are orthogonal sets of functions
π
Z
Z
0
π
(πn − τn )(πn − τn ) sin θdθ = 0 (m 6= n) (3.89)
(πn + τn )(πn + τn ) sin θdθ =
0
The vector spherical harmonics(with m = 1), then are
Moln = cos φ πn (cos θ)zn (kr)b
eθ − sin φ τn (cosθ)zn (kr)b
eφ
(3.90)
Meln = − sin φ πn (cos θ)zn (kr)b
eθ − cos φ τn (cosθ)zn (kr)b
eφ
(3.91)
zn (kr)
b
er
kr
0
[krzn (kr)]
b
eθ
+ sin φ τn (cos θ)
kr
Noln = sinφ n(n + 1) sin θπn (cos θ)
0
[krzn (kr)]
b
+ cos φπn (cos θ)
eφ (3.92)
kr
zn (kr)
b
er
kr
0
[krzn (kr)]
b
+ cos φ τn (cos θ)
eθ
kr
Neln = cos φ n(n + 1) sin θπn (cos θ)
0
[krzn (kr)]
b
− sin φπn (cos θ)
eφ (3.93)
kr
It was shown how the functions jn and nn behave at the origin (θ = 0). It only
remains for us to show the behavior of the functions πn and τn which determine
the θ dependence of the fields. Polar plots of πn and τn for n = 1 − 5 are shown
in Figure 3-4.1 (θ = 0, 2π). There exists a symmetry for all plots around x-axis,
3-4 The internal and scattered fields
40
which is expected because the scattering particle is spherical and the incident light is
propagating along the axis of symmetry. Also as the n increases the number of lobes
increases, and the width of the forward and back - scattered lobes is decreasing.
Furthermore, the absence of a backscattering lobe in the polar plots of πn and τn
indicates that they are negative for backward directions, and all plot have a forward
lobe but the backscattered is vanished for alternative values of n. From the behavior
of these functions it follows that the scattered field will be distributed in a smaller
spatial area as the particle size increases and as the particle size increases there will
be decreased backscattering and increased forward scattering.
3-4.2
Scattering Coefficients
The next step is to obtain explicit expressions for the scattering coefficients. For a
given n there are four unknown coefficients an , bn , cn , and dn . The four independent
equations needed, which are obtained from the Equation 3.81 at the boundary r = α
in component form:
Eiθ + Esθ = E1θ
Eiφ + Esφ = E1φ
Hiθ + Hsθ = H1θ
Hiφ + Hsφ = H1φ
(3.94)
To obtain the four linear equations in the expansion coefficients the orthogononality
of sin φ and cos φ, the relations for the angle functions 3.89 the above boundary
relations with the expansions for the internal and scattering fields 3.79, 3.80, 3.82
and 3.85 are required. Along with the vector harmonics Equations 3.90, 3.91, 3.92
and 3.93 eventually result in four independent linear equations
jn (mx)cn + h1n (x)bn = jn (x),
0
0
(3.95)
0
µ[mxjn (mx)] cn + µ1 [xh1n (x)] bn = µ1 [xjn (x)] ,
(3.96)
µmjn (mx)dn + µ1 h1n (x)αn = µ1 jn (x),
(3.97)
0
0
[mxjn (mx)] dn + m[xh1n (x)] αn = m[xjn (x)],
(3.98)
where the prime indicates differentiation with respect to the argument in parentheses and the size parameter x and the relative refractive index m are
x = kα =
2πN α
,
λ
m=
k1
N1
=
k
N
(3.99)
N1 and N are the refractive indices of particle and medium, respectively. The four
simultaneous linear equations are easily solved for the coefficients of the field inside
3-4 The internal and scattered fields
41
10
10
5
5
0
0
5
5
10
10
15
10
5
0
5
10
15
10
10
5
5
0
0
5
5
10
15
10
5
0
5
10
15
15
10
5
0
5
10
15
15
10
5
0
5
10
15
15
10
5
0
5
10
15
15
10
5
0
5
10
15
10
15
10
5
0
5
10
15
10
10
5
5
0
0
5
5
10
10
15
10
5
0
5
10
15
10
10
5
5
0
0
5
5
10
10
15
10
5
0
5
10
15
15
10
8
10
6
4
5
2
0
0
2
5
4
6
10
8
15
10
15
10
5
0
a
5
10
15
b
Figure 3.4: Polar plots of the first five angle-dependent functions πn (a) and τn (b)
(n = 1, . . . , 5). All functions are plotted to the same scale.
3-5 Scattering Matrix
42
the particle
0
0
µ1 jn (x)[xh1n (x)] − µ1 h1n (x)[xjn (x)]
cn =
µ1 jn (mx)[xh1n (x)]0 − µh1n (x)[mxjn (mx)]0
0
(3.100)
0
µ1 mjn (x)[xh1n (x)] − µ1 mh1n (x)[xjn (x)]
dn =
µ1 m2 jn (mx)[xh1n (x)]0 − µh1n (x)[mxjn (mx)]0
(3.101)
and the scattering coefficients
0
0
µm2 jn (mx)[xjn (mx)] − µ1 jn (x)[mxjn (mx)]
an =
µm2 jn (mx)[xh1n (x)]0 − µ1 h1n (x)[mxjn (mx)]0
0
(3.102)
0
µ1 jn (mx)[xjn (x)] − µjn (x)[mxjn (mx)]
bn =
µ1 jn (mx)[xh1n (x)]0 − µh1n (x)[mxjn (mx)]0
(3.103)
The scattering coefficients can be simplified somewhat by introducing the RiccatiBessel functions:
ψn (kr) = krjn (kr),
ξn (kr) = krh1n (kr))
(3.104)
If the permeability of the particle and the surrounding medium are considered to
be the same, then
0
0
mψn (mx)ψn (x) − ψn (mx)ψn (mx)
an =
mψn (mx)ξn0 (x) − ξn (mx)ψn0 (mx)
0
(3.105)
0
ψn (mx)ψn (x) − mψn (mx)ψn (mx)
bn =
ψn (mx)ξn0 (x) − mξn (mx)ψn0 (mx)
(3.106)
Note that an and bn vanish as m approaches unity; this is as it should be: when
the particle disappears, so does the scattered field.
3-5
Scattering Matrix
The infinity series expansion, Equation 3.79 and 3.80 of the scattered field is assumed to be uniformly convergent, if nc terms are taken then the resulting error
will be arbitrarily small for all kr if nc is sufficiently large. Also if kr n2c the
3-5 Scattering Matrix
43
asymptotic expressions of the spherical Hankel can be substituted
h1n (kr)
(−j)n ejkr
,
∼
jkr
kr n2
(3.107)
dh1n (kr)
(−j)n ejkr
∼
d(kr)
jkr
(3.108)
in the truncated series; the resulting transverse components of the scattered electric
field are
ejkr
cos φS2 (cos θ)
−jkr
ejkr
sin φS1 (cos θ)
∼ −E0
−jkr
Esθ ∼ E0
Esφ
(3.109)
(3.110)
where
X 2n + 1
(αn πn + bn τn )
n(n + 1)
n
X 2n + 1
S2 =
(αn τn + bn πn )
n(n
+
1)
n
S1 =
(3.111)
(3.112)
and the series are terminated after nc terms. The resulting scattering matrix for
the relation between incident and scattered field amplitudes is
Eks
E⊥s
!
ejk(r−z)
=
−jkr
S2 0
0 S1
!
Eki
E⊥i
!
(3.113)
The relation between incident and scattered Stokes parameters


Is
 
Qs 
 = 1
 U  k2 r2
 s
Vs

 
S11 S12
0
0
Ii

 
S12 S11
 
0
0 

 Qi 
 0


0
S33 S34  

 Ui 
0
0 −S34 S33
Vi
1
S11 = (|S2 |2 + |S1 |2 ),
2
1 ∗
S33 = (S2 S1 + S2 S1∗ ),
2
1
S12 = (|S2 |2 − |S1 |2 )
2
i
S34 = (S1 S2∗ − S2 S1∗ )
2
(3.114)
(3.115)
(3.116)
2
2
2
2
Only three of these four matrix elements are independent:S11
= S12
+ S33
+ S34
.
From the Stokes parameters, the relation between incident light and the scattered light for unpolarized light is
Is = S11 (θ)Ii ,
Qs = S12 (θ)Ii ,
Us = Vs = 0
(3.117)
3-6 Computation of Scattering Coefficients
44
where the factor 1/k 2 r2 was ommited.
3-6
Computation of Scattering Coefficients
To obtain quantitative results from the Mie theory one must calculate the angular
functions πn and τn together with the scattering coefficients an and bn and sum
the series 3.111, 3.112 for the scattering matrix elements. However the scattering
coefficients an and bn are complicated functions of spherical Bessel functions and
their derivatives which make their computation cumbersome.
To simplify Equations 3.105 and 3.106 the logarithmic derivative can be employed [57]
d
ln ψn (ρ)
(3.118)
Dn (ρ) =
dρ
which simplifies Equations 3.105 and 3.106 to
an =
[Dn (mx)/m + n/x] ψn (x) − ψn−1 (x)
[Dn (mx)/m + n/x] ξn (x) − ξn−1 (x)
(3.119)
bn =
[mDn (mx) + n/x] ψn (x) − ψn−1 (x)
[mDn (mx) + n/x] ξn (x) − ξn−1 (x)
(3.120)
0
0
For eliminating ψn and ξn the reccurence relations were used
0
ψn = ψn−1 (x) −
nψn−1 (x)
,
x
0
ξn = ξn−1 (x) −
nξn−1 (x)
x
(3.121)
The logarithmic derivative satisfies Dn (mx), from the properties of spherical Bessel,
the recurrence relation
1
n
(3.122)
Dn−1 = −
ρ Dn + n/ρ
and it can be calculated with downward recurrence. Also ψn and ξn satisfy
ψn+1 (x)
2n + 1
ψn − ψn−1 (x)
x
(3.123)
Now to calculate ψn , ξn = ψn − jξn by using upward recurrence with starting
functions
ψ−1 (x) = cos x,
ψ0 (x) = sin x
(3.124)
ξ−1 (x) = − sin x,
ξ0 (x) = cos x
For speeding up the calculation a stop–criterion can be used, the whole process
will be stopped when Nstop terms are calculated. Where Nstop is the integer closest
to [58]
Nstop = x + 4x1/3 + 2
(3.125)
3-7 Mie Theory for Biomedical Optics
45
Perhaps the best known program for computing Mie scattering coefficients is
that by Bohren and Huffman, called BHMIE [54]. Although BHMIE is one of the
earliest programs for Mie theory, it has a wide distribution and is considered simple,
efficient and easy to modify.
