fourier-transform infrared spectroscopic imaging of

127
University of South Florida Scholar Commons Graduate eses and Dissertations Graduate School 5-20-2003 Fourier-Transform Infrared Spectroscopic Imaging of Prostate Histopathology Daniel Celestino Fernandez University of South Florida Follow this and additional works at: hps://scholarcommons.usf.edu/etd Part of the American Studies Commons is Dissertation is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion in Graduate eses and Dissertations by an authorized administrator of Scholar Commons. For more information, please contact [email protected]. Scholar Commons Citation Fernandez, Daniel Celestino, "Fourier-Transform Infrared Spectroscopic Imaging of Prostate Histopathology" (2003). Graduate eses and Dissertations. hps://scholarcommons.usf.edu/etd/1366

Upload: others

Post on 14-Nov-2021

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Fourier-Transform Infrared Spectroscopic Imaging of

University of South FloridaScholar Commons

Graduate Theses and Dissertations Graduate School

5-20-2003

Fourier-Transform Infrared Spectroscopic Imagingof Prostate HistopathologyDaniel Celestino FernandezUniversity of South Florida

Follow this and additional works at: https://scholarcommons.usf.edu/etdPart of the American Studies Commons

This Dissertation is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion inGraduate Theses and Dissertations by an authorized administrator of Scholar Commons. For more information, please [email protected].

Scholar Commons CitationFernandez, Daniel Celestino, "Fourier-Transform Infrared Spectroscopic Imaging of Prostate Histopathology" (2003). Graduate Thesesand Dissertations.https://scholarcommons.usf.edu/etd/1366

Page 2: Fourier-Transform Infrared Spectroscopic Imaging of

Fourier-Transform Infrared Spectroscopic Imaging of Prostate Histopathology

by

Daniel Celestino Fernandez

A dissertation submitted in partial fulfillment of the requirements for the degree of

Doctor of Philosophy Department of Pathology and Laboratory Medicine

College of Medicine University of South Florida

Co-Major Professor: Santo V. Nicosia, M.D. Co-Major Professor: Ira W. Levin, Ph.D.

Wenlong Bai, Ph.D. Luis H. Garcia-Rubio, Ph.D.

Maria Kallergi, Ph.D. Patricia A. Kruk, Ph.D.

Date of Approval: May 20, 2003

Keywords: FT-IR, adenocarcinoma, vibrational, spectroscopy, classification

© Copyright 2004 , Daniel Celestino Fernandez

Page 3: Fourier-Transform Infrared Spectroscopic Imaging of

Dedication

To my parents, for their love and support,

and

my wife, for always believing in me.

Page 4: Fourier-Transform Infrared Spectroscopic Imaging of

Acknowledgements

• Howard Hughes Medical Institute – National Institutes of Health Research

Scholars Program

• National Institutes of Health Graduate Partnership Program

• University of South Florida – College of Medicine – Department of Pathology

and Laboratory Medicine

• National Institute of Diabetes, Digestive and Kidney Diseases

• Ira W. Levin, Ph.D.

• Santo V. Nicosia, M.D.

• Stephen M. Hewitt, M.D., Ph.D.

• Rohit Bhargava, Ph.D.

• Michael D. Schaeberle, Ph.D.

• Scott W. Huffman, Ph.D.

• Patricia McCarthy, Ph.D.

• Jamie Winderbaum Fernandez, M.D.

Page 5: Fourier-Transform Infrared Spectroscopic Imaging of

i

Table of Contents

List of Tables ..................................................................................................................... iv

List of Figures ......................................................................................................................v

Abstract ............................................................................................................................. vii

Chapter One - Introduction ..................................................................................................1

1.1 Electromagnetic Spectrum......................................................................................... 1 1.1.1 Interactions of Electromagnetic Radiation with Matter.............................. 4

1.2 Basis of Infrared Absorption...................................................................................... 5 1.2.1 Requirements for IR Absorption................................................................. 6 1.2.2 Number of Vibrational Modes .................................................................... 8 1.2.3 Group Frequencies ...................................................................................... 9

1.3 IR Spectral Feature of Tissues ................................................................................. 10 1.3.1 Proteins ..................................................................................................... 10 1.3.2 Carbohydrates ........................................................................................... 15 1.3.3 Lipids ........................................................................................................ 15 1.3.4 Nucleic Acids............................................................................................ 17

1.4 FTIR Spectroscopy Background.............................................................................. 17 1.4.1 FTIR Spectrometers .................................................................................. 18 1.4.2 Infrared Microscopy.................................................................................. 20 1.4.3 Mapping with Single-Point Detectors....................................................... 22 1.4.4 Raster-scan Imaging Using Multichannel Detectors ................................ 24 1.4.5 Global FTIR Spectroscopic Imaging ........................................................ 25

1.5 Spectroscopic Imaging: Data Structure and Applications ...................................... 27 1.5.1 Image Classification Methods................................................................... 30

1.6 Prostate Background ................................................................................................ 31 1.6.1 Anatomy and Histology ............................................................................ 31 1.6.2 Prostate Pathology .................................................................................... 33

Chapter Two - Methods .....................................................................................................44

2.1 Tissue microarrays ................................................................................................... 44 2.1.1 Construction of Prostate Tissue Microarrays............................................ 44 2.1.2 Array P-16 Design .................................................................................... 45 2.1.3 Array P-40 Design .................................................................................... 45 2.1.4 Array P-80 Design .................................................................................... 46

Page 6: Fourier-Transform Infrared Spectroscopic Imaging of

ii

2.2 Tissue Array Section preparation............................................................................. 47 2.2.1 Optical Substrates for Tissue Array Sections ........................................... 47 2.2.2 Deparaffinization ...................................................................................... 48 2.2.3 Optical imaging of H&E sections ............................................................. 48

2.3 Spectroscopic Imaging Instrumentation .................................................................. 49 2.3.1 Tissue Array FT-IR Data Collection Parameters...................................... 51 2.3.2 Modifications and Environmental Considerations.................................... 52

2.4 Data Handling and Computational Considerations.................................................. 53 2.4.1 Data Pre-Processing.................................................................................. 53 2.4.2 Spectral Baseline Correction..................................................................... 54

Chapter Three - Infrared Spectroscopic Histology of Prostate..........................................56

3.1 Visualization of Spectral Images and Verification of Histologic Features.............. 56

3.2 Creation of Ground Truth Data Regions of Interest ................................................ 58

3.3 Spectral analysis of histologic features and metric selection................................... 62

3.4 Construction of a Supervised Classification Model for Prostate Histology ............ 64 3.4.1 Spectral Data Reduction ........................................................................... 64 3.4.2 Image Classification................................................................................. 66 3.4.3 Array P-16, 20-metric, GML Self-classification results........................... 68 3.4.4 Leave-one-out metric evaluation .............................................................. 70 3.4.5 Array P-16, 18-metric GML Classification Results ................................. 72

3.5 Validation of Prostate Histology Classification Model ........................................... 75 3.5.1 Cross-Array Validation............................................................................. 76

3.6 Conclusions and Further Directions......................................................................... 79

Chapter Four - Infrared Spectroscopic Histopathology of Prostate...................................81

4.1 Classification strategy.............................................................................................. 81

4.2 Array P-80 H&E Stained Section Pathology Analysis ............................................ 82

4.3 Array P-80 Histology Classification Results ........................................................... 83 4.3.1 Spatial Filtering of Histology Classification Results................................ 84

4.4 Construction of a Supervised Classification Model for Prostate Pathology............ 88 4.4.1 Creation of pathology ground truth ROIs ................................................. 88 4.4.2 Pathology Spectral Data Reduction .......................................................... 89 4.4.3 Histogram analysis of Spectral Metric Data ............................................. 91 4.4.4 Mean-centering of epithelial metric data. ................................................. 92 4.4.5 Metric Statistical Analysis ........................................................................ 92 4.4.6 GML Pathology Classification of Array P-80 .......................................... 94

4.5 Individual Patient Evaluation of P-80 Pathology Classification.............................. 96

4.6 Cross-Array Validation............................................................................................ 97

Page 7: Fourier-Transform Infrared Spectroscopic Imaging of

iii

4.7 Conclusions and Further Directions......................................................................... 98

References........................................................................................................................101 About the Author ................................................................................................... End Page

Page 8: Fourier-Transform Infrared Spectroscopic Imaging of

iv

List of Tables

Table 1.1 - Spectroscopic techniques utilizing different regions of the electromagnetic spectrum.............................................................................. 5

Table 1.2 - Staging of primary tumor (T) ......................................................................... 42

Table 1.3 - Staging of regional lymph node involvement (N).......................................... 43

Table 2.1 - Spectral frequencies used for spectroscopic baseline correction ................... 55

Table 3.1 - Histologic class population data..................................................................... 61

Table 3.2 - Histology Spectral Metric Definitions............................................................ 66

Table 3.3 - Error Matrix of supervised GML Classification results using 20 spectroscopic metrics .................................................................................. 69

Table 3.4 - Confusion matrix of supervised GML Classification attempt using 18 spectroscopic metrics .................................................................................. 72

Table 3.5 - Revised 6-class histology ground truth ROIs for Array P-16 and Array P-40 ............................................................................................................. 77

Table 3.6 - Error Matrix for 6-Class, GML Classification Results .................................. 78

Table 4.1 - Pathology spectral metric parameters............................................................. 90

Table 4.2 - Results of t-test on mean adenocarcinoma metric values from population of 25 patients on array P-80 for 54 candidate pathology metrics ......................................................................................................... 93

Table 4.3 - Error matrix for 20-metric pathology GML classification of epithelial tissue on array P-80 ..................................................................................... 96

Page 9: Fourier-Transform Infrared Spectroscopic Imaging of

v

List of Figures

Figure 1.1 - The electromagnetic spectrum ........................................................................ 2

Figure 1.2 - The infrared region of the electromagnetic spectrum ..................................... 4

Figure 1.3 - Vibrational modes and IR activity of water vapor (A) and carbon dioxide (B) molecules ................................................................................... 8

Figure 1.4 - Vibrational modes of methylene group........................................................... 9

Figure 1.5 - Structure of a typical amino acid .................................................................. 11

Figure 1.6 - Basic polypeptide structure ........................................................................... 12

Figure 1.7 - Common Protein Secondary Structures: α-helix and β–sheet....................... 13

Figure 1.8 - Michelson Interferometer.............................................................................. 19