3-7
Mie Theory for Biomedical Optics
The interaction between electromagnetic waves and tissue is important for the development of diagnostic and therapeutic applications of light in medicine. To better
understand the interaction of light within tissue it is useful to have a theoretical
model from which the tissue optical properties can be derived. By considering
cellular organelles of biological tissue as spheroidal scatterers whose interactions
with light are governed by Mie theory, optical properties such as refractive index,
scatterer size and distribution can be derived [52, 53]. The most relevant for OCT
imaging from Mie theory, is the relationship between backscattering intensity and
illuminating wavelength. Figure 3-4.1 suggests that as particle size increases there
will be decreased backscattering, meaning that smaller scattering particles in biological tissue will produce most of the OCT signal.
Chapter 4
Spectroscopic OCT
4-1
Introduction
As described at Chapter 2 OCT images are generated by analyzing the envelope
of the demodulated interference signal to produce two-dimensional cross-sectional
images. However a great deal of information in the backscattered signal is not utilize with traditional OCT systems. In addition, two tissues with the same amount
of backscattered intensity cannot be differentiate, even if those differ in cellular
and subcellular structure. To resolve this problem other methods can been used
for analyzing the backscatter signal such as spectroscopic analysis methods. For
example, different types of tissue will have different absorption and scattering spectra depending on the cell sizes, blood content, water content and whether the area
is oxygenated or deoxygenated. An extension of OCT is the Spectroscopic OCT
(SOCT) for performing both cross-sectional tomographic and imaging of the spectral content. SOCT may find applications in functional imaging by detecting variations in blood oxygenation, early detection of cancer by detecting variations in
cell size and nuclear density and for the detection of other pathologies by detecting
changes in spectroscopic properties of tissue types not visible by traditional OCT.
The first demonstration of SOCT had used the spectral centroid shift as the
indicator of spectral modification to enhanced the contrast of OCT images [9, 11].
SOCT rapidly become a powerful extension of OCT with many applications [10, 6,
59] and it has been demonstrated that SOCT could be combined with LSS [50, 60]
to eliminate the diffuse scattering background. As mentioned in Chapter 3 the
size of the scatterers be determined from the intensity of the scattered light as a
function of wavelength[50, 61], as well as from measurements of the scattering as a
function of angle [51, 12].
4-2 Mie theory in SOCT
4-2
47
Mie theory in SOCT
As discussed in Chapter 3, depending on the scatterer size, the spectral scattering
can be divided into three regimes. For particles much smaller compared to the
wavelength, the spectral scattering is governed by the Rayleigh scattering. For
particles much larger than the wavelength, the solution of spectral scattering can
be approximated by geometric optics. For particles comparable to the wavelength,
the spectral scattering can be solved using Mie theory. The most relevant process
in OCT is Mie theory because of the wavelength emission of OCT sources which is
at near–infrared spectra (1300 nm) and the dysplastic alterations which are of our
interest are in the same scale.
When the scatterer size is comparable to the wavelength, there is no simple
means to solve for the scattering contribution, other than by the formal solutions
of Maxwells equations and applying appropriate boundary equations. These calculations, collectively called Mie theory, dictate that the scattered wave for incident
plane waves depends on the distribution of the refractive index of the scatterers,
the incident wave frequency, and the scattering angle [55]. Mie theory is highly
spectrally dependent, often forming characteristic modulation patterns.
Figure 4-2 shows the characteristic difference of backscattered intensity for scatterers with different size. The backscattered intensity calculations are based on the
Fortran subroutine BHMIE [54] which is described at Chapter 3-6, BHMIE have
been modified and convert in Matlab source code in order to have uniformly sources
code. The results for the refractive index of the tissue phantoms from section 4-5,
and our source characteristics were used in the calculations. Because of the relative
narrow bandwidth of our source, modulation appears only for larger scatterer. However, spectral differences do exist even for smaller scatterers. By quantifying the
degree of spectral modulation in the backscattered spectrum information about the
size of the scattering particles can be obtained. This information was used as contrast modality for providing enhanced differentiation in areas where the scattering
particles sizes varies significantly.
4-3
Speckle
As mentioned already, OCT has micrometer-scale axial and lateral resolution, so in
the imaging volume of an OCT image many individual scattering particles exists simultaneously, such as mitochondria and nuclei. The existence of multiple scatterers
in the imaging volume of the OCT gives rise to speckle patterns. These patterns are
created by random interference of partial waves scattered by the scatterers. With
interference, stochastic random phase shifts occur between the waves which result in
statistical properties of the amplitude, phase, and intensity of speckle-modulated
4-3 Speckle
48
scattered fields. In addition, the spectral contents of the backscattered signal is
altered due to the size effects of these scatterers [62].
Speckle patterns formed by tissue scattering are often called biospeckles. The
possibility of obtaining information about structure and dynamic properties of tissues by using statistical or correlation analysis of biospeckles have been studied
extensively in many areas [63, 64, 65]. Speckle is an unresovable signal and its
properties depend on the size and the distribution of scatterers i.e. concentration,
spacing and periodicity. In SOCT, speckle appears in both the intensity of the
backscattered light and the modulation patterns of the local spectra. The modulation can be considered as an outcome of a stochastic process where the properties
of the process are dependent on the size and distribution of the scatterers within
the imaging volume. By analyzing the statistical properties of the backscattered
spectrum one can extract information about the scattering particles in the imaging
volume of OCT.
−4
8
x 10
−4
1.2
1.15
Backscattered Intensity (A.U.)
7
Backscattered Intensity (A.U.)
x 10
6
5
4
1.1
1.05
1
0.95
0.9
3
0.85
2
1.22
1.23
1.24
1.25
1.26
Wavelength (µ m)
1.27
1.28
0.8
1.22
1.29
1.23
1.24
1.25
1.26
Wavelength (µ m)
−6
x 10
(a)
1.27
1.28
1.29
(b)
−6
9
x 10
Backscattered Intensity (A.U.)
8
7
6
5
4
3
1.22
1.23
1.24
1.25
1.26
Wavelength (µ m)
1.27
1.28
−6
x 10
1.29
−6
x 10
(c)
Figure 4.1: Example of back-scattered intensity. (a) Back-scattered intensity for
1 µm scatterer. (b) Back-scattered intensity for 2 µm scatterer. (c) Back-scattered
intensity for 4 µm scatterer.
4-4 SOCT Instrumentation
49
Statistical properties of speckle can be divided into statistics of the first and
second order. Statistics of the first order describe the properties of speckle fields
at each point. Such description employs the intensity probability density function
ρ(I) which is represented as a negative exponential function
I
1
exp −
ρ(I) =
hIi
hIi
(4.1)
This expression is valid for fully developed speckles and when the scattering amplitude follows gaussian statistics.
Statistics of the second order show how fast the intensity changes from point to
point in the speckle pattern and they can characterize the size and the distribution
of speckle in the pattern. Goodman [66] and Goldfischer [67] were the first that use
the power spectrum density (PSD) and autocorrelation analysis for studying the
statistics of the second order. The correlation of intensity for two different points
is (x1 , y1 ) and (x2 , y2 ) is
rxx (x, y) = E {I(x1 , y1 )I(x2 , y2 )}
(4.2)
and its Fourier transform represents the power spectrum of a stochastic process. By
analyzing the power spectrum statistical properties diagnostic useful information
can be extracted.
4-4
SOCT Instrumentation
Both standard TDOCT and FDOCT instrumentation can be used for SOCT. For
the purpose of this study, a fiber-based TDOCT system with a broad bandwidth
source was used, Figure 4.2 shows the schematic of the OCT system. The system
uses a broadband superluminescent diodes (Superlum Broadlighter 1300), with a
center wavelength of 1300 nm and a bandwidth of 55 nm. Resulting an axial
resolution of 13.5 µm and lateral resolution of 25.6 µm. The interfering signal and
reference light was detected using a New Focus Model 2011 photoreceiver detector.
The interference fringes are acquired without demodulation using a 16-bit A/D card
(National Instruments, PCI-6251) and the signal was digitized at 6x the carrier
frequency.
4-4.1
Superluminescent diodes (SLDs)
As described in Section 2-3, the laser source spectrum and power are important
factors in determining the OCT resolution and sensitivity. In SOCT, this is especially important because SOCT uses the broad-band laser spectra to perform
4-5 Tissue Phantoms
50
spectroscopic analysis. The spectral features to be determined must lie within the
laser source spectrum.
At present, the most popular light sources in OCT are the Superluminescent
diodes (SLDs). SLDs are optoelectronic semiconductor devices which are emitting broadband optical radiation based on superluminescence. Their structure and
operation is similar to the edge-emitting laser diodes (EELD), containing an electrically driven p-n junction and an optical waveguide, but they lack an optical
resonator, thus no laser action can occur. Optical resonators are suppressed by
tilting the facets relative to the waveguide and can be suppressed even further with
anti-reflection coatings. Essentially, an SLD is a semiconductor optical amplifier
with no input signal.