Figure 1.9 - Three Instrumental Approaches for collection of spatially resolved FTIR spectroscopic data.............................................................................. 22

Figure 1.10 - Schematic representation of the image cube............................................... 28

Figure 1.11 - Zonal Anatomy of the Prostate ................................................................... 32

Figure 2.1 - Array P-80 Layout......................................................................................... 47

Figure 2.2 - Spectrum Spotlight 300 Microscope Optical Configuration......................... 51

Figure 3.1- A) Baseline-corrected N-H stretching (3290cm-1) absorbance intensity image of four tissue array spots from a single patient on Array P-16 B) Optical images of corresponding H&E stained section.......................................................................................................... 57

Figure 3.2 - Absorbance Band Ratio Images of tissue array spots from Array P-16 ....... 60

Figure 3.3 - Histologic class mean spectra ....................................................................... 63

Figure 3.4- Histograms of metric value class frequency distribution for the three most populated classes (epithelium, mixed stroma, & fibrous stroma) for: A) Metric 02 (band ratio 1080/1544cm-1), and B) Metric 11 (band ratio 1400/1390 cm-1) ...................................................... 67

Page 10: Fourier-Transform Infrared Spectroscopic Imaging of

vi

Figure 3.5 - Graphical Representation of results of the leave-one-out analysis ............... 71

Figure 3.6 - Classification results for 2 tissue array spots from the same patient ............ 73

Figure 4.1- Array P-80 histology classification results .................................................... 83

Figure 4.2 - Optical images of H&E stained section of Array P-80 ................................. 84

Figure 4.3 - Spatial filtering techniques for classified image results................................ 86

Figure 4.4 - Sieve operation spatial filtering of histology classfication results for patient 2 from array P-80 ............................................................................ 88

Figure 4.5 - Array P-80 pathology ground truth ROIs...................................................... 89

Figure 4.6 - Patient-to-patient metric variation................................................................. 91

Figure 4.7 - Array P-80 pathology classification results .................................................. 95

Figure 4.8 - Individual patient analysis of 20-metric GML pathology classification....... 97

Page 11: Fourier-Transform Infrared Spectroscopic Imaging of

vii

Abstract

Fourier-Transform Infrared Spectroscopic Imaging of Prostate Histopathology

Daniel Celestino Fernandez

ABSTRACT

Vibrational spectroscopic imaging techniques have emerged as powerful methods

of obtaining sensitive spatially resolved molecular information from microscopic

samples. The data obtained from such techniques reflect the intrinsic molecular

chemistry of the sample and in particular yield a wealth of information regarding

functional groups which comprise the majority of important molecules found in cells and

tissue. These spectroscopic imaging techniques also have the advantage of acquisition of

large numbers of spectral measurements which allow statistical analysis of spectral

features which are characteristic of the normal histological state as well as different

pathologic disease states. Databases of large numbers of samples can be acquired and

used to build model systems that can be used to predict spatial properties of unknown

samples.

The successful construction and application of such a model system relies on the

ability to compile high-quality spectral database information on a large number of

samples with minimal sample-to-sample preparation artifact. Tissue microarrays provide

a consistent sample preparation for high-throughput infrared spectroscopic profiling of

histologic specimens. Tissue arrays consisting of representative normal healthy prostate

tissue as well as pathologic entities including prostatitis, benign prostatic hypertrophy,

Page 12: Fourier-Transform Infrared Spectroscopic Imaging of

viii

and prostatic adenocarcinoma were constructed and used as sample populations for

infrared spectroscopic imaging at high spatial and spectral resolutions.

Histological and pathological features of the imaged tissue were correlated with

consecutive tissue sections stained with standard histologic stains and visualized via

traditional optical microscopy and reviewed with a trained pathologist. Spectral analysis

of histologic class mean spectra and subsequent cross-sample statistical validation were

used to classify reliable spectral metrics for class discrimination. Multivariate Gaussian

maximum likelihood classification algorithms were used to reliably classify all pixels in

an image scene to one of six different histologic subclasses: epithelium, smooth muscular

stroma, fibrous stroma, corpora amylacea, lymphocytic infiltration, and blood. The

developed database-dependent classification methods were used as a tool to investigate

subsequent microarrays designed with both normal epithelial tissue as well as

adenocarcinoma from a large population of patients. Such investigation led to the

identification of spectral features that proved useful in the preliminary discrimination of

benign and malignant prostatic epithelial tissue.

Page 13: Fourier-Transform Infrared Spectroscopic Imaging of

1

Chapter One - Introduction

Spectroscopy deals with the interaction of various forms of electromagnetic (EM)

radiation with matter. Vibrational spectroscopy provides information regarding the

molecular composition and structure of a wide range of materials including biological

tissues. Recent technological advances have led to powerful vibrational imaging

approaches involving both near and mid-infrared, as well as Raman-based platforms

providing spatially-resolved chemical information on a microscopic scale[1]. Infrared

spectroscopic imaging microscopy, in particular, benefits from many decades of

instrumentation advances and database compilations. A brief background into the theory

and techniques of infrared spectroscopy follows.

1.1 Electromagnetic Spectrum

The wave nature of electromagnetic (EM) radiation treats the radiation in terms of

oscillating electric and magnetic fields perpendicular to one another and to the direction

of wave propagation traveling with the velocity of light. Certain continuous regions of

the EM spectrum have been designated and appear in Figure 1.1[2]. Vibrational

absorption spectra result from the interaction of oscillating dipole moments, which occur

during molecular vibrations, with the electric field of the radiation, resulting in an energy

exchange between the radiation and the molecular system.

Electromagnetic radiation is characterized by its wavelength λ. The specific units

typically used to express wavelength vary across the spectrum from angstroms (Å) in the

Page 14: Fourier-Transform Infrared Spectroscopic Imaging of

gamma ray region to meters in the radio wave region or ~10-10 to 102 cm, respectively.

The units of µm are practical for describing radiation in the mid-infrared spectral region.

In the near-infrared (NIR) region the unit nm typically employed just as it is in the visible

(VIS) and ultraviolet (UV) spectral regions.

Ultraviolet Microwave Radio Waves

Visible

Infrared

MidNear Far

X-rays

γ-rays&

cosmic rays

.05Å 10nm 350nm 770nm 2.5µm 50µm 1mm 300mm

wavelength (λ)

frequency(ν) , energy (E)

Visible

Infrared

MidNear Far

X-rays

γ-rays&

cosmic rays

.05Å 10nm 350nm 770nm 2.5µm 50µm 1mm 300mm

wavelength (λ)

frequency(ν) , energy (E)

VisibleUltraviolet Microwave Radio Waves

Visible

Infrared

MidNear Far

X-rays

γ-rays&

cosmic rays

.05Å 10nm 350nm 770nm 2.5µm 50µm 1mm 300mm

wavelength (λ)

frequency(ν) , energy (E)

Ultraviolet Microwave Radio Waves

Visible

Infrared

MidNear Far

X-rays

γ-rays&

cosmic rays

.05Å 10nm 350nm 770nm 2.5µm 50µm 1mm 300mm

wavelength (λ)

frequency(ν) , energy (E)

Visible

Infrared

MidNear Far

X-rays

γ-rays&

cosmic rays

.05Å 10nm 350nm 770nm 2.5µm 50µm 1mm 300mm

wavelength (λ)

frequency(ν) , energy (E)

Visible

Figure 1.1 - The electromagnetic spectrum

Electromagnetic radiation can also be characterized by its frequency ν, defined as

the number of oscillations of the magnetic or electric field radiation vector per unit of

time[2]. The frequency unit is s-1 (oscillations per second), often specified in Hertz (Hz).

The energy (E) of EM radiation is directly related to its frequency (ν) by the equation

νhE = (1.1)

where h is Plank’s constant with a value h = 6.63 · 10-34 J s.

The frequency and wavelength (λ) of EM radiation are related by the proportionality

constant c (the speed of light) according to the equation

λν c= (1.2)

2

Page 15: Fourier-Transform Infrared Spectroscopic Imaging of

where c has a value of ~2.99793 · 1010 cm s-1 (in a vacuum).

Infrared spectroscopists have adopted the convention of expressing frequency in

terms of wavenumber with the units of cm-1[3]. A simple expression for wavenumber is

given by

λν 1= (1.3)

The units of wavenumber provide a convenient scale for IR spectroscopy,

especially the mid-infrared region that spans 200-4000 cm-1. The units of wavenumber

are also desirable for IR spectroscopists because they are directly proportional to the

energy of radiation, which varies inversely with wavelength as described by equation

1.4[4].

λhcE = (1.4)

The relationships between energy, frequency, and wavelength and the various

regions of the electromagnetic spectrum are detailed in figure 1.1. The infrared region of

the electromagnetic spectrum is subdivided into three contiguous regions; the near, mid

and far infrared regions. The nomenclature of these prefixes refers to the individual sub-

region’s position relative to the visible region. Figure 1.2 shows these three regions of

the infrared spectrum and the ranges they occupy on the wavelength, frequency and

wavenumber scales.

3

Page 16: Fourier-Transform Infrared Spectroscopic Imaging of

12900 4000 200 10

3.9· 1014 1.2· 1014 6.0· 1012 3.0· 1011

0.77 2.50 50 1000

InfraredNear Mid Far

wavelength (µm)

frequency (Hz)

wavenumber (cm-1)12900 4000 200 10

3.9· 1014 1.2· 1014 6.0· 1012 3.0· 1011

0.77 2.50 50 1000

InfraredNear Mid Far

wavelength (µm)

frequency (Hz)

wavenumber (cm-1)

Figure 1.2 - The infrared region of the electromagnetic spectrum

1.1.1 Interactions of Electromagnetic Radiation with Matter

All forms of spectroscopy deal with the interaction of radiation and matter.

Numerous possible types of interactions exist and many involve transitions between

specific molecular energy states. The monitoring of the absorption and emission of

radiation from different regions of electromagnetic spectrum provides information

regarding these molecular transitions and consequently gives information regarding the

atomic and molecular composition of samples[5].

Quantum mechanical treatments describe both the wave and particle nature of

electromagnetic radiation[5, 6]. As seen in figure 1.1, the electromagnetic spectrum

spans an extremely wide range of frequencies, and therefore, energies. There are a

variety of energy levels that molecules can occupy leading to the possibility of many

transitions between states. These energy transitions are of varying magnitudes with

corresponding frequencies depending upon the specific regions of the spectrum in which

they occur. Radiation from different regions of the electromagnetic spectrum are used as

4

Page 17: Fourier-Transform Infrared Spectroscopic Imaging of

the basis of the many spectroscopic techniques that exist, for which each technique

provides molecular information regarding the sample[2].

Table 1.1 contains examples of different types of spectroscopy based on specific

regions of the electromagnetic spectrum and the type of chemical information probed.

rotational tranisitionsmicrowave spectroscopy1mm to 300 mmMicrowave

nuclear spin transitions (in magnetic field)

Molecular StructureNMR Spectroscopy> 300 mmRadio Waves

50 µm to 1mmFar Infrared

IR Absorption spectroscopyIR Relection spectroscopy

IR emission spectroscopy

2.5 µm to 50 µm Mid Infraredvibrational transitions

thermal emission

770 nm to 2.5 µmNear Infrared

350 nm to 770 nmVisible (VIS)

electronic transitionsfluorescence emissionvibrational transitions

UV-VIS spectroscopyfluorescence spectroscopyRaman spectroscopy

10 nm to 350 nmUltraviolet (UV)

electronic structuremolecular structure

x-ray spectroscopyx-ray crystallography

0.05 Å to 10 nmX-rays

nuclear decay emissionγ - ray spectroscopy< 0.05 Åγ - rays

informationspectroscopywavelength range (λ)

spectral region

rotational tranisitionsmicrowave spectroscopy1mm to 300 mmMicrowave

nuclear spin transitions (in magnetic field)

Molecular StructureNMR Spectroscopy> 300 mmRadio Waves

50 µm to 1mmFar Infrared

IR Absorption spectroscopyIR Relection spectroscopy

IR emission spectroscopy

2.5 µm to 50 µm Mid Infraredvibrational transitions

thermal emission

770 nm to 2.5 µmNear Infrared

350 nm to 770 nmVisible (VIS)

electronic transitionsfluorescence emissionvibrational transitions

UV-VIS spectroscopyfluorescence spectroscopyRaman spectroscopy

10 nm to 350 nmUltraviolet (UV)

electronic structuremolecular structure

x-ray spectroscopyx-ray crystallography

0.05 Å to 10 nmX-rays

nuclear decay emissionγ - ray spectroscopy< 0.05 Åγ - rays

informationspectroscopywavelength range (λ)

spectral region

Table 1.1 - Spectroscopic techniques utilizing different regions of the electromagnetic spectrum

1.2 Basis of Infrared Absorption

Photons in the infrared spectral region have energies representative of transitions

between molecular vibrational energy levels. While spectroscopic techniques exist

which make use of the reflection and emission of infrared radiation, we are most

concerned with the absorption of infrared radiation. Nearly all molecules exhibit an

infrared spectrum, the noted exceptions being homonuclear diatomics, such as the

common gases N2, O2, and H2[5].

5

Page 18: Fourier-Transform Infrared Spectroscopic Imaging of

Various interactions can occur between radiation and matter that result in the

transfer of energy. Quantum mechanical principles require that molecules exist in

quantized energy states and thus the absorption of energy results in bands that

characterize an infrared spectrum.

1.2.1 Requirements for IR Absorption

The wave nature of quantum mechanics is most simply represented by the time

independent Schrödinger equation

ψψ EH = (1.5)

where ψ is the wavefunction of the system, H is the Hamiltonian operator, and E is the

energy of a state characterized by ψ[6]. The wavefunction can be used to calculate the

transition moment R as shown in the equation

τ∂= ∫ ψµψ*jiR (1.6)

for a transition between states i and j, where µ is the electric dipole moment operator (µ =

er, e is the electronic charge, r is the distance between the charges), and dτ indicates the

integration over all space. For vibrational motions, the electric dipole moment µ is

expressed as

...µ)(µ)(µµ0

2

22

21

00 +⎟⎟

⎞⎜⎜⎝

⎛∂∂

−+⎟⎠⎞

⎜⎝⎛∂∂

−+=r

rrr

rr ee (1.7)

where µ0 is the permanent dipole moment, r is the internuclear distance and re is the

equilibrium bond distance[5]. If we consider only the first two terms in equation 1.7 and

substitute for µ in equation 1.6 we obtain

6

Page 19: Fourier-Transform Infrared Spectroscopic Imaging of

τ∂⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛∂∂

−+= ∫ ψµ)(µψ0

0*

jei rrrR (1.8)

which reduces to

τ∂⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛∂∂

−= ∫ ψµ)(ψ0

*jei r

rrR (1.9)

since µ0 is a constant and because of the orthogonality of the

wavefunctions[2].

0=τ∂∫ ψψ*ji

From equation 1.8 it is clear that there must be a change in dipole moment during

the vibration in order for a molecule to absorb infrared radiation. The selection rules

predict that the fundamental absorption will occur with vibrational quantum number

∆υ = ±1 for a harmonic oscillator, with much weaker overtone absorption corresponding

to ∆υ = ±2 etc. for anharmonic conditions[6].

All molecules that are more complex than diatomics have multiple vibrational

modes. These vibrational modes each have associated energies that correspond to the

particular frequency or wavenumber of infrared radiation. The number, type, and

energies of these vibrations are dictated by the molecular structure of the system in terms

of the bonds, geometry, atomic masses, and force fields and are thus representative of

specific molecules[2].

Vibrational modes that produce a change in dipole moment result in the absorption

of IR radiation and are termed infrared-active. Vibrational modes that do not induce in a

change in dipole moment are termed infrared-inactive. The requirement for a change in

dipole moment during a molecular vibration explains why, for example, homonuclear

diatomic molecules do not absorb infrared radiation[4]. 7

Page 20: Fourier-Transform Infrared Spectroscopic Imaging of

1.2.2 Number of Vibrational Modes

While diatomic molecules can vibrate only in one dimension or mode, more

complicated molecular structures present other possible vibrational modes. Linear

molecules with N atoms exhibit 3N-5 vibrational modes, while nonlinear molecules have

3N-6 vibrational modes[5]. Water (a nonlinear triatomic) and carbon dioxide (a linear

triatomic) are illustrative examples. As seen in figure 1.4, the carbon dioxide molecule’s

additional symmetry provides it with four possible vibrational modes while the water

molecule has only three. Note also that the symmetric stretch of the carbon dioxide

molecule produces no net change in dipole moment and is thus infrared-inactive[4].

HH

O

HH

O

HH

O

HH

O

H H

O

H H

O

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

O OC

O OC

O OC

O OC

O OCOO OOCC

O OCOO OOCC

O OCO OCOO OOCC

O OCOO OOCC

A Bwater carbon dioxide

HH

O

HH

O

HH

O

HH

O

H H

O

H H

O

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

O OC

O OC

O OC

O OC

O OCOO OOCC

O OCOO OOCC

O OCO OCOO OOCC

O OCOO OOCC

A B

HH

O

HH

O

HH

O

HH

O

H H

O

H H

O

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

O OC

O OC

O OC

O OC

O OCOO OOCC

O OCOO OOCC

O OCO OCOO OOCC

O OCOO OOCC

HH

O

HH

O

HH

O

HH

O

H H

O

H H

O

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

Bend

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν2

ν1

νn

1596 cm-1

3756 cm-1

3652 cm-1

band position

IR-active

IR-active

IR-active

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

IR-active666 cm-1

(degenerate)ν2

Bending (in plane)

Bending (out-of-plane)

Asymmetric Stretch

Symmetric Stretch

Vibration

ν3

ν1

νn

2350 cm-1

1340 cm-1

band position

IR-active

IR-inactive

infrared activity

O OC

O OC

O OC

O OC

O OCOO OOCC

O OCOO OOCC

O OCO OCOO OOCC

O OCOO OOCC

A Bwater carbon dioxide

Adapted from [2] Figure 1.3 - Vibrational modes and IR activity of water vapor (A) and carbon dioxide (B) molecules

As molecular structural complexity increases, other types of vibrational modes

become possible. The methylene group, for example is capable of six different

vibrational modes as illustrated in figure 1.4.

8

Page 21: Fourier-Transform Infrared Spectroscopic Imaging of

rocking

symmetric stretch

wagging

asymmetric stretch

Methylene Normal Modes

twisting

scissoring

rocking

symmetric stretch

wagging

asymmetric stretch

Methylene Normal Modes

twisting

scissoring

H H

C

H H

C

H H

C

H H

C

H H

C C

H H

C

H H

CC

H H

C

H H

CC

H H

C

H H

CC

H H

C

H H

CC

H H

C

H H

CC CCC

H H H H H H

H H H H H H

rocking

symmetric stretch

wagging

asymmetric stretch

Methylene Normal Modes

twisting

scissoring

rocking

symmetric stretch

wagging

asymmetric stretch

Methylene Normal Modes

twisting

scissoring

H H

C

H H

C

H H

C

H H

C

H H

C C

H H

C

H H

CC

H H

C

H H

CC

H H

C

H H

CC

H H

C

H H

CC

H H

C

H H

CC CCC

H H H H H H

H H H H H H

Adapted from [2] Figure 1.4 - Vibrational modes of methylene group

1.2.3 Group Frequencies

Various chemical functional groups exhibit specific infrared frequencies

representative of their structures. Frequencies such as these are known as characteristic

or group frequencies[4]. Many of the most common functional groups with characteristic

group frequencies are familiar organic groups. Functional group frequencies allow the

spectroscopist to use IR spectra to qualitatively identify structural elements in samples.

Since vibrational frequency absorption profiles parallel functional group structure, the

spectroscopist investigating biological material using vibrational techniques often

depends upon existing databases and extensive compilations of spectral information.

9

Page 22: Fourier-Transform Infrared Spectroscopic Imaging of

10

1.3 IR Spectral Feature of Tissues

Modern approaches to histology categorize cells into different types based on their

primary physiological function[7]. In such a system cells belong to one or more of the

following groups: epithelial cells, support cells, contractile cells, nerve cells, germ cells,

blood cells, immune cells, or hormone-secreting cells.

From a molecular point of view, all of these various types of specialized cells

encountered in biological tissue are predominately comprised of four major types of

biomolecules or their subunits: proteins, carbohydrates, lipids, and nucleic acids.

Additionally, all four of these types of molecules each have a great deal of structural

redundancy. That is, they tend to form polymeric molecules based on subunits that while

different, reflect structural similarity. For example, thousands of different proteins exist

in a typical cell, and while the individual structure of each protein is different, they are all

made from the same set of amino acids, and share a common backbone structure.