Depending on the material and structure, SLDs have different center wavelengths and bandwidths. The broadband SLD source that used for this thesis was
developed by Superlum Diodes, Ltd., and consists of two single-mode fiber coupled SLD modules with slightly shifted central wavelengths. The first module has
emission wavelength of 1285.5 nm with a bandwidth of 57.4 nm and the second
module has emission wavelength of 1285.1 nm with a bandwidth of 55.3 nm. These
emissions are combined by appropriate broadband singlemode fiber couplers to give
a centered wavelength of 1300 nm and a bandwidth of 55 nm.
4-5
Tissue Phantoms
For the purpose of this study solid tissue phantoms were made using polystyrene
sphere solutions (Polybead Microspheres) of diameter 1 µm, 2 µm and 4 µm, and
with a refractive index n = 1.59, embedded in an acrylamide gel. The imaging
volume of the OCT system was approximately 30 µm x 30 µm x 15 µm. For
reference mirror
SLD
50/50
Beam Splitter
PC
D
photoreceiver
sample
Figure 4.2: Schematic of the Time Domain OCT system that we used.
4-5 Tissue Phantoms
51
Table 4.1: Recipes for tissue phantoms (acrylamide gel) of different imaging volume
concentration
Microspheres
diameter
Concentration
(Imaging Volume)
1 µm
2 µm
4 µm
50 spheres
5 spheres
2 spheres
dH2 O (µL)
82.5
684
375
Polystyrene
Microspheres (µL)
667.5
66
375
30% acrylamide (µL)
250
250
250
10% APS (µL)
20
20
20
TEMED (µL)
20
20
20
the 1 µm phantom the concentration was 50 spheres per imaging volume, for the
2 µm phantom 5 spheres per imaging volume and for the 4 µm 2 spheres per
imaging volume. These phantoms were selected because they result in samples
with similar backscattering coefficient and have enough scatterers in the imaging
volume to create speckle modulation.
An acrylamide gel is a separation matrix used in electrophoresis of biomolecules,
such as proteins or DNA fragments. The following steps must be followed for the
production of the phantoms:
• Determine the appropriate gel composition and microspheres concentration(Table 4.1) for a given experiment.
• Combine all reagents (except TEMED) in the order listed.
• Add TEMED only when completely ready for polymerization to occur!
• Allow gel to polymerize (∼ 2min)
As mentioned in Chapter 2 OCT measures the group delay imposed by the
sample, (Equation 2.45), if the thickness of the sample z is known the group index
can be calculated from the group delay. By placing the sample on top of a planar
reflecting surface and acquiring an OCT image, the refractive index of the sample
4-5 Tissue Phantoms
52
can be determined. The refractive index of the sample is defined as [68]:
n=
z0 + z
z
(4.3)
where z 0 is the additional optical path length. From such a calculation the refractive
index of the phantom was found to be n = 1.47.
Figure 4.3: OCT image of a piece of the phantom placed on an unpolished metal
substrate. The two vertical bars represent z, z 0 , from top to bottom.
Chapter 5
Spectral Analysis of OCT Signals
5-1
Introduction
In SOCT, the information on the spectral content of backscattered light is obtained
by the time-frequency analysis of the interferometric OCT signal. The backscattered spectrum can be acquired over the entire available optical bandwidth in a
single A-scan, allowing the spectroscopic information to be analyzed in a depthresolved way [69]. In the literature, the short-time Fourier transform (STFT) and
the continuous wavelet transform (Morlet transform) have been used for obtaining
the backscattering spectrum.
In both of these techniques, performance is complicated by the time-frequency
uncertainty principle, which states that there exists an inherent trade-off between
the spectral resolution and the time resolution. Improvement in one implies degradation in the other. In our case, high spatial resolution is required because spectral
back-scattering is a short-range effect in that large spectral variations can happen
within submicrometer scale distance (cell or tissue boundaries).
From the work done in LSS area there are currently two approaches for sizing
the scatterers based on the measured spectra. The first approach is based on pitch
detection with the use of Fourier transform or determining the autocorrelation of the
transform. The principle behind the first approach is that the oscillation frequency
in the wavelength-dependent scattering is size dependent, such that larger scatterers
tend to produce more oscillatory patterns [45]. The second approach is based on
curve fitting using least-square or χ2 methods [12] using Mie theory to fit the
normalized experimental measurement to the theoretical predicted.
The scope of this thesis is to introduce a novel spectral analysis method for
SOCT signals. It is known from previous work in the SOCT area that the spectral
features can be indicators of scatterer size by using spectral analysis methods such
as the Fourier transform or the autocorrelation [11, 60] There are many kinds of
spectral analysis methods, including Fast Fourier Transformation (FFT) method,
5-2 Spectral Analysis Based on Fourier Transformation
54
Yule-Walker method, Burg method, Least Squares method and Maximum likelihood
method. The Burg method was explored over other methods because it:
• Offers high resolution
• Converge faster for shorter segments
• Is computationally more efficient than other spectral techniques.
In this chapter the methods that were used for spectral analysis of the OCT signals,
and the methods for statistical multivariate analysis will be described.
5-2
Spectral Analysis Based on Fourier Transformation
The original or “classical” methods for spectrum estimation are based directly on
the Fourier transform and were used in the early 1950s. These methods gained
additional popularity with the development of the fast Fourier transform (FFT) in
the mid-1960s. In spite of development, of newer, more “modern” techniques, the
classical methods are still preferred for long and stationary data.
5-2.1
Energy Spectrum Density (ESD)
The Spectral Density of a sampled signal or time series x(n), −∞ < n < ∞ is a
positive real function of a frequency variable associated with a stationary stochastic
process, or a deterministic function of time, which has dimensions of power per Hz,
or energy per Hz. The Energy Spectrum Density (ESD) describes how the energy
(or variance) is distributed with frequency and is defined by the module square of
its Fourier transform. The Fourier transform of the sequence x(n) is
X(ω) =
∞
X
x(n)exp(−jωn)
(5.1)
x(n)exp(−j2πf n)
(5.2)
n=−∞
or, equivalently
X(f ) =
∞
X
n=−∞
The ESD is defined
∞
X
2
Sxx (ω) = |X(ω)| = X(ω)X (ω) =
x(n)exp(−jωn)
2
∗
n=−∞
(5.3)
5-2 Spectral Analysis Based on Fourier Transformation
55
The autocorrelation function is defined by
rxx (k) =
∞
X
x∗ (n)x(n + k)
(5.4)
n=−∞
The Fourier transform of the autocorrelation function with the use of WienerKhintchine theorem
∞
X
Sxx (ω) =
rxx (k)exp(−jkωn)
(5.5)
n=−∞
Hence the ESD also can be obtained by the Fourier transform of the autocorrelation.
From the relations above there are two different methods for computing the ESD.
The first method is the direct method which involves computing the Fourier transform of x(n) and then the square power of it to obtain Sxx (ω). The second method
is called the indirect method because it requires two steps. First the computation
of autocorrelation rxx (k) and then the Fourier transform of the autocorrelation is
computed as in Eq. 5.5 to obtain the ESD.
In practise, only the finite-duration sequence x (ns ), 0 ≤ ns ≤ N − 1 is the
total number of data samples for computing the spectrum of the signal. In effect,
limiting the duration of the sequence is equivalent to multiplying by a rectangular
window. Thus we have

x(n ) 0 ≤ n ≤ N − 1
s
s
(5.6)
x̃(ns ) = x(n)w(n) =
0
otherwise
The ESD of the windowed sequence x̃(ns ) is given by
−1
N
2
X
Sx̃x̃ (f ) = |X̃(f )| =
x̃(n)exp(−jωn)
2
(5.7)
ns =0
The spectrum given by Eq. 5.7 can be computed numerically from the Discrete
Fourier Transform (DFT) of N points [70].
5-2.2
Power Spectrum Density (PSD)
The SOCT signals due to speckle, as we mention at Chapter 4-3, are stochastic process and such signals have finite energy and are characterized by a Power Spectrum
Density (PSD). Lets define the autocorrelation function of a stationary stochastic
process x(t)
γxx (τ ) = E {x∗ (t)x(t + τ )}
(5.8)
5-2 Spectral Analysis Based on Fourier Transformation
56
where E is defined as the mean operator. The autocorrelation estimation for the
sampled sequence x(n) of the signal is
1
rxx (m) =
N
1
rxx (m) =
N
N −m−1
X
n=0
N
−1
X
x∗ (n)x(n + m)
0≤m≤N −1
(5.9)
x∗ (n)x(n + m)
m = −1, −2, . . . , 1 − N
n=|m|
With the use of Wiener–Khintchine theorem we obtain the PSD
Pxx (ω) =
N
−1
X
rxx (m)exp(−jωm)
(5.10)
m=−(N −1)
By inserting 5.9 into Equation 5.10 we have
Pxx (ω) =
N −1
2
1 X
1
x(n)exp(−jωn)
= |X(ω)|2
N n=0
N
(5.11)
This form of the PSD is defined as the Periodogram and it is popular for spectrum
estimation because the computation is especially convenient.
5-2.3
Computation using the FFT
As given by Equations 5.7 and 5.11 the estimated ESD Sxx (ω) and the periodogram
Pxx (ω) respectively can be computed by the use of the DFT. The DFT is efficiently
computed by Fast Fourier Transform (FFT) algorithm. However if the data segment consists of N points, then an N -point FFT produces only N samples of the
corresponding Fourier transform. This computation does not provide a very good
representation of the spectrum, to overcome this the Fourier transform can be evaluated at a larger and more closely spaced frequency values by “zero padding” the
data sequence and using a larger FFT. Zero-padding does not change the shape of
the spectrum, it only provides a way to evaluate the spectrum at a more closely
spaced set of frequency values.
Spectral analysis based on Fourier Transformation provides a simple, reliable
and efficient way for obtaining the spectrum. It requires only two computational
steps function and then with the use of FFT the spectrum can be obtained. The
major disadvantage of this method for spectral estimation is the limited resolution,
i.e. the limited ability to distinguish closely spaced features of the spectrum.