1.3.1 Proteins

Protein molecules play many fundamental roles in the life of every cell in addition

to serving various important extracellular functions in many tissues. The significance of

proteins to biological organisms cannot be understated and their utility is evident in the

many functions they perform including: enzymatic catalysis, transport and storage,

coordinated motion, mechanical support, immune protection, generation and transmission

of nerve impulses, and control of growth and differentiation[8].

All proteins are formed as linear chains of amino acid building blocks that can form

various secondary and tertiary structures. Eukaryotic proteins are typically assembled

Page 23: Fourier-Transform Infrared Spectroscopic Imaging of

from a set of 20 different α-amino acids that share a common template and are

distinguished by unique side chain structures[9]. Figure 1.5 shows the molecular

structure of a typical amino acid.

Amino group

Carboxylateion

Side chain is distinctive for each amino acid

R

H

+H3N C

O

O-Cα

Amino group

Carboxylateion

Side chain is distinctive for each amino acid

R

H

+H3N C

O

O-Cα

R

H

+H3N C

O

O-Cα

Figure 1.5 - Structure of a typical amino acid

All amino acids share a common structure that includes a central or α-carbon atom

bonded to a carboxyl group, an amino group and a hydrogen atom. At physiologic pH

the amino group is protonated (NH3+) and the carboxyl group exists as the carboxylate

ion (COO-)[9], displayed in figure 1.5. Each different amino acid contains a distinctive

structure at the side chain position designated as R in figure 1.5.

The primary protein or polypeptide structure is formed by linking these amino acid

subunits together in a linear chain via a condensation reaction between the amino and

carboxyl groups of adjacent amino acids in a linear chain[10]. The linkage that is formed

between these amino acid subunits is known as a peptide bond and polypeptide chains

that result form a repeating backbone structure that is the same for all proteins. Figure

1.6 shows the basic protein primary structure and the locations of these peptide bonds.

11

Page 24: Fourier-Transform Infrared Spectroscopic Imaging of

12

N

RA

H

N C

O

C

H

RB

C

O

NC

RC

H

C

O

N…C

Peptide bonds

…C

Amino Acid A Amino Acid B Amino Acid C

H

H

H

H

H

H

H

HRA

H

N C

O

NC

H

RB

C

O

NC

RC

H

C

O

N…C

Peptide bonds

…C

Amino Acid AAmino Acid A Amino Acid BAmino Acid B Amino Acid CAmino Acid C

HH

H

H

H

H

H

H

H

Figure 1.6 - Basic polypeptide structure

The polypeptide backbone structure consists of several functional groups, including

a C-N group, a C-H group, an NH2 group, and a carbonyl group (C=O). Since these

functional groups repeat for every amino acid in a protein regardless of the protein’s

identity or higher-order structure, the absorbance bands resulting from these structures

dominate the IR spectra of most proteins. The most prominent of these absorbances

include; the Amide I absorption near 1650 cm-1 arising from C=O stretching vibrations

(80%) weakly coupled to C-N stretching vibrations (20%), the Amide II absorption near

1545 cm-1 arising from N-H bending vibrations (60%) coupled to C-N stretching

vibrations (40%), the Amide III absorption near 1236 cm-1 arising from C-N stretching

vibrations, and the Amide A absorbance near 3290cm-1 arising from N-H stretching

vibrations[11].

In their native states, most proteins do not exist as simple linear polypeptide

structures, but instead form complex secondary and tertiary structures that impart a

distinct three-dimensionality to a particular protein. The most common protein

Page 25: Fourier-Transform Infrared Spectroscopic Imaging of

secondary structures are the α-helix and β-pleated sheet configurations depicted in figure

1.7.

R

R

C

C

C

CO

O

O

O

N

N

N

N

Cα H

H

H

H

H

H

H

H

H

R

R

R

R

R

C

C

C

C O

O

O

O

N

N

N

N

CαH

H

H

H

H

H

H

H

H

R

R

R

R

R

C

C

C

CO

O

O

O

N

N

N

N

Cα H

H

H

H

H

H

H

H

H

R

R

R

R

R

C

C

C

C O

O

O

O

N

N

N

N

CαH

H

H

H

H

H

H

H

H

R

R

R

R

N

N

N

N

C

C

C

C

C

C

C

N

N

H

H

H

HCα

H

H

H

H

H

H

H

H

HCα

R

RH

N

R

R

RNCα

CH

R

O

O

O

H

O

OC

O

N

OR

H

O

R

O

R

N

N

N

N

C

C

C

C

C

C

C

N

N

H

H

H

HCα

H

H

H

H

H

H

H

H

HCα

R

RH

N

R

R

RNCα

CH

R

O

O

O

H

O

OC

O

N

OR

H

O

R

O

α-helix β-sheet (antiparallel)

R

R

C

C

C

CO

O

O

O

N

N

N

N

Cα H

H

H

H

H

H

H

H

H

R

R

R

R

R

C

C

C

C O

O

O

O

N

N

N

N

CαH

H

H

H

H

H

H

H

H

R

R

R

R

R

C

C

C

CO

O

O

O

N

N

N

N

Cα H

H

H

H

H

H

H

H

H

R

R

R

R

R

C

C

C

C O

O

O

O

N

N

N

N

CαH

H

H

H

H

H

H

H

H

R

R

R

R

N

N

N

N

C

C

C

C

C

C

C

N

N

H

H

H

HCα

H

H

H

H

H

H

H

H

HCα

R

RH

N

R

R

RNCα

CH

R

O

O

O

H

O

OC

O

N

OR

H

O

R

O

R

N

N

N

N

C

C

C

C

C

C

C

N

N

H

H

H

HCα

H

H

H

H

H

H

H

H

HCα

R

RH

N

R

R

RNCα

CH

R

O

O

O

H

O

OC

O

N

OR

H

O

R

O

R

N

N

N

N

C

C

C

C

C

C

C

N

N

H

H

H

HCα

H

H

H

H

H

H

H

H

HCα

R

RH

N

R

R

RNCα

CH

R

O

O

O

H

O

OC

O

N

OR

H

O

R

O

R

N

N

N

N

C

C

C

C

C

C

C

N

N

H

H

H

HCα

H

H

H

H

H

H

H

H

HCα

R

RH

N

R

R

RNCα

CH

R

O

O

O

H

O

OC

O

N

OR

H

O

R

O

α-helix β-sheet (antiparallel)

Figure 1.7 - Common Protein Secondary Structures: α-helix and β–sheet

β–pleated sheet structures can form between parallel polypeptide chains, or between strands with antiparallel orientation, as shown in the figure. The dotted lines indicate hydrogen bonds.

Both of these recurrent secondary structures involve hydrogen bonding between the

oxygen atoms of backbone carbonyl groups and the hydrogen atoms of backbone N-H

groups indicated in the figure as dotted lines. These structural arrangements change bond 13

Page 26: Fourier-Transform Infrared Spectroscopic Imaging of

14

angles and other structural parameters, causing frequency shifts of absorbance bands

arising from backbone vibrations. As a result, the relationship between IR band positions

of protein backbone absorbances, most notably the Amide I absorbance near 1650 cm-1,

and protein structure has been the subject of much work over the past decade[12-16].

For example, several studies have examined the amide I bands of polypeptides and

proteins whose structures are known to be dominated by one of the common secondary

structure motifs, such as α-helix, β–sheet, or unordered structures[17-19]. Such studies

have led to the development of some empirical rules for the correlation of amide I band

features and common secondary structural motifs.

On the basis of these empirical rules, IR bands in the 1660-1650 cm-1 spectral

region are assigned to α-helices, 1640-1620 cm-1 to β–sheets, 1695-1660 cm-1 to β-sheets

and β-turns, and 1650-1640 cm-1 to unordered structures[20]. Such empirical rules are

useful guidelines for obtaining structural information from vibrational spectroscopic

information, however, many studies show that such rules are not free from

shortcomings[19]. For instance, IR studies of proteins such as myoglobin and

hemoglobin, for which x-ray crystallographic data suggests highly helical-structures with

no β-sheets, have shown Amide I absorbances in the 1640-1620 cm-1 region[21, 22].

While no conclusive evidence exists to explain the presence of such lower-frequency α-

helix amide I bands, some have suggested that strong hydrogen bonding of peptide

groups with solvent molecules and distortion of helix structures may contribute to such

findings[23, 24].

Page 27: Fourier-Transform Infrared Spectroscopic Imaging of

15

1.3.2 Carbohydrates

Carbohydrates are aldehyde or ketone compounds with multiple hydroxyl groups.

These important biomolecules play three central roles in all organisms: First, they serve

as energy stores and metabolic intermediates. Stored glycogen can be readily broken

down into glucose, a preferred metabolic fuel. Glucose is broken down to yield

adenosine triphosphate (ATP), a phosphorylated sugar derivative and universal currency

of energy in the organism. The second important role of carbohydrates is as basic

structural components of nucleic acids. Ribose and deoxyribose sugars are structural

units of all nucleotides and ribonucleotides whose sequence in nucleic acids is

responsible for the storage and expression of genetic information. A third important role

of carbohydrates in organisms is that they are often linked to proteins and lipids on cell

membranes, many playing critical roles in cell signaling and recognition[25, 26].

Common cellular carbohydrates have many vibrational spectral features in the

fingerprint region of the mid-IR spectrum due to various vibrational modes of C-O, C-C,

and carboxylate groups. Infrared spectroscopy has been used extensively to help

characterize biologically important polysaccharide cell-surface components, including

glycolipids like diacyl sugars[27], cerebrosides[28, 29], gangliosides[30, 31],

lipopolysaccharides[32-34], and mucopolysaccharides[35].

1.3.3 Lipids

Lipids form another important class of biomolecules found in tissue that play many

important roles. Like carbohydrates, lipids provide an important source of energy for

metabolism. The hydrophobic nature of lipids contributes significantly to their central

Page 28: Fourier-Transform Infrared Spectroscopic Imaging of

16

role in cellular membrane function, providing barriers which partition cells and

subcellular organelles. Additionally, lipids perform a variety of other important

functions, from the coenzyme roles of fat-soluble vitamins to the regulatory roles of

prostaglandins and steroid hormones to structural and functional roles in the nervous

system.

Lipids all share the characteristic of having non-polar, hydrophobic domains. In

many cases, long chain fatty acids are responsible for this hydrophobicity, and such lipids

have many vibrational modes associated with C-H groups across the fingerprint region of

the mid-IR. The spectral frequency region between 3000-2800 cm-1 also contains four

prominent absorbance bands common to many lipids: the methyl antisymmetric stretch

(asυCH3) at 2962 cm-1, the methyl symmetric stretch (sυCH3) at 2872 cm-1, the

antisymmetric CH2 stretch (asυCH2) between 2936-2916 cm-1, and the symmetric CH2

stretch (sυCH2) between 2863-2843 cm-1[36].

Unfortunately, most standard methods for the preparation of sectioned tissue

involve the use of one or more nonpolar solvents such as ethanol or xylenes that remove

lipids from the tissue section[37, 38]. As a tissue source for FT-IR spectroscopic studies,

formalin-fixed paraffin-embedded tissue offers some advantages over frozen tissue

including higher-quality preservation and access to large libraries of preserved tissue,

however, paraffin exhibits many of these common lipid absorbances, and therefore must

be removed from tissue sections intended for spectroscopic analysis. Effective paraffin

removal requires the use of strong nonpolar solvents such as hexane for several hours at

temperatures of 40°C further contributing to the extraction of physiologic lipids from

paraffin-embedded tissue.

Page 29: Fourier-Transform Infrared Spectroscopic Imaging of

17

1.3.4 Nucleic Acids

Nucleic Acids have been studied extensively in both purified state as well via model

compounds[39]. The most prominent absorbances reported are due to vibrations of

several functional groups on the repeating backbone structure of nucleic acids. These

include absorbances near 1080cm-1 and 1240 cm-1 attributed respectively to the

symmetric and asymmetric stretch of phosphodiester (PO2-) moieties[40]. However, the

ability of IR spectroscopy to attain vibrational information from quiescent nuclear DNA

from cell preparations or tissue sections has recently been called into question and some

theoretical analyses of chromatin density and packing used to support the idea that

nuclear DNA is too dense to produce appreciable absorbances in transmission IR

spectroscopic experiments[41].

1.4 FTIR Spectroscopy Background

Modern instrumental approaches to the collection of spatially-resolved infrared

spectroscopic data share many characteristics and all benefit from the extensive advances

made in the field of Fourier transform infrared (FTIR) spectroscopy over the past three

decades. Several excellent books[4, 42, 43] have been written on the subject of FT-IR

spectroscopy and contain comprehensive information on the technology that has been

implemented for years in commercial FT-IR spectroscopy systems.

Infrared microspectroscopic imaging systems share many common features. Most

consist of a research-grade FT-IR spectrometer that provides an output beam of

modulated infrared radiation used as a source for an infrared microscope equipped with

infrared detectors[44]. Modern approaches to the collection of spatially-resolved spectral

Page 30: Fourier-Transform Infrared Spectroscopic Imaging of

18

data are best differentiated in terms of the type of infrared detection employed. The

following sections discuss instrumental aspects of spectrometers and infrared

microscopes, as well as strategies for collecting FT-IR spectroscopic imaging data with

three different types of infrared detection: single-point mapping, raster scanning with

linear multichannel detectors, and global FT-IR imaging with Focal Plane Array (FPA)

detectors.

1.4.1 FTIR Spectrometers

The majority of commercial research-grade FTIR spectrometers incorporate a

broadband infrared source, Michelson interferometer, sample compartment, and infrared

detection with either deuterated triglycine sulfate (DTGS) or mecury cadmium telluride

(MCT) single-point detectors. Many commercial FTIR instruments exist for dedicated

analyses typically implemented in industrial settings for process assessment and quality

control analyses. Such spectrometers are typically designed to be lower in cost than

research-grade spectrometers, which offer more flexibility in the types of measurements

that are possible as well as increased sensitivity and higher spectral signal-to-noise ratios

(SNRs).

Figure 1.8 shows the schematic design of the Michelson interferometer, which is

the optical portion of the spectrometer that is used to modulate the radiation. The

interferometer is composed of two perpendicular beam paths often referred to as separate

arms of the interferometer. These beampaths intersect at the beamsplitter, an optical

component that when placed at 45-degree angle to the normal both reflects and transmits

exactly 50% of incident radiation. In the mid-IR region, beamsplitters are typically

Page 31: Fourier-Transform Infrared Spectroscopic Imaging of

constructed from potassium bromide (KBr) with a thin coating of germanium (Ge) or

silicon (Si), and many commercial instruments allow beamsplitters to be changed to other

materials for coverage of specific spectral regions[42].

Figure 1.8 - Michelson Interferometer

As depicted in Figure 1.8, polychromatic radiation from an infrared source,

typically a ceramic globar, is passed through an aperture to form a beam. This beam

strikes the beamsplitter at a 45° angle, dividing the beam in half. Half of the beam is

directed at a fixed mirror, while the other half is diverted to a mirror whose displacement

can be varied along the axis of the incident beam. After striking these mirrors, the beams

in the two arms of the interferometer are sent back to the beamsplitter, where they

recombine and interfere with each other. The beamsplitter divides the recombined beam

in half again, sending half back toward the source, while the other half is used for

spectroscopy and is directed through sample and subsequently detected[4].

19

Page 32: Fourier-Transform Infrared Spectroscopic Imaging of

20

When the moving mirror occupies a displacement where the pathlengths in the two

arms of the interferometer are equal, then the recombining beams are precisely in-phase

and only interfere constructively. This mirror position produces the most intense beam

for every frequency of radiation. As the mirror moves from this position, a pathlength

difference is created in the two arms of the interferometer that causes specific

interference patterns for different mirror displacements. If the mirror is continuously

scanned, then the intensity of the recombined beam will vary with respect to time in a

frequency or wavelength dependent manner[2].

The function of the spectrometer is to encode a modulation on the polychromatic IR

source radiation such that detection of the intensity of the encoded radiation with respect

to time or in the “time domain” yields spectral information in the “frequency domain”.

The “Fourier transform” part of the technique’s name refers to the mathematical

operation that is required to transform the raw data collected by the instrument in the time

domain, known as the interferogram, into a intensity profile in the frequency domain,

otherwise know as an infrared spectrum.

1.4.2 Infrared Microscopy

Infrared microspectroscopic imaging systems typically couple the modulated output

beam of a FTIR spectrometer to an infrared microscope for use as source radiation for

obtaining spectroscopic information from microscopic regions of a sample. Infrared

microscopes perform similarly to conventional optical microscopes and are typically set

up to image with visible light along the same optical path. However, they have many

structural differences that stem from some fundamental properties of infrared radiation.

Page 33: Fourier-Transform Infrared Spectroscopic Imaging of

21

One major limitation of infrared spectroscopy is related to its exceptional molecular

sensitivity. As mentioned in section 1.2, all covalently bonded molecules, with the

exception of homonuclear diatomics, absorb infrared radiation. Optical components used

in conventional microscopes are composed almost exclusively of borosilicate glass or

quartz, both of which have broad absorbances over much of the infrared spectrum. For

this reason, infrared microscopes are designed to use reflective optics wherever possible,

and refractive optics have to be manufactured from alternative materials, such as halide

salts, which are transparent over the spectral regions of interest[42].

Most Infrared microscopes use Cassegrain condenser and objective lenses and can

be operated in either transmission or reflectance modes. In reflectance mode, one side of

the Cassegrain objective primary mirror is typically used to direct the radiation onto the

sample while the opposite portion of the primary mirror is used to collect the reflected

radiation. Infrared microscopes are often outfitted with automated high-precision

motorized mapping stages, which permit the sample to be positioned precisely in the

plane perpendicular to the optical path. Most microscopes incorporate a visible light

source and detection system, typically a video camera. Adjustable mirrors are used to

switch between visible and infrared modes and some models incorporate a beamsplitter to

allow for simultaneous imaging in both spectral regions[45].

The different strategies that can be employed to collect spatially-resolved infrared

microspectroscopic data depend on the types of infrared detection systems available of

the microscope[44]. Panels A-C of Figure 1.9 depict three different approaches based

respectively on single-point, linear-array, and focal plane array (FPA) detection. A

discussion of each approach follows.

Page 34: Fourier-Transform Infrared Spectroscopic Imaging of

Sample

Aperture

Aperture

Singl e ElementInfrared Detector

Microscope

CCD Visible Detector

Turning Mirror

Visible LightSource

Prec ision Stage

Rapid-ScanInterferometer

Sample

MultichannelInfrared Detector

Microscope

CCD Visible Detector

Turning Mirror

Visible LightSource

Precision Stage

Rapid-ScanInterferometer

Sample

MultichannelInfrare d Detector

Microscope

CCD Visible Detector

Turning Mirror

Visible LightSource

Microscop e Stage

Rapid- or Ste p-ScanIn terferometer

Focal Plane ArrayDe tector

A B

C

Sample

Aperture

Aperture

Singl e ElementInfrared Detector

Microscope

CCD Visible Detector

Turning Mirror

Visible LightSource

Prec ision Stage

Rapid-ScanInterferometer

Sample

MultichannelInfrared Detector

Microscope

CCD Visible Detector

Turning Mirror

Visible LightSource

Precision Stage

Rapid-ScanInterferometer

Sample

MultichannelInfrare d Detector

Microscope

CCD Visible Detector

Turning Mirror

Visible LightSource

Microscop e Stage

Rapid- or Ste p-ScanIn terferometer

Focal Plane ArrayDe tector

A B

C

Figure 1.9 - Three Instrumental Approaches for collection of spatially resolved FTIR spectroscopic data A) Point-mapping using single element detection; B) Raster-Scan imaging using linear multichannel detection; and C) Global FT-IR imaging using 2-D focal plane

1.4.3 Mapping with Single-Point Detectors

In single element microspectroscopic instrumentation, spectral information from a

small, specified area of the sample is obtained by restricting the area illuminated by the

infrared beam using opaque apertures of controlled size. The collected radiation is then

diverted to a sensitive detector. To identify the area to be examined, however, a

corresponding white light optical image is also required. Clearly, focusing the infrared

22

Page 35: Fourier-Transform Infrared Spectroscopic Imaging of

23

beam for maximal throughput and minimal dispersion in the sample plane requires the

optical and infrared paths be parfocal and collinear[45].