5-3 Spectral Analysis Based on Burg’s Method
5-3
57
Spectral Analysis Based on Burg’s Method
The non-parametric methods described above make no assumption about how the
data were generated and the estimation is based entirely on a finite record of data.
The assumption in Equation 5.10 that the autocorrelation estimate rxx is zero for
m ≤ N severely limits the frequency resolution and the quality of the estimated
PSD. The parametric or model–based methods of spectral estimation do not require
such assumptions and in fact Burg’s method bypasses any estimation of the autocorrelation function. Parametric methods for rational spectra and Burg’s method
for AR spectrum parameter estimation are discussed below. AR spectrum parameter estimation has been used in data forecasting, speech coding and recognition,
model-based spectral analysis, model-based interpolation, signal restoration. Also
it has been used for ultrasonic tissue characterization [71, 72].
5-3.1
Rational Power Spectra
The power spectrum density of a stochastic process could be expressed as a rational
function
2
2 B(ω) (5.12)
Pxx (ω) = σ A(ω)
where σ 2 is a positive scalar, and the polynomials A(ω) and B(ω) are
A(ω) = 1 +
q
X
αk exp(−jkω)
k=1
B(ω) =
p
X
(5.13)
bk exp(−jkω)
k=0
This function can similarly be expressed in the z–domain
φ(z) =
∗
∗
2 B(z)B (1/z )
σ
A(z)A∗ (1/z ∗ )
(5.14)
where the polynomials A(z) and B(z) have roots that fall inside the unit circle in
the z–plane. The rational PSD can be associated with a signal obtained by filtering
white noise of power σ 2 with transfer function H(ω) = H(z)|z=exp(jω = B(z)/A(z)
[73]
The filtering gives us a relation between the PSD as a signal x(n) with the white
noise w(n)
B(p)
x(n) =
w(n)
(5.15)
A(q)
It can be transformed into a difference equation, from where three specific cases
5-3 Spectral Analysis Based on Burg’s Method
58
can be separated
x(n) +
p
X
αk x(n − k) =
k=1
q
X
bk w(n − k)
(5.16)
k=0
1.
Autoregressive (AR) process. Where b0 = 1, bk = 0, k > 0 resulting
a linear, all-pole filter H(z) = 1/A(z) and the difference equation
x(n) +
p
X
αk x(n − k) = w(n)
(5.17)
|b0 |2
|A(ω)|2
(5.18)
k=1
The PSD for AR model is
AR
Pxx
(ω) =
For the AR model the constant |b0 |2 can be replaced by the error variance σP2 in
the PSD giving
σP2
AR
(5.19)
Pxx
(ω) =
|A(ω)|2
2.
Moving Average (MA) process. Where αk = 0, k ≤ 1 resulting a
linear, all-zero filter H(z) = B(z) and the relationship below
x(n) =
q
X
bk w(n − k)
(5.20)
k=0
The PSD for MA model is
MA
Pxx
(ω) = |B(ω)|2
(5.21)
3.
Autoregressive Moving Average (ARMA) process. In this case
the linear filter H(z) = B(z)/A(z) has both finite poles and zeros in the z –plane
and the corresponding difference equation is given by Equation 5.16. The PSD for
ARMA model is
|B(ω)|2
ARM A
(5.22)
Pxx
(ω) =
|A(ω)|2
5-3.2
Burg’s Method
It is obvious from the relations above that the estimation of the ARMA parameters, ak and bk is essential for the estimation of the PSD. Burg devised a method
for estimating the AR parameters as an order–recursive least square lattice method
[74]. The AR parameters with Burg’s method can be directly estimated from the
data without the intermediate step of computing a correlation matrix and solving
Yule-Walker equations, like other methods. Burg’s method is based on the minimization of the forward and backward in linear errors with the constraint that the
AR parameters satisfy the Levinson-Durbin recursion.
5-3 Spectral Analysis Based on Burg’s Method
59
Assume we have data measurements x(n), n = 1, 2, . . . , N , we define the forward
and backward prediction errors for a pth–order model as
êf,p (n) = x(n) +
m
X
α̂p,k x(n − k)
t = p + 1, . . . , N
(5.23)
k=1
êb,p (n) = x(n − p) +
m
X
∗
α̂p,k
x(n + k − m)
t = p + 1, . . . , N
(5.24)
k=1
where the AR parameters α̂p,k are related to the reflection coefficients K̂p with the
constraint that they satisfy the Levinson-Durbin recursion [? ] given by
α̂p,k =

α̂
p−1,k
∗
, i = 1, . . . , p − 1
+ K̂p α̂p−1,p−i
k̂ ,
p
(5.25)
i=p
The next step in Burg’s method for estimating the AR parameters is to find the
reflection coefficients K̂p that minimize the least–squares error
h
i
f
b
Em = min Êm
(p) + Êm
(p)
k
where
f
Êm
(p)
=
N
X
|êf,p (k)|2
k=p+1
b
Êm
(p) =
N
X
(5.26)
(5.27)
|êb,p (k)|2
k=p+1
Also the prediction errors satisfy the following recursive
êf,p (n) = êf,p−1 (n) + K̂p ê∗b,p−1 (n − 1)
êb,p (n) = êb,p−1 (n − 1) + K̂p ê∗f,p−1 (n)
(5.28)
Inserting Equation 5.28 in Equation 5.25 and performing minimization of Em
with respect to the reflection coefficients K̂m results in
N
X
−2
K̂m =
êf,p−1 (n) ê∗b,p−1 (n − 1)
k=p+1
N
X
(5.29)
|êf,p−1 (n)|2 + |êb,p−1 (n − 1)|2
k=p+1
The denominator is the least-squares estimate of the forward and backward errors
5-4 Statistical Multivariate Analysis
60
f
b
Êm−1
and Êm−1
respectively. Thus
−2
K̂m =
N
X
êf,p−1(n) ê∗b,p−1 (n − 1)
k=p+1
h
i
f
b
Êm−1
+ Êm−1
(5.30)
f
b
where Êm−1
+ Êm−1
is an estimate of the total error Em . A recursion could be
developed for estimation the total least square error
Em = (1 − |K̂m |2 )Em−1 − |êf,p−1 (m − 1)|2 − |êb,p−1 (m − 2)|2
(5.31)
To summarize, the Burg’s method estimates the reflections coefficients from
Equations 5.30 and 5.31 and the Levinson-Durbin algorithm is used to obtain the
AR model parameters α̂(k). From Equation 5.19 and the estimates of the AR
parameters we form the PSD
Êp
Bu
P̂xx
(ω) = p
2
X
1
+
α̂
(k)exp(−jωk)
p
(5.32)
k=1
5-4
Statistical Multivariate Analysis
The statistical analysis of the spectra obtained from the methods that described
above is essential for scatterer size estimation and for the classification. Backscattered OCT spectra are multivariate functions of wavelength. An algorithm for
studying such spectra should use multivariate statistical analysis. Advanced analysis techniques, such as Principal Component Analysis (PCA) and Multivariate
Analysis of Variance (MANOVA), were used for providing more insight into the
spectral changes. PCA is used as a dimension reducing technique and for extracting features from the training set to be used for size estimation and for classification.
MANOVA is used for discriminant analysis and classification.
5-4.1
Principal Component Analysis (PCA)
Principal component analysis (PCA) is a method of statistical analysis useful in
data reduction and interpretation of multivariate data sets [75]. This is achieved
by transforming to a new set of variables, the principal components (PCs), which
are mutually-orthogonal linear combinations and are order based on the variance of
the original variables. PCs is a smaller data set with less redundancy that give as
a very good representation of the original data set and this makes PCA a powerful
and versatile analysis method.
5-4 Statistical Multivariate Analysis
61
The starting point for PCA is a random vector x with n elements, and x has
zero empirical mean, is centered by subtracting its mean. A linear combination of
the vector x can be stated
y1 =
n
X
wk1 xk = w1T x
(5.33)
k=1
where w11 , . . . , wn1 are elements of an n–dimensional weight vector w1 .
If y1 has maximum variance then is the first PC of x, variance depends on both
the norm and orientation of the weight vector w1 the constraint that the norm
equals to 1 must be imposed, kw1 k = 1. The variance of y1 is defined as
E{y12 } = E{(w1T x)2 } = w1T E{xxT }w1 = w1T Cx w1
(5.34)
where Cx is the m x n covariance matrix of x. The weight vector that maximizes Equation 5.34 must be calculated so that kw1 k = 1 , it is well known that
the eigenvectors of the covariance matrix Cx , e1 , . . . , en are the solutions for the
maximization of variance [76]. Thus the first PC is given by
y1 = eT1 x
(5.35)
Generalized Equation 5.34 to k PCs, yk = wkT x one can find the rest PCs under
the constraint that yk is uncorrelated with all the previously found PCs. It follows
that
wk = ek
(5.36)
Thus the kth PC is
yk = eTk x
(5.37)
It have been showed that the PC basis vector wk are eigenvectors ek of the
covariance matrix Cx , it follows that
2
E{ym
} = eTm E{xxT }em = eTm Cx em = dm
(5.38)
where dm are the eigenvalues of Cx . Thus by ordering the eigenvectors found from
the covariance matrix by eigenvalue, from highest to lowest, this gives PCs in order
of significance. A very important practical problem is to choosing the number of
PCs, m in Equation 5.38. There is a trade-off between error and the amount of
data needed because sometimes a rather small number of PCs are sufficient.