By restricting the infrared beam to a small spatial area of the sample, and

sequentially moving to different regularly-spaced sample locations with a high precision

microscope stage, spatially-resolved spectroscopic data from large sample areas can be

mapped out point by point. This strategy, often referred to as point-mapping, suffers from

several limitations.

The cross-sectional diameter of the beams used in such infrared microscopes must

be large enough to fully illuminate the area passed by the largest aperture setting that may

be employed, for example a 100x100 um square. There is a tradeoff between the spatial

resolution of mapping data that can be acquired and corresponding throughput due to the

need to block out more and more of the available radiation. Aperture use decreases the

instrumental throughput due to diffraction when the aperture is of the same dimension as

the wavelength of light (~3-14um), thus limiting the highest achievable data spatial

resolution. Apertures also permit the passage of some diffracted light from outside the

apertured region. The use of a second set of apertures in tandem to reject stray radiation

can improve spatial fidelity, unfortunately at the cost of additional throughput loss.

Throughput is important because it directly affects the spectral signal to noise ratio

(SNR), and losses in throughput require larger acquisition times for signal recovery[42].

Data acquisition time is the major drawback to single-point mapping approaches.

Spectral information is acquired for each spatial location in the final map one-by-one and

there is significant time overhead for moving the sample to each new sampling location.

Page 36: Fourier-Transform Infrared Spectroscopic Imaging of

24

1.4.4 Raster-scan Imaging Using Multichannel Detectors

While single element microspectroscopy provides the capability to obtain spectra

from small spatial regions, poor SNR characteristics, diffraction effects and stray light

issues resulting from the use of apertures limit the applicability of this point mapping

approach. A multichannel detection approach to circumvent some of these issues has

recently been implemented[46] with a linear array detector employed to image an area

corresponding to a rectangular spatial area on the sample. The sample stage is moved

precisely to sequentially image a selected spatial area on the sample. This data collection

strategy is referred to as push-broom mapping or raster scanning. The process is

conceptually similar to point-by-point mapping but takes advantage of the multiple

channels of detection. Hence, imaging a large sample area is faster by a factor of n, for a

linear array detector containing n elements. The instrument is schematically displayed in

Figure 1.9B.

Point mapping detectors are typically 100 – 250 µm in size; in contrast, an

individual detection element in a linear array detector is of the order of tens of

micrometers. Employing a linear array eliminates the need for apertures, as small

detector elements directly image different sample spatial regions. For example, a detector

element 25 µm in size can be operated at 1:1 magnification or 4:1 magnification to

provide a 25 µm or a 6.25 µm effective pixel size with available, relatively aberration-

free infrared optics. This approach circumvents the debilitating diffraction effects

resulting from the use of small apertures in single channel detection systems and provides

higher quality data when desired spatial resolutions approach the wavelengths of light

being used. In addition, the spatial resolution, data quality, and time for data acquisition

Page 37: Fourier-Transform Infrared Spectroscopic Imaging of

25

are no longer coupled as in point mapping methods. The data acquisition time depends

solely on the size of the image and quality of data desired, and is correlated less with the

spatial resolution, which is determined by the employed optics.

A high-precision, motorized stage that reproducibly steps in small increments is

used and the interferometer is operated in a continuous scan mode. In combination with

high performance multichannel detectors, this mode combines high performance

multichannel detectors with the most desirable properties of rapid-scan interferometry to

yield high quality spectroscopic imaging data.

1.4.5 Global FTIR Spectroscopic Imaging

The state of the art in FTIR microspectroscopic imaging instrumentation is the

combination of an infrared microscope equipped with a focal plane array (FPA) detector

and an FTIR spectrometer[47, 48], as shown in Figure 9C. FPA detectors are

constructed of thousands of individual detection elements laid out in a two-dimensional

grid pattern. An FPA matched to the characteristics of the optical system is capable of

imaging the entire field of view afforded by the optics and of utilizing a large fraction of

the infrared radiation spot size at the plane of the sample. The increase in the number of

individual detectors with respect to a linear array provides a correspondingly larger

multichannel advantage. For example, an FPA with pixel dimensions p x p, provides a p2

time savings relative to a single element detector and a p2/n time savings compared to a

linear array detector containing n elements. For a 128 x 128 element FPA detector

relative to the single element case, the advantage is a factor of 16,384, while compared to

a 16-element linear array detector; the multichannel advantage is a factor of 2048. FPA

Page 38: Fourier-Transform Infrared Spectroscopic Imaging of

26

detectors are also capable of imaging large spatial areas simultaneously without inherent

inefficiencies of moving the sample or re-setting the interferometer to scan a different

area. The considerable reduction in data acquisition times allows for imaging large areas,

as well as the examination of dynamic processes in a single field of view[49].

The first and, to date, most popular approach to FTIR micro-imaging spectrometers

incorporates a step-scan interferometer[50]. While continuous or rapid-scan

spectrometry involves scanning the moving mirror at a constant velocity, a step-scan

interferometer is capable of stepping the moving mirror to discrete, evenly-spaced

intervals and maintaining individual mirror positions with very little displacement error.

A constant retardation over an extended time period allows suitable time for signal

averaging and for data readout and storage. Short time delays prior to data acquisition

are necessary for mirror stabilization at the onset of the step. Detector signal is integrated

for only a fraction of the total time required for collection of each frame. The integration

time, number of frames co-added, and number of interferometer retardation steps (a

function of desired spectral resolution) determine the total time required for the

experiment. Since the integration time determines the data quality, efforts have been

made to increase the ratio of the integration time to the total data acquisition time[51].

Imaging configurations that utilize a rapid scan interferometer have been proposed

for small arrays[52]. Slow data readout and storage rates for many FPA detectors

preclude conventional rapid-scan mirror velocities, thus approaches must make use of so

called slow-scan mirror velocities of ≤ 0.01 cm/s. A generalized data acquisition scheme

that permits true rapid scan data acquisition for FPA detectors has been proposed[53],

where the integration time of individual frames collected by the FPA detector is

Page 39: Fourier-Transform Infrared Spectroscopic Imaging of

27

negligible with respect to the complete interferogram acquisition. For most FPA

detectors available today, the motion of the moving mirror does not allow co-addition of

frames at individual retardations in the continuous scanning mode, but successive single-

frame acquisitions can be averaged to increase data SNRs. Compared to step-scan data

acquisition, rapid scan data collection (mirror velocity > 0.025 cm/s) allows for fast

interferogram capture as no time is spent on mirror stabilization. The error arising from

the deviation in mirror position during frame collection is hypothesized to be the next

largest contributor of noise compared to the dominant contribution from random detector

noise[50]. At present, the advantages of continuous-scan relative to step-scan approaches

are a decreased cost of instrumentation and an increased data collection efficiency.

1.5 Spectroscopic Imaging: Data Structure and Applications

Spectroscopic imaging data, regardless of its method of collection, can be

conceptualized as an image cube with two dimensions corresponding to the spatial axes

of the sample and the third dimension to the spectral frequency or wavelength. Digital

image data is represented as a collection of rectangular picture elements or pixels, each

with an associated brightness value or magnitude. Spectroscopic image data can be

thought of as a collection of super-imposable and spectrally consecutive image planes,

whose pixel values consist of the spatially independent absorbance at the spectral

frequency or wavelength specified by the image plane. Alternatively, the data structure

can be conceptualized to consist of individual spatial locations or pixels each with an

associated absorbance spectrum. The concept of the image cube is represented

schematically in figure 1.10.

Page 40: Fourier-Transform Infrared Spectroscopic Imaging of

x

y

Wavelength Axis

Spatial Axes

Figure 1.10 - Schematic representation of the image cube

These alternative views of the data structure influence the type of information that

can be extracted from the data. For example, we can specify distinct spatial locations in a

spectroscopic image, and display the associated spectra for simultaneous comparison of

absorption features across the full spectral region collected. Alternatively we can specify

a particular absorption feature of interest and display the associated spectral image plane.

The brightness values of pixels in such an image will correspond to the sample’s spatial

distribution of the species responsible for the absorption at the associated spectral

frequency.

FTIR imaging of biological systems has demonstrated a potential to complement

other imaging approaches. For biomedical applications, the technique may be used to

28

Page 41: Fourier-Transform Infrared Spectroscopic Imaging of

29

examine chemical changes due to pathological abnormalities and to follow histological

alterations with high accuracy. Non-destructive morphological visualization of chemical

composition rapidly provides structural and spatial information at an unprecedented level.

Specifically, thousands of spectra routinely acquired in an imaging experiment may be

employed for statistically meaningful data analyses, which in the example of biological

tissue samples may prove ultimately useful in medical diagnoses. Since the visualization

contrast is dictated by inherent chemical and molecular properties, no sample treatments,

such as histopathological staining techniques required for optical microscopy, are

necessary.

A typical example of the type of tissue information that can be retrieved was

demonstrated by examining monkey cerebellum sections[54]. Distributions of lipid

relative to protein allowed easy differentiation of white and gray matter areas. Purkinje

cells in rat cerebella, which strongly influence motor coordination and memory

processes, were visualized using FTIR imaging techniques[55, 56]. Neuropathologic

effects of a genetic lipid storage disease, Niemann-Pick type C (NPC)[57], were

distinguishable on the basis of spectral data without the use of external histological

staining. Statistical analysis provided a numerical confirmation of these determinations

consistent with a significant demyelination within the cerebellum of the NPC mouse. IR

spectroscopy has been used for a number of years to characterize mineralized structures

in living organisms (notably, bone). FTIR imaging spectroscopy[58, 59] of bone allows

spatial variations of a number of chemical components to be non-destructively monitored.

Correlations in bone between FTIR imaging and optical microscopy involving chemical

composition, regional morphologies and the developmental processes have been made,

Page 42: Fourier-Transform Infrared Spectroscopic Imaging of

30

and an index of crystallinity/bone maturity could be determined providing structural

information in a non-destructive manner[60].

1.5.1 Image Classification Methods

One of the most useful approaches to extracting data from such data structures is

the process of image classification. Image classification algorithms automatically assign

each pixel in an image scene to a specific class or group based on its spectral properties

or pattern. Unsupervised Classification refers to the automatic partitioning of pixels into

classes of spectral similarity without the use of any class training data. Supervised

Classification is the process of classifying pixels into specific classes based on their

spectral similarity to user-supplied training data for each class.

Unsupervised classification methods have the advantage that no extensive prior

knowledge of the image scene is necessary and the potential for human error is far less

than with supervised methods. Additionally, they are useful for finding natural spectral

patterns and groups in spectral images. However, they are limited in their usefulness by

the need to identify the resulting classes after the classification is performed[61]. For this

reason, such unsupervised methods are of little usefulness for diagnostic implementation.

Supervised classification methods have several advantages relative to unsupervised

strategies. First, the analyst has control over the specific number and identity of class

categories and can tailor them for specific tasks. Supervised classification is tied to areas

of known identity, determined through the process of selecting training regions.

Additionally, regions of training data can be used during the process of classifier

development to evaluate classifier performance. While inaccurate classification of

Page 43: Fourier-Transform Infrared Spectroscopic Imaging of

31

training data indicates serious classification problems and/or problems with training data

selection, accurate classification of training data does not always assure accurate

classification of other image data[62].

Supervised image classification methods have several disadvantages and limitations

as well. By creating classes and assigning training populations, the analyst imposes a

classification structure on the data. If the user-defined class structure does not match the

natural class structure within the data, the classes may not be distinct or well defined in

multidimensional space. Training populations that do not accurately represent the natural

distribution of values within a class may result in severe classification error[63]. Finally,

classes unknown to the analyst and not included in the training data may also be

misclassified and thereby remain undiscovered.

1.6 Prostate Background

1.6.1 Anatomy and Histology

In men, the prostate is a retroperitoneal gland located just below the bladder that

surrounds the urethra. The gland is divided into four zones: peripheral, central,

transitional, and periurethral as shown in Figure 3.1. Distinctions between these zones

are important because proliferative lesions vary according to the zone in which they

occur. For instance, nodular hyperplasia, also known as benign prostatic hypertrophy or

hyperplasia (BPH), occurs predominantly in the central zone, whereas most

adenocarcinomas occur in the peripheral zone[64].

Page 44: Fourier-Transform Infrared Spectroscopic Imaging of

adapted from [64]

Figure 1.11 - Zonal Anatomy of the Prostate

Histologically, the prostate is a compound tubuloalveolar gland in which glandular

spaces are lined by epithelium. Specifically, the gland is lined by a layer of low cuboidal

epithelium at the basal surface, which is covered by a layer of columnar mucus-secreting

cells. The glands contain a discrete basement membrane and are separated by abundant

fibromuscular stroma. Some ducts in the gland are lined by tall columnar epithelium, but

as they approach the urethra, the epithelium changes to more cuboidal and eventually into

the transitional epithelium that lines the urethra and urinary bladder[65].

32

Page 45: Fourier-Transform Infrared Spectroscopic Imaging of

33

While prostatic epithelial tissue and fibromuscular stroma make up the bulk of the

gland, there are several other important histological features seen in the prostate.

Numerous blood vessels run throughout the prostate, as well as peripheral nervous tissue

innervating the gland. Prostates from older men frequently contain small, spherical

corpora amylacea comprosed primarily of condensed glycoprotein in the glandular

lumina[7].

1.6.2 Prostate Pathology

1.6.2.1 Incidence

Prostatic carcinoma is the most common form of cancer in men and it is estimated

that 221,000 new cases will be diagnosed in the United States in 2003[66]. The

incidence of newly diagnosed cases of prostate cancer in the US was 100,000 in 1988,

and has risen steadily since then to just under 200,000 in 1994[67]. Mortality in the US

due to prostate cancer rose from 28,000 to 36,000 during the same time period, however

recent evidence suggests that mortality has peaked and may be falling[68]. The estimated

mortality for US men in 2003 is 29,000[66]. This decline has been attributed to increased

screening efforts and active treatment of localized disease by radiation and radical

prostatectomy[69].

1.6.2.2 ‘Latent’ Prostate Cancer

In 1954, Franks observed an extraordinarily high prevalence of microscopic foci of

what he termed ‘latent’ prostate cancer during autopsy of men who died from other

diseases[70]. His observations have been corroborated by several investigators[71, 72]

Page 46: Fourier-Transform Infrared Spectroscopic Imaging of

34

and the occurrence of these incidental cancers has been shown to increase with age

affecting approximately 20% of men in their 20’s, 30% of men in their 50’s, and 70% of

men in their 80’s[73]. The lifetime chance that a man will develop clinically apparent

prostate cancer is less than 10%[74], thus the majority of these tiny cancers detected at

autopsy are clinically insignificant. While it is clear that early diagnosis and treatment of

prostate adenocarcinoma leads to an improved mortality and morbidity, these findings

point out the importance of being able to differentiate potentially dangerous cancers from

the very small, well-differentiated, slow-growing lesions which are unlikely to present

clinically during the patient’s natural lifespan.

1.6.2.3 Etiology and Risk Factors

It has become clear that genetics play a significant role in the pathogenesis of

prostate adenocarcinoma. Male relatives of men who have died from prostate cancer

have a greater-than-expected incidence of the disease. An early study by Woolf of 228

men dying of prostate cancer found the relative nearly 3-fold increase in the relative risk

of first-degree relatives compared to a control group[75]. Subsequent studies have

confirmed this familial association[76-78], and demonstrated the importance of screening

PSA values in asymptomatic men from families with 3 or more members affected by

prostate cancer[74, 79].

Recent evidence supports the existence of a genuinely hereditary form of early

onset prostate cancer exhibiting Mendelian autosomal dominant inheritance[80]. The

exact gene defects have not been elucidated for these families but possible locations have

been mapped to chromosome 1q24-25[81] as well as the X chromosome suggesting the

Page 47: Fourier-Transform Infrared Spectroscopic Imaging of

35

possibility of X-linked inheritance[82]. Recent evidence suggest that mutations in the

tumor suppressor genes BRCA-1[83] and BRCA-2[84, 85] confer increased risk of

developing prostatic adenocarcinoma, and attempts to screen for those at risk are

currently being studied[86]. The most influential factor conferring risk of developing

prostate cancer besides familial inheritance is age[87]. African-American men have

roughly twice the lifetime risk of their white counterparts and higher PSA and tumor

volume in a study adjusted for age, stage, pathologic stage, Gleason score, and volume of

benign disease[88].

Other predisposing factors for clinical prostate cancer include the presence of

testosterone and dihydrotestosterone (DHT), sexual history positive for early first sexual

experience and multiple sexual partners[89], a diet high in saturated animal fat and low in

yellow and green vegetables, and environmental or occupational exposure to several

pollutants including cadmium[90] and the radioactive agents 51Cr, 59Fe, 60Co, and

65Zn[91]. Vasectomy has been suggested as a possible risk conferring event[92-94]

though some studies failed to demonstrate a conclusive link[95, 96].

1.6.2.4 Diagnosis

1.6.2.4.1 Clinical Presentation

With the recent widespread increase of PSA testing in men at risk for prostate

cancer, a large proportion of patients presenting with the disease are asymptomatic.

Clinically apparent prostate cancer presents with a spectrum of symptoms related to the

extent of disease progression. Urinary symptoms occur in localized as well as advanced

disease states as well as in extremely common condition of benign prostatic hyperplasia

Page 48: Fourier-Transform Infrared Spectroscopic Imaging of

36

(BPH). Symptoms related to bladder outflow obstruction, such as hesitancy, poor stream,

and a sensation of incomplete voiding arise from urethral occlusion by the tumor or

nodular mass. Urinary frequency and urgency are irritative symptoms that develop due to

detrusor muscle instability secondary to outflow obstruction or directly by tumor invasion

of the trigone of the bladder and pelvic nerves. Invasive cancer can produce other

symptoms both locally and at distant sites. Local extension of prostate cancer can present

with hematuria and/or hemospermia due to invasion of the prostatic urethra or seminal

vesicles. Direct invasion of the distal urinary sphincter can cause urinary symptoms

unrelated to outflow obstruction, while similar invasion of the neurovascular bundles

posteriorly can lead to erectile dysfunction and pain. Significant posterior invasion of

prostate cancer can produce lower bowel symptoms including rectal bleeding and

constipation due to large intestine obstruction near the rectum. Symptoms that indicate

local metastatic disease include bone pain, paraplegia due to cord compression, lymph

node enlargement, lower limb lymphedema, and loin pain while lethargy, cachexia, and

hemorrhage may indicate significant systemic metastases[97].

1.6.2.4.2 Digital Rectal Examination (DRE)

Digital rectal examination (DRE) is an inexpensive method of prostate cancer

detection which has been the focus of many clinical studies[98-103]. One problem with

the test is that it is subjective and consequently depends on the experience of the

examiner. Another is that several other conditions can lead to a false-positive DRE

finding, including BPH, prostatitis, prostatic calculi, ejaculatory duct anomaly, seminal

vesicle anomaly, and rectal wall phlebolith or polyp/tumor. Early stages of prostate

Page 49: Fourier-Transform Infrared Spectroscopic Imaging of

37

cancer (T2a) are characterized by a firm peripheral nodule that does not distort the

capsule, while more advanced cancers feel hard and more diffuse. T3 stage tumors often

present an altered prostate contour while retaining movement of the gland as a whole

contrasted with the fixed, immobile presentation of T4 stage tumors.

1.6.2.4.3 Prostate Specific Antigen (PSA)

Prostate-specific antigen is a 34 kD glycoprotein specifically found in prostate

epithelium. It is a neutral serine protease designed to lyse seminal-vesicle protein. A

small percentage of PSA normally escapes the prostatic ducts and enters the bloodstream

where it exists bound mainly to the proteins alpha-1-antichymotrypsin (ACT) and alpha

macroglobulin (αMG), leaving a small proportion of free PSA in the serum. Prostate-

specific antigen has established utility for the immunohistochemical identification of

metastatic disease of prostatic origin, for monitoring of “biochemical recurrence” after

therapy and for assessment of disease status in men who are at high risk for biopsy

complications.

Screening measures for serum PSA levels have increased the detection rate of early-

stage prostate cancer and are thought to be in part responsible for the downward stage

migration trend seen in the disease. Considerable variability exists in the world of PSA

testing. The cutoff for normal total PSA is accepted to be 4.0 ng/mL though some

evidence suggests lowering this cutoff in at risk populations. While most clinical assays

measure total PSA (bound + free) a significant advantage is afforded when an additional

test for free PSA is performed. Strong evidence exists that PSA complexed with ACT

increases in prostatic carcinoma[104, 105] and the lack of availability of a test to

Page 50: Fourier-Transform Infrared Spectroscopic Imaging of

38

specifically measure serum ACT-complexed PSA led to the use of percent free-to-total

PSA ratio to approximate complexed PSA[106]. Such ratios proved to be especially

useful in the population of men with total PSA values in the ‘gray zone’ of 2.5 to 10

ng/ml[107]. Recent development of a reliable assay for ACT-PSA complex[108] looks

promising and may outperform both total PSA and free-to-total PSA ratio as a more

specific analyte for cancer[109]. Other methods to improve PSA performance that have

been studied include PSA density[110, 111], transitional zone density[112], PSA

velocity[113, 114], and age-specific PSA[115].