5-4 Statistical Multivariate Analysis
5-4.2
62
Multivariate Analysis of Variance (MANOVA)
Multivariate Analysis of Variance (MANOVA) is used to obtain new linear combinations of variables with maximum separation between categories. The data in
MANOVA may be regarded as forming a matrix X in which each row corresponds
to an object nj and each column corresponds to a measured variable. X is an m x n
matrix of data values and each row is a vector of measurements on n variables for
a single observation. We have performed a one-way MANOVA for comparing the
multivariate means of the columns of X, grouped by group. Group is a grouping
variable defined as the size array in our case, two observations are in the same
group if they have the same value in the group array. The MANOVA test makes
the following assumptions about the data in X:
• The populations for each group are normally distributed.
• The variance-covariance matrix is the same for each population.
• All observations are mutually independent.
5-4.3
Classification
A subject closely related to MANOVA is that of classification and discriminant
analysis. A classification algorithm is used to determine in which category each
sample belongs. For this classification, a discriminant quadratic type function can
be used which fits a multivariate normal density to each group, with covariance
estimates stratified by group. From the discriminated data set we can compute the
sensitivity and the specificity of our classification.
Sensitivity is defined as
sensitivity =
number of True Positives
number of True Positives + number of False Negatives
(5.39)
and in diagnostic methods tell us what percentage of the data set have been recognized with the disease.
Specificity is defined as
specificity =
number of True Negatives
number of True Negatives + number of False Positives
(5.40)
and in diagnostic methods tell us what percentage of the data set have been recognized without the disease.
Chapter 6
Results
6-1
Introduction
As described at chapter 4-4 a fiber optic Michelson interferometer was used to measure the relative optical path difference between a reference arm and the phantoms
tissues layers at the sample arm 4.2. By mechanically varying the path length in the
reference arm with a scanning galvanometer an interference fringe is obtained for
an arbitrary path length in the sample. To generate a single Axial scan (A-scan) in
the image, an optical beam is incident on the sample at a single transverse location
while the reference arm is scanned. Each A-scan resembles a high frequency carrier,
amplitute modulated with a low frequency signal. The carrier signal contains all
the spectroscopic information of the sample, while the modulation contains information about the scattering structure. For generating a 2 − d image (B–scan), the
beam is translated in a transverse direction over the sample.
As mentioned before, our OCT system has an imaging volume approximately
30 µm x 30 µm x 15 µm, and the phantoms enough scatterers in the imaging volume
for creating speckle modulated patterns. Speckle is essential for our spectral analysis
method. For each phantom 200 A-scans were acquired with the first 100 A-scans
used as training set and the rest 100 as sample set. Axial scans were sampled at
6 x the carrier frequency. Collected data were analyzed using Matlab.
6-2
Spectra Calculation
Following the acquisition, preprocessing the obtained data begun with the normalazition of the Root Mean Square (RMS) value to unity. This removed the effect
of intensity differences between scans and samples. Subsequently, 100 scans were
used as a training set and the remaining as a sample set. Average spectra for the
training sample were obtained using both methods described in Chapter 5. The
Fourier transform based method can be summarized as follows:
6-2 Spectra Calculation
64
1. Take the Fourier transform of the OCT intensity-based image data matrix to
generate the backscattered spectrum
2. Take the autocorrelation of the backscattering spectra
The Burg’s based method can be summarized as follows:
1. Reshape OCT intensity-based image data matrix
2. Use Burg’s algorithm to obtained an AR-parameter estimate of the Power
Spectral Density (PSD) of the reshaped matrix and the corresponding frequencies
The Fourier transform based method, as expected, was more noisy and with
limited resolution, eventhough spectral differences can be seen in the spectra from
different phantoms. The full width magnitude around 75% is an indicator of the
scatterer size (Figure 6-2), a feature exploited by others in LSS [45, 46]. Burg’s
method results in characteristic spectral differences with much greater spectral resolution. Figure 6-2 we can noticed the different tilt for each phantom which is
predicted by Mie theory.
There is a significant amount of spectral modulation due to the random positioning of the scatterers inside the imaging volume, which makes difficult to distinguish
their scattering pattern. Theoretical studies have showned that spectral analysis
is possible with the assumption that the scattering from different locations has the
same general wavelength-dependent scattering profile [61]. Biologically this is true,
epithelium tissue demonstrates layered or regional structure where adjacent scatterers are more or less homogeneous. Also experimental observations agree that
is extremely difficult to distinguish the wavelength-dependent scattering pattern
based on only one SOCT measurement [11]. One simple method to resolve this
problem is to average the OCT signal around the scan of interest before obtaining
the backscattered spectrum. A sliding window was used for averaging the scans
adjacent to the scan of interest.
For spectral analysis a rectangular window of the OCT image for each phantom
was used with 3001(axial) x 25(transverse) pixel dimensions. Following the averaging each spectrum was normalized to the system spectrum. Although not necessary
for the analysis, the normalization accentuates the differences between the spectra
which makes studying the spectral variations easier. Also for the Fourier transform
based method, median filtering was applied in an effort to reduce the effects of the
high level of noise present.
6-2 Spectra Calculation
65
Normalized ACF
1
1 um
2 um
4 um
0.9
0.8
Ampl
0.7
0.6
0.5
0.4
0.3
0.2
1.3
1.305
1.31
lag
1.315
1.32
5
x 10
Figure 6.1: The autocorrelation of the backscattering spectra obtained with FFT
Normalized Average Training Spectra
1
1 um
2 um
0.9
4 um
0.8
0.7
Ampl
0.6
0.5
0.4
0.3
0.2
0.1
0
3.25
3.3
3.35
3.4
3.45 3.5
3.55
frequency(kHz)
3.6
3.65
3.7
3.75
4
x 10
Figure 6.2: Back-scattered normalized average power spectra obtained with Burg’s
method
6-3 Spectral Analysis
6-3
66
Spectral Analysis
The statistical analysis of the spectra of the above samples begins with a reduction
of the number of variables per sample. The reduction is achieved using Principal
Component Analysis (PCA), PCA can identifying patterns in data, and expressing
the data in such a way as to highlight their similarities and differences. All the
training groups for each phantom were merged to form a new training group data
ˆ tr
(AllP
xx ) for PCA, where each column corresponds to the training data from each
phantom. Then a so-called Feature Vector (FV) is constructed by their Principal
Components (PCs), FV is constructed by taking the first n PCs and forming a
matrix with these PCs in the columns
FV = (P C1 P C2 . . . P Cn )
(6.1)
FV with the first 35 PCs describes more than 99.99% of the training data group
variance. To obtain a new orthogonal basis for both the data groups with no
redundant information, the FV can be used by simply left multiplying both group
data.
P̂PxxCA tr = AllPtr
(6.2)
xx FV
The new transformed data have been expressed in terms of the PCs in FV and
the principal differences of the PCs between each phantom will be exploited for
scatterer size estimation and classification.
A classification scheme followed to determine in which category each sample
belongs. First one-way Multivariate Analysis of Variance (MANOVA) was applied
to the new training group P̂PxxCA tr to obtain new linear combinations of variables.
For comparing the multivariate means of the new training group P̂PxxCA tr a matrix
label as size matrix (d̂) was constructed and each column of d̂ is the size of each
phantom. Size matrix d̂ in MANOVA is defined as a categorical variable for grouping the new training data group. MANOVA have resulted in two eigenvectors on
which the data must be projected. A discriminant quadratic type function can then
be used which fits a multivariate normal density to each group, with covariance estimates stratified by group. From the discriminated data set, the sensitivity and
the specificity of our classification can be computed.
In addition to the classification scheme, an estimate of the diameter of the
microspheres was calculated. A linear dependence was assumed between particle
diameter and frequency and the new training data group P̂PxxCA tr were used to give
solution to the set of linear equations
 = P̂PxxCA
tr
−1
d̂
(6.3)
6-3 Spectral Analysis
67
Then coefficient matrix A can be used for scatterer size estimation with sample
training group
(6.4)
d̂ = P̂PsaCA Â
The success of the algorithm can be assessed by evaluating the mean error and the
standard deviation for all estimates.
Therefore, by studying the mean error and the classification results, the optimum settings for both techniques can be found. Figure 6.3(a) shows the relation
between the mean error estimate for the scatterer size versus the parameters p for
the AR model, there is no obvious relation between the mean erron and the AR
parameters N Burg. The mean error for the size estimation is minimized for AR
parameters N Burg = 120. Also, the classifications results Figure 6.3(b) versus
the AR parameters have the same behaviour and the best classification results is
for AR parameters N Burg = 120. The number of PCs for the construction FV
also affect the results, Figures 6.3(c)-(d) show mean error and classification results
for different numbers of PCs, for both cases the best results are for 35 PCs. The
same study was made for the Fourier based method, the number points of the FFT
(N F F T ) and the number of PCs were examined. From Figures 6.4(a)-(d), again 35
PCs are giving the best results and the points of the FFT are N F F T = 4096 = 212 .
The results from the whole process have been more than promising for Burg’s
method, has successfully classified the scatterers by category with sensitivity and
specificity more than 85% (Figure 6.5) and the mean error for the size estimation
was less than 17 %, (Table 6.1). Results from the Fourier based method, as expected
were quite poor due to the low resolution and the high level of noise, (Table 6.2).
A number of features were study, especially for the backscattered spectrum. For
studying the relationship between scatterer size and frequency the square power
and the derivative of the backscattered spectrum were used for PCA. Table 6.3 and
Table 6.4 are the results for the square power and the derivative respectively.There
is no improvement therefore a combination between the single power with the square
tr
power and the derivative. The dimension of matrix AllP̂xx is doubled, new columns
are added for the square power and the derivative respectively. The results now
are improve for the combination of single power and derivative (Table 6.5), the
other combination results were satisfying. Figure 6.5 are the MANOVA results for
the single power and the two combinations, single power with the derivative have
exactly the same results. Because an approximative derivative was used, which
essentially is the difference between adjacent elements of the single power of the
spectrum.
6-3 Spectral Analysis
68
Table 6.1: Results for Burg’s Method with window 3001 x 25 pixels and p = 120
AR parameters
D
Concentration Sensitivity Specificity Mean Error St. Dev.
(µm)
Sph/I.V.
(%)
(%)
(%)
Error(%)
1
50
100
100
2
5
100
100
16.44
19.10
4
2
85
100
Table 6.2: Results for FFT/Autocorrelation with window 3001 x 25 pixels and
N F F T = 212
D
Concentration Sensitivity Specificity Mean Error St. Dev.
(µm)
Sph/I.V.
(%)
(%)
(%)
Error(%)
1
50
75
82
2
5
69
82
65.35
48.52
4
2
62
82
D
(µm)
1
2
4
Table 6.3: Results for square power of the spectrum P̂Bu
tr
Concentration Sensitivity Specificity Mean Error St. Dev.
Sph/I.V.
(%)
(%)
(%)
Error(%)
50
100
97.50
5
94
97.50
32.83
32.71
2
95
97.50
D
(µm)
1
2
4
ˆ Bu
Table 6.4: Results for derivative of the spectrum dP
tr
Concentration Sensitivity Specificity Mean Error St. Dev.
Sph/I.V.
(%)
(%)
(%)
Error(%)
50
100
100
5
91
100
24.21
21.56
2
94
100
D
(µm)
1
2
4
Table 6.5: Results for single power and derivative of P̂Bu
tr
Concentration Sensitivity Specificity Mean Error St. Dev.
Sph/I.V.
(%)
(%)
(%)
Error(%)
50
100
100
5
100
100
15.79
18.31
2
85
100
Table 6.6: Results for single power and square power of P̂Bu
tr
D
Concentration Sensitivity Specificity Mean Error St. Dev.
(µm)
Sph/I.V.
(%)
(%)
(%)
Error(%)
1
50
100
100
2
5
82
100
16.81
15.96
4
2
93
100
6-3 Spectral Analysis
69
1 um (50)
2 um (5)
4 um (2)
Sensitivity Vs #NBurg
Mean Error Vs #NBurg
40
110
100
90
30
Sensitivity (%)
mean error (%)
35
25
80
70
20
60
15
50
20
40
60
80
100
120
140
160
20
40
60
80
# NBurg
(a)
120
140
160
(b)
1um (50)
2 um (5)
4 um (2)
Sensitivity Vs #PC
Mean Error Vs #PC
28
100
# NBurg
110
26
105
100
Sensitivity (%)
Sensitivity (%)
24
22
20
95
90
85
18
80
16
75
20
25
30
35
40
45
20
25
30
# PC
35
40
45
# PC
(c)
(d)
Figure 6.3: Plots of mean error and sensitivity versus N Burg (a), (c) and principal
components (b), (d) for Burg’s method
Mean error Vs Nfft
1 μm
2 μm
4 μm
Sensitivity Vs Nfft
90
68
80
70
66
Sensitivity (%)
Mean error (%)
60
64
62
50
40
30
60
20
58
10
0
2000
4000
6000
8000
0
10000 12000 14000 16000 18000
2000
4000
(a)
Sensitivity (%)
Mean Error (%)
72
70
68
66
64
35
40
# PC
(c)
10000 12000 14000 16000 18000
Sensitivity Vs #PC
Mean Error Vs #PC
30
8000
(b)
74
25
6000
# NFFT
# NFFT
45
50
1 um
2 um
4 um
76
74
72
70
68
66
64
62
60
58
56
54
52
50
48
46
44
42
25
30
35
40
45
50
# PC
(d)
Figure 6.4: Plots of mean error and sensitivity versus N F F T (a), (c) and principal
components (b), (d) for FFT/Autocorrelation
6-3 Spectral Analysis
70
Sample Data Scatter Plot
50
Training Data Scatter Plot
50
1
2
4
40
1
2
4
40
MANOVA Eigenvec 1
30
MANOVA Eigenvec 1
30
20
10
20
10
0
0
-10
-10
-20
-25
-20
-25
-20
-15
-10
MANOVA Eigenvec 2
-5
-20
-15
-10
MANOVA Eigenvec 2
0
-5
0
(a)
Training Data Scatter Plot
50
Sample Data Scatter Plot
1
2
4
40
50
1
2
4
40
30
MANOVA Eigenvec 1
MANOVA Eigenvec 1
30
20
10
20
10
0
0
-10
-10
-20
-25
-20
-15
-10
MANOVA Eigenvec 2
-5
-20
-25
0
-20
-15
-10
MANOVA Eigenvec 2
-5
0
(b)
Training Data Scatter Plot
Sample Data Scatter Plot
-20
-20
1
2
4
-40
-60
MANOVA Eigenvec 1
MANOVA Eigenvec 1
-60
-80
-100
-80
-100
-120
-120
-140
-140
-160
1
2
4
-40
-160
0
5
10
15
MANOVA Eigenvec 2
20
25
0
5
10
15
MANOVA Eigenvec 2
20
(c)
Figure 6.5: Results of the MANOVA analysis of 100 training and 100 sample
spectra for the single power (a), the single power and the derivative (b), the single
and square power (c). (The dark lines are the classification area borders resulting
from the discriminant analysis)
25
6-4 Depth Analysis
6-4
71
Depth Analysis
Previous studies on SOCT have used each pixel in the OCT image for calculating
local spectra [9, 11]. In our approach a part of the OCT image was only used, a
box window dimension of 3001(axial) x 25(transverse) pixel. By varying the axial
initial position of the window, a significant differences to the acquired spectrum
was noticed. Theoretical studies of the OCT signal have indicated that there is
a depth dependency for the OCT signal using the small-angle approximation of
the Radiative Transfer Equation (RTE) [77, 78] and the extended Huygens-Fresnel
principle [79]. In the literature only one work refers to the depth-dependency of the
SOCT spectrum by using theoretical studies based on Wigner distribution [80].
To verify the depth dependency for the SOCT spectra a number of OCT images
were acquired from the 2 µm phantom. The first step was to acquire images for
different focus locations, this can be achieved by varying the focus position of the
OCT system, and three images were acquired with the focus, midway and low in the
sample. Also three images were acquired by varying the position of the reference
mirror, for obtaining images with different pathlengths.
The spectra from the first OCT image where the focus was at the top of the sample are shown in Figure 6.6. Spectra were obtained from different depth positions
inside OCT image, in focus (s1) and below focus (s2). The spectral differences
indicate that, s2 is shifted and has different shape from s1. Depth analysis was
continued for the image where the focus was in the middle of the sample. The
corresponding spectra are s3, s4 and s5 for above, in and below focus (Figure 6.7).
Again there are significant spectral differences and also the amplitute of spectra is
different in this case. The third image was obtained with focus at the bottom of
the sample, with s6, s7 and s8 taken above, in and below focus (Figure 6.8). For
the images with different pathlengths the spectra also have different shape and they
are shifted (Figure 6.9).
Further analysis was done for comparing spectra from the OCT images. First,
spectra from the same depth (s2, s4 and s6) were compared, again spectral differences such as shift and different shape were observed (Figure 6.10). Also the in
focus spectra were compared (s1, s4 and s7), here the spectral differences are only
for the shape of the spectrum (Figure 6.11). It is clear that there is a depth dependency in the SOCT signals. Temporal coherence have been suggested for explaining
this phenomenon [80]. However, further study is required to understand what is
happening since it is hard to decouple focus and pathlength dependencies.
6-4 Depth Analysis
−8
7
72
Focus at the top of the sample
x 10
focus (s1)
below (s2)
6
5
4
3
2
1
0
24.5
25
25.5
26
Frequency (kHz)
26.5
27
27.5
Figure 6.6: Spectra from different depths with the focus at the top of the sample
−9
8
Focus in the middle of the sample
x 10
above s3
focus s4
below s5
7
6
5
4
3
2
1
0
24.5
25
25.5
26
26.5
Frequency (kHz)
27
27.5
Figure 6.7: Spectra from different depths with the focus at the middle of the sample
6-4 Depth Analysis
−9
8
73
Focus at the bottom of the sample
x 10
above s6
focus s7
below s8
7
6
5
4
3
2
1
0
24.5
25
25.5
26
Frequency (kHz)
26.5
27
27.5
Figure 6.8: Spectra from different depths with the focus at the bottom of the sample
Spectra with same pathlength and different depth
1
long (l)
medium (m)
small (s)
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
24.5
25
25.5
26
26.5
Frequency (kHz)
27
27.5
Figure 6.9: Spectra from below the surface at different pathlengths
6-4 Depth Analysis
74
Spectra from the same depth
1
focus above (s2)
in focus (s4)
focus below (s6)
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
24.5
25
25.5
26
26.5
Frequency (kHz)
27
27.5
Figure 6.10: Spectra from the same depth from three different OCT images
In Focus Spectra
1
top (s1)
middle (s4)
bottom (s7)
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
24.5
25
25.5
26
26.5
Frequency (kHz)
27
27.5
Figure 6.11: In focus spectra from three different OCT images
Chapter 7
Summary and Future Work
7-1
Summary
In this thesis a review was given for the background knowledge and important concepts of OCT, especially in the Time–Domain. A thorough study of Mie scattering
theory and its applications in SOCT was given, a Matlab source code was developed for calculating backscattered Mie spectra. A brief description of speckle and
its properties was given, and the prospect of SOCT. Relative small volume tissue
phantoms with high microspheres concentration have bee constructed for imaging purposes. Spectral analysis method for OCT signals were reviewed, based on
classic non–parametric methods such as Periodogram/FFT and based on autoregressive parametric methods such as Burg’s method. Based on the previous sections
an autoregressive spectral estimation was suggested for acquiring the backscattered
spectrum from the OCT signal. Then a PCA-based algorithm was used to extract
diagnostic features and to reduce the dimension of the spectra. A classification
scheme based on MANOVA followed to determine in which category each sample
belongs. In addition to the classification scheme, an estimate of the diameter of
the microspheres was calculated, based on the assumption that a linear dependence
between scatterer size and frequency exists.
7-2
Future Work
The technique developed in this thesis can result in an extremely valuable tool for
the investigation of disease tissue features which now remain below the resolution
of OCT. For this study a Time–Domain OCT system was used, implementation of
this technique in Fourier–Domain OCT systems, required to enable clinical studies.
FDOCT has higher resolution and high acquition rate for real time imaging. Clinical applications such as dissease imaging tissues will be a major step for OCT cancer
studies. Also different imaging scenarios should be investigated such as tissue phan-
7-2 Future Work
76
toms with different concentration, to estimate the concentration in addition to the
diameter of the phantom using a Mie based approach. Another issue that rises from
this study is the depth dependency of the SOCT spectra, an issue that demands
rigorous possibly analysis using joint Time–Frequency Distributions (jTFD). There
are many other potential future SOCT research directions and application areas
for this study. The additional information provided by SOCT increases our ability for tissue characterization, disease diagnosis, and other medical and biological
investigations.
Bibliography
[1] D. Huang, E. 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, 254(5035):1178–1181, 1991. 12, 15
[2] J. A. Izatt, M. R. Hee, E. A. Swanson, C. P. Lin, D. Huang, J. S. Schuman,
C. A. Puliafito, and J. G. Fujimoto. Micrometer-scale resolution imaging of
the anterior eye with optical coherence tomography. Arch. Ophthalmol., 112:
1584–1589, 1994. 13
[3] Wei Luo, Daniel L. Marks, Tyler S. Ralston, and Stephen A. Boppart. Threedimensional optical coherence tomography of the embryonic murine cardiovascular system. Journal of Biomedical Optics, 11(2):021014, 2006. 13
[4] J. Ik-Kyung, G. Tearney, B. MacNeill, M. Takano, F. Moselewski, N. Iftima,
M. Shishkov, S. Houser, T. Aretz, E. Halpern, and B. Bouma. In Vivo Characterization of Coronary Atherosclerotic Plaque by Use of Optical Coherence
Tomography. Circulation, 111(12):1551–1555, 2005. 13
[5] Wolfgang Drexler. Ultrahigh-resolution optical coherence tomography. Journal
of Biomedical Optics, 9(1):47–74, 2004. 13
[6] Chenyang Xu, Jian Ye, Daniel L. Marks, and Stephen A. Boppart. Nearinfrared dyes as contrast-enhancing agents for spectroscopic optical coherence
tomography. Opt. Lett., 29(14):1647–1649, 2004. 13, 14, 46
[7] B. Hyle Park, Mark C. Pierce, Barry Cense, and Johannes F. de Boer. Jones
matrix analysis for a polarization-sensitive optical coherence tomography system using fiber-optic components. Opt. Lett., 29(21):2512–2514, 2004. 13
[8] Adam M. Zysk, Freddy T. Nguyen, Amy L. Oldenburg, Daniel L. Marks,
and Stephen A. Boppart. Optical coherence tomography: a review of clinical
development from bench to bedside. Journal of Biomedical Optics, 12(5):
051403, 2007. 9, 13
BIBLIOGRAPHY
78
[9] U. Morgner, W. Drexler, F. X. Kärtner, X. D. Li, C. Pitris, E. P. Ippen, and
J. G. Fujimoto. Spectroscopic Optical Coherence Tomography. Opt. Lett., 25
(2):111–113, 2000. 13, 46, 71
[10] Dirk J. Faber, Egbert G. Mik, Maurice C. G. Aalders, and Ton G. van Leeuwen.
Light absorption of (oxy-)hemoglobin assessed by spectroscopic optical coherence tomography. Opt. Lett., 28(16):1436–1438, 2003. 14, 46
[11] Desmond Adler, Tony Ko, Paul Herz, and James Fujimoto. Optical coherence
tomography contrast enhancement using spectroscopic analysis with spectral
autocorrelation. Opt. Express, 12(22):5487–5501, 2004. 14, 46, 53, 64, 71
[12] Adam Wax, Changhuei Yang, Vadim Backman, Kamran Badizadegan,
Charles W. Boone, Ramachandra R. Dasari, and Michael S. Feld. Cellular Organization and Substructure Measured Using Angle-Resolved Low-Coherence
Interferometry. Biophys. J., 82(4):2256–2264, 2002. 14, 46, 53
[13] Robert C. Youngquist, Sally Carr, and D. E. N. Davies. Optical coherencedomain reflectometry: a new optical evaluation technique. Opt. Lett., 12(3):
158, 1987. 15
[14] K. Takada, I. Yokohama, K. Chida, and J. Noda. New measurement system
for fault location in optical waveguide devices based on an interferometric
technique. Appl. Opt., 26(9):1603, 1987. 15
[15] H. H. Gilgen, R. P. Novak, R. P. Salathe, W. Hodel, and P. Beaud. Submillimeter optical reflectometry. Journal of Lightwave Technology, 7:1225–1233,
1989. 15
[16] A. F. Fercher, K. Mengedoht, and W. Werner. Eye-length measurement by
interferometry with partially coherent light. Opt. Lett., 13(3):186, 1988. 15
[17] Christoph K. Hitzenberger. Measurement of corneal thickness by low-coherence
interferometry. Appl. Opt., 31(31):6637, 1992. 15
[18] X. Clivaz, F. Marquis-Weible, R. P. Salathe, R. P. Novak, and H. H. Gilgen.
High resolution reflectometry in biological tissues. Opt. Lett., 17(1):4, 1992.
15
[19] J. M. Schmitt, A. Knuttel, and R. F. Bonner. Measurement of optical properties of biological tissues by low-coherence reflectometry. Appl. Opt., 32(30):
6032, 1993. 15
[20] A. Fercher, C. Hitzenberger, W. Drexler, G. Kamp, and H. Sattmann. In vivo
optical coherence tomography. Amer. J. Ophthalmol, 116:113–114, 1993. 15
BIBLIOGRAPHY
79
[21] J. G. Fujimoto, M. E. Brezinski, G. J. Tearney, S. A. Boppart, B. E. Bouma,
M. R. Hee, J. F. Southern, and E. A. Swanson. Optical biopsy and imaging
using optical coherence tomography. Nature Med., 1:970–972, 1995. 15, 24
[22] J. M. Schmitt, A. Knuttel, M. Yadlowsky, and R. F. Bonner. Optical coherence
tomography of a dense tissue: statistics of attenuation and backscattering.
Phys. Med. Biol., 42:1427–1439, 1994. 15, 24
[23] B. E. Bouma and G. J. Tearney, editors. Handbook of Optical Coherence Tomography. Marcel Dekker, New York, 2001. 16, 25
[24] Max Born and Emil Wolf. Principles of Optics. Cambridge University Press,
Cambridge, seventh edition, 1999. 16
[25] Eugene Hecht. Optics. Addison Wesley, New York, fourth edition, 2001. 16
[26] J. W. Goodman. Statistical Optics. Wiley Classics Library. John Wiley and
Sons, New York, 1985. 20
[27] Joseph M. Schmitt and Gitesh Kumar. Optical Scattering Properties of Soft
Tissue: A Discrete Particle Model. Appl. Opt., 37(13):2788–2797, 1998. 24
[28] Brett E. Bouma, Lynn E. Nelson, Guillermo J. Tearney, David J. Jones,
Mark E. Brezinski, and James G. Fujimoto. Optical Coherence Tomographic
Imaging of Human Tissue at 1.55µm and 1.81µm Using Er- and Tm-Doped
Fiber Sources. Journal of Biomedical Optics, 3(1):76–79, 1998. 24, 25
[29] Parwane Parsa, Steven L. Jacques, and Norman S. Nishioka. Optical properties
of rat liver between 350 and 2200 nm. Appl. Opt., 28(12):2325, 1989. 24
[30] Jean-Luc Boulnois. Photophysical processes in recent medical laser developments: A review. Lasers in Medical Science, 1(1):47–66, 1986. 9, 24
[31] C. K. Hitzenberger, A. Baumgartner, and A. F. Fercher. Dispersion induced
multiple signal peak splitting in partial coherence interferometry. Optics Communications, 154:179–185, 1998. 24
[32] M. Kempe, A. Thon, and W. Rudolph. Resolution limits of microscopy through
scattering layers. Optics Communications, 10:492–496, 1994. 24
[33] E.A. Chinn, S.R.; Swanson. Blindness limitations in optical coherence domain
reflectometry. Electronics Letters, 29(23):2025–2027, 1993. 25
[34] J.M. Schmitt. Optical coherence tomography (OCT): a review. Selected Topics
in Quantum Electronics, IEEE Journal of, 5(4):1205–1215, 1999. 11, 25, 26
BIBLIOGRAPHY
80
[35] D. J. Derickson, P. A. Beck, T. L. Bagwell, D. M. Braun, J. E. Fouquet, F. G.
Kellert, M. J. Ludowise, W. H. Perez, T. R. Ranganath, G. R. Trott, , and
S. R. Sloan. High-power, low-internal-reflection, edge emitting light-emitting
diodes. Hewlett-Packard J., 46:43–49, 1995. 25
[36] C. F. Lin and B. L. Lee. Extremely broadband AlGaAs/GaAs superluminescent diodes. Appl. Phys. Lett., 71:1598–1600, 1997. 25, 26
[37] P. J. Poole, M. Davies, M. Dion, Y. Feng, S. Charbonneau, R. D. Goldberg,
and I. V. Mitchell. The fabrication of a broad-spectrum light-emitting diode
using high-energy ion implantation. IEEE Photon. Technol. Lett., 8:1145–1147,
1996. 25, 26
[38] Hsiao-Hua Liu, Po-Hsiu Cheng, and Jyhpyng Wang. Spatially coherent whitelight interferometer based on a point fluorescent source. Opt. Lett., 18(9):678,
1993. 25
[39] B. Bouma, G. J. Tearney, S. A. Boppart, M. R. Hee, M. E. Brezinski, and
J. G. Fujimoto. High-resolution optical coherence tomographic imaging using
a mode-locked T i : Al2 O3 laser source. Opt. Lett., 20(13):1486, 1995. 25
[40] G. J. Tearney, M. E. Brezinski, B. E. Bouma, S. A. Boppart, C. Pitris, J. F.
Southern, and J. G. Fujimoto. In vivo endoscopic optical biopsy with optical
coherence tomography. Science, 276:2037–2039, 1997. 25
[41] R. Paschotta, J. Nilsson, A. C. Tropper, and D. C. Hanna. Efficient superfluorescent light sources with broad bandwidth. Selected Topics in Quantum
Electronics, IEEE Journal of, 3(4):1097–1099, 1997. 25
[42] Jon Y. Wang. Detection efficiency of coherent optical radar. Appl. Opt., 23
(19):3421, 1984. 26
[43] J. M. Schmitt, S. L. Lee, and K. M. Yung. An optical coherence microscope
with enhanced resolving power in thick tissue. Optics Communications, 142:
203–207, 1997. 26
[44] G. Mie. Beitrage zur optik triiber medien speziell kolloidaler metallosungen.
Ann. Phys., 25(2):377–445, 1908. 27
[45] L. T. Perelman, V. Backman, M. Wallace, G. Zonios, R. Manoharan, A. Nusrat,
S. Shields, M. Seiler, C. Lima, T. Hamano, I. Itzkan, J. Van Dam, J. M.
Crawford, and M. S. Feld. Observation of periodic fine structure in reflectance
from biological tissue: a new technique for measuring nuclear size distribution.
Phys. Rev. Lett., 80(3):627–630, 1998. 27, 53, 64
BIBLIOGRAPHY
81
[46] V. Backman, V. Gopal, M. Kalashnikov, K. Badizadegan, R. Gurjar, A. Wax,
I. Georgakoudi, M. Mueller, C. Boone, R. Dasari, and M. Feld. Measuring cellular structure at submicrometer scale with light scattering spectroscopy. Selected
Topics in Quantum Electronics, IEEE Journal of, 7(6):887–893, Nov/Dec 2001.
ISSN 1077-260X. doi: 10.1109/2944.983289. 27, 64
[47] M. Wallace, L. T. Perelman, V. Backman, J. M. Crawford, M. Fitzmaurice,
M. Seiler, K. Badizadegan, S. Shields, I. Itzkan, R. R. Dasari, J. Van Dam,
and M. S. Feld. Endoscopic detection of dysplasia in patients with Barrett’s
esophagus using light-scattering spectroscopy. Gastroenterology, 119(3):677–
682, 2000. 27
[48] V. Backman, M. B. Wallace, L. T. Perelman, J. T. Arendt, R. Gurjar, M. G.
Muller, Q. Zhang, G. Zonios, E. Kline, T. McGillican, S. Shapshay, T. Valdez,
K. Badizadegan, J. M. Crawford, M. Fitzmaurice, S. Kabani, H. S. Levin,
M. Seiler, R. R. Dasari, I. Itzkan, J. Van Dam, and M. S. Feld. Detection of
preinvasive cancer cells. Nature, 406:35–36, 2000. 27
[49] V. Backman, R. Gurjar, K. Badizadegan, I. Itzkan, R. Dasari, L. Perelman, and
M. Feld. Polarized light scattering spectroscopy for quantitative measurement
of epithelial cellular structures in situ. Selected Topics in Quantum Electronics,
IEEE Journal of, 5(4):1019–1026, 1999. 27
[50] Adam Wax, Changhuei Yang, and Joseph A. Izatt. Fourier-domain lowcoherence interferometry for light-scattering spectroscopy. Opt. Lett., (14):
1230–1232. 27, 46
[51] Adam Wax, Changhuei Yang, Vadim Backman, Maxim Kalashnikov, Ramachandra R. Dasari, and Michael S. Feld. Determination of particle size
by using the angular distribution of backscattered light as measured with lowcoherence interferometry. J. Opt. Soc. Am. A, 19(4):737–744, 2002. 27, 46
[52] J. Beuthan, O. Minet, J. Helfmann, M. Herrig, and G. Muller. The spatial
variation of the refractive index in biological cells. Phys. Med. Biol., 41(3):
369–382, 1996. 27, 45
[53] P. Sloot, A. Hoekstra, and C. Figdor. Osmotic response of lymphocytes measured by means of forward light scattering: theoretical considerations. Cytometry, 9(6):636–641, 1988. 27, 45
[54] C. Bohren and D. Huffman. Absorption and Scattering of Light by Small
Particles. Wiley Science Paperback Series. John Wiley and Sons, New York,
1998. 27, 37, 45, 47
BIBLIOGRAPHY
82
[55] H. C. Van de Hulst. Light Scattering by Small Particles. Dover Publications,
New York, 1981. 27, 37, 47
[56] G. B. Arfken and H. J. Weber. Mathematical Methods for Physicists. Academic
Press, 2005. 33
[57] A. L. Aden. Electromagnetic scattering from spheres with sizes comparable to
the wavelength. J. Appl. Phys., 22:601–605, 1951. 44
[58] W. J. Wiscombe. Improved Mie scattering algorithms. Appl. Opt., 19(9):1505,
1980. 44
[59] B. Hermann, K. Bizheva, A. Unterhuber, B. Považay, H. Sattmann,
L. Schmetterer, A. Fercher, and W. Drexler. Precision of extracting absorption
profiles from weakly scattering media with spectroscopic timedomain optical
coherence tomography. Opt. Express, (8):1677–1688. 46
[60] C. Xu, P. S. Carney, W. Tan, and S. A. Boppart. Light-scattering spectroscopic
optical coherence tomography for differentiating cells in 3D cell culture. Proc.
of SPIE, 6088(608804):1–10, 2006. 46, 53
[61] Chenyang Xu, P. Carney, and Stephen Boppart. Wavelength-dependent scattering in spectroscopic optical coherence tomography. Opt. Express, 13(14):
5450–5462, 2005. 46, 64
[62] J. M. Schmitt, S. H. Xiang, and K. M. Yung. Speckle in optical coherence
tomography. Journal of Biomedical Optics, 4:99–105, 1999. 48
[63] JD Briers and AF Fercher. Retinal blood-flow visualization by means of laser
speckle photography. Invest. Ophthalmol. Vis. Sci., 22(2):255–259, 1982. 48
[64] Hitoshi Fujii, Toshimitsu Asakura, Kunihiko Nohira, Yoshihisa Shintomi, and
Takehiko Ohura. Blood flow observed by time-varying laser speckle. Opt. Lett.,
10(3):104, 1985. 48
[65] Michael D. Stern. Laser doppler velocimetry in blood and multiply scattering
fluids: theory. Appl. Opt., 24(13):1968, 1985. 48
[66] J. W. Goodman. Statistical properties of laser speckle pattern. Applied Physics,
9, 1984. 49
[67] Lester I. Goldfischer. Autocorrelation function and power spectral density of
laser-produced speckle patterns. J. Opt. Soc. Am., 55(3):247, 1965. 49
BIBLIOGRAPHY
83
[68] G. J. Tearney, M. E. Brezinski, J. F. Southern, B. E. Bouma, M. R. Hee,
and J. G. Fujimoto. Determination of the refractive index of highly scattering
human tissue by optical coherence tomography. Opt. Lett., 20(21):2258, 1995.
52
[69] R. Leitgeb, M. Wojtkowski, A. Kowalczyk, C. K. Hitzenberger, M. Sticker, and
A. F. Fercher. Spectral measurement of absorption by spectroscopic frequencydomain optical coherence tomography. Opt. Lett., 25(11):820–822, 2000. 53
[70] John Proakis and Dimitris Manolakis. Digital Signal Processing, Principles,
Algorithms and Applications. Prentice Hall, 2006. 55
[71] P. Chaturvedi and M. Insana. Autoregressive Spectral Estimation in Ultrasonic
Scatterer Size Imaging. Ultrasonic Imaging, 18:10–24(15), 1996. 57
[72] K.A. Wear, R.F. Wagner, and B.S. Garra. A comparison of autoregressive spectral estimation algorithms and order determination methods in ultrasonic tissue characterization. Ultrasonics, Ferroelectrics and Frequency Control, IEEE
Transactions on, 42(4):709–716, Jul 1995. 57
[73] Petre Stoica and Randolph L. Moses. Introduction to Spectral Analysis. Prentice Hall, 1997. 57
[74] J. P. Burg. Maximum Entropy Spectral Analysis. PhD thesis, Stanford University, 1975. 58
[75] I. T. Jolliffe. Principal Component Analysis. Springer-Verlag, second edition,
2002. 60
[76] J. E Jackson. A User’s Guide to Principal Components. Wiley Series in Probability and Statistics. John Wiley and Sons, New York, 2003. 61
[77] L. S. Dolin. A theory of optical coherence tomography. Radiophysics and
Quantum Electronics, 41(10):850–873, 1998. 71
[78] Ilya V. Turchin, Ekaterina A. Sergeeva, Lev S. Dolin, Vladislav A. Kamensky,
Natalia M. Shakhova, and Rebecca Richards-Kortum. Novel algorithm of processing optical coherence tomography images for differentiation of biological
tissue pathologies. Journal of Biomedical Optics, 10(6):064024, 2005. 71
[79] Lars Thrane, Harold T. Yura, and Peter E. Andersen. Analysis of optical coherence tomography systems based on the extended huygens–fresnel principle.
J. Opt. Soc. Am. A, 17(3):484–490, 2000. 71
BIBLIOGRAPHY
84
[80] Robert N. Graf and Adam Wax. Temporal coherence and time-frequency distributions in spectroscopic optical coherence tomography. J. Opt. Soc. Am. A,
24(8):2186–2195, 2007. 71