1.6.2.4.4 Diagnostic Imaging

Transrectal ultrasound imaging (TRUS) produces high-resolution images of the

prostate which are useful for assessing extent of tumor involvement and extension as well

for guiding needle biopsies to sample areas suspected of harboring tumor foci. Prostate

cancers are frequently hypoechoic on TRUS, but can also be isoechoic and more rarely

hyperechoic[116]. Characteristics of prostate cancer that can be evaluated by TRUS

include asymmetry of prostate size, shape, indefinite differentiation between the central

and peripheral zones, and bulging or disruption of the capsule. Advances in color

Doppler TRUS allowing analysis of abnormal blood flow look promising for the

identification of hypervascular regions in the peripheral zone[117]. Computed

tomography (CT) scanning is useful in metastatic disease to identify the presence of

lymphadenopathy in the pelvis and is suggested only when other factors identify risk of

tumor spread (i.e. PSA>20ng/mL and Gleason grade > 7)[118]. Advances in Magnetic

resonance (MR) imaging endorectal coil design[119] have allowed the acquisition of

Page 51: Fourier-Transform Infrared Spectroscopic Imaging of

39

high-resolution differentially weighted MR images of prostatic disease that are probably

the most accurate technique currently available for assessing the extent of tumor

involvement. Additionally, dynamic contrast enhanced MR imaging may provide tumor

angiogenesis information[120].

1.6.2.5 Biopsy Interpretation and Grading of Prostatic Adenocarcinoma

The definitive diagnosis of prostatic adenocarcinoma involves the cytological and

histological confirmation of the established criteria of malignancy. The diagnostic

criteria for carcinomas in biopsies of the prostate involve both architectural and cytologic

findings[121]. Low to medium power analysis of the arrangement of the glandular acini

is useful and is the basis of the Gleason scale for grading prostatic adenocarcinoma, the

predominant scoring system used in the United States[122]. Malignant acini are typically

scattered haphazardly in the stroma either singly or in clusters. The acini in cancer are

typically small to medium sized with contours that are less smooth than adjacent normal

and hyperplastic acini. Cytologic abnormalities in adenocarcinoma include nuclear and

nucleolar enlargement present in a majority of malignant cells. Nucleolar size greater

than 1.5 mm suggests malignancy while identification of two or more nucleoli in a single

cell is virtually diagnostic of malignancy[123].

1.6.2.5.1 Gleason Grading System

The Gleason Grading system is the most widely used system for grading prostatic

adenocarcinoma. It relies heavily on the examination of low power architectural features

of the arrangement of prostatic acini. The Gleason scale rates glandular patterns of

proliferation on a scale of 1 (most differentiated) to 5 (least differentiated). Most prostate

Page 52: Fourier-Transform Infrared Spectroscopic Imaging of

40

cancers contain more than one of these patterns and thus the Gleason score for a biopsy

interpretation is reported as the combination of the two most prominent patterns. Scores

range from 2-10 and should be reported as the composite score and its component

patterns with the most prevalent pattern listed first[124]. For example a biopsy sample

with a predominant pattern of 3 and a secondary pattern of two would be reported as

3+2=5. In practice most cancers have at least one score of 3, and the score of 1 is rarely

used.

Gleason grade 1 architecture is described as very well differentiated and is

minimally distorted. Neoplastic glands are round, closely packed, single, separate,

uniform in shape and diameter, and are sharply delineated from fibrovascular stroma.

Hyperplastic glands also fulfill these criteria, therefore a classification as grade 1

adenocarcinoma also requires occasional enlarged nucleoli > 1mm in diameter. In

practice a Gleason score of 1 is rarely used. Gleason grade 2 pattern (well differentiated)

consists of glands which still exhibit a mild but definite stromal separation between

glands with more variation in the shape and size of glands than is seen in grade 1, but less

than that of grade 3. Grade 2 tumors remain circumscribed, and definite separation of the

malignant glands exists at the tumor periphery suggesting ability to spread to the

surrounding stroma. Tumor gland separation is usually less than one average gland

diameter. Gleason grade 3 cancers exhibit more extreme variation in size, shape, and

separation than grade 2 and are typically spaced more than one average gland diameter

apart. The cytoplasm of grade 3 tumor cells tends to be more basophilic than lower grade

cancers and nuclei are variable but still larger than lower grades and almost always

contain prominent nucleoli. Gleason grade 4 cancers may exhibit any of 4 different

Page 53: Fourier-Transform Infrared Spectroscopic Imaging of

41

morphologic patterns. Glands with a cribiform pattern have large masses of tumor cells

punctuated by sieve-like spaces. Such a pattern was classified as grade 3 by Gleason,

however, subsequent reclassification to grade 4 was based on the conclusion that most, if

not all examples of cribiform carcinoma are equivalent to grade 4 carcinoma growing

within preexisting lumina[125]. The distinctive feature of grade 4 tumors is ragged and

invading edges in contrast to the smooth edges of grade 3. Other architectural variants of

grade 4 adenocarcinoma include solid, microacinar, and papillary. Gleason grade 5

tumors completely lack glandular differentiation. Such tumors can be arranged in solid

masses, cords, trabeculae, sheets, or may appear as single cells infiltrating the stroma.

1.6.2.5.2 Importance of Histologic Grading

Cancer grade at time of diagnosis has been investigated extensively for correlations

with other tumor characteristics and clinical behavior. Every measure of survival and

recurrence is strongly correlated with cancer grade. These measures include crude

survival, tumor-free survival after treatment, metastasis-free survival, and cause-specific

survival. Such correlation has been described and validated in numerous studies[126-

129]. Age-adjusted, fifteen-year, cancer-specific mortality rates for men with Gleason

scores of 2 -4, 5, 6, 7 , and 8-10 are 4-7%, 6-11%, 18-30%, 42-70%, and 60-87%

respectively[130]. Tumor volume has been correlated with histologic grade in both

transurethral and radical prostatectomy specimens. A study by McNeal showed that in

Gleason grade 4 and 5 tumors, 22 of 38 tumors >3.2 cm3 had tumor-positive nodes while

positive nodes were present in only 1 out of 171 tumors <3.2 cm3. Two studies

Page 54: Fourier-Transform Infrared Spectroscopic Imaging of

independently confirmed that the strongest predictor of progression of poorly

differentiated cancer is tumor volume[129, 131].

Other studies have found correlations between Gleason grade and PSA levels[132].

Gleason grade is also one of the strongest and most useful predictors of pathologic stage

in many studies including the progression of capsular perforation, seminal vesicle

invasion, and lymph node and bone metastases and can be correlated with expression

levels of MIB-1 (Ki-57), a tissue marker for proliferation[133-136].

1.6.2.6 Staging of Prostatic Adenocarcinoma

Accurate assessment of the clinical stage of prostatic adenocarcinoma is important

for the estimation of prognosis, selection of treatment, and evaluation of therapeutic

results. The Tumor Node Metastasis (TNM) staging system is used to stage prostatic

adenocarcinoma. The current TNM clinical staging is shown below in tables 1.2 and 1.3.

TX Primary tumor cannot be assessed T0 No evidence of primary tumor T1 Clinically inapparent tumor not palpable or visible by imaging T1a Tumor incidental histological finding in 5% or less of tissue resected T1b Tumor incidental histological finding in more than 5% of tissue resected T1c Tumor identified by needle biopsy. Nonpalpable, not visible in imaging. T2 Tumor confined within the prostate T2a Tumor involves one lobe T2b Tumor involves both lobes T3 Tumor extends through the prostate capsules T3a Unilateral extracapsular extension T3b Bilateral extracapsular extension T3c Tumor invades the seminal vesicle(s) T4 Tumor invades any of bladder neck, external sphincter, or rectum T4a Tumor invades any of bladder neck, external sphincter, or rectum T4b Tumor invades levator muscles and/or the pelvic wall

adapted from [69] Table 1.2 - Staging of primary tumor (T)

42

Page 55: Fourier-Transform Infrared Spectroscopic Imaging of

NX Regional lymph nodes cannot be assessed N0 No regional lymph node metastasis N1 Metastasis in regional node(s)

adapted from [69]

Table 1.3 - Staging of regional lymph node involvement (N)

43

Page 56: Fourier-Transform Infrared Spectroscopic Imaging of

44

Chapter Two - Methods

2.1 Tissue microarrays

Tissue microarray technology provides a platform for the high throughput analysis

of tissue speciemens in research[137]. They are used for the target verification of cDNA

microarray results[138], expression profiling of tumors and tissues[139], as well as

epidemiology based investigations. Well-designed tissue arrays reduce the variability of

experiments performed in a repetitive fashion on large populations, and provide

consistent sample-to-sample preparation.

There are currently no reported studies applying vibrational spectroscopic imaging

techniques to the analysis of tissue microarray specimens. The tissue microarray is an

attractive sample platform for pathological spectroscopic imaging approaches for several

reasons. First, tissue arrays can be constructed from archival material, allowing for large

sample populations representative of normal tissue and disease processes to be examined.

Second, tissue microarrays provide consistent sample preparation across a large sample

population, minimizing sample-to-sample data variation. Finally, serial sections of tissue

microarrays can be analyzed with other techniques to provide complementary

information invaluable to the interpretation of spectroscopic imaging results.

2.1.1 Construction of Prostate Tissue Microarrays

Sections from three prostate tissue microarrays constructed in the Tissue Array

Research Program Laboratory, Laboratory of Pathology, Center for Cancer Research, of

Page 57: Fourier-Transform Infrared Spectroscopic Imaging of

45

the National Cancer Institute by Dr. Stephen M. Hewitt were used as samples for the

experiments in this study. The tissue array donor material was obtained from formalin-

fixed paraffin-embedded blocks from radical prostatectomy specimens taken from cases

of confirmed prostate adenocarcinoma from specimens obtained from the Cooperative

Human Tissue Network (CHTN) with approval of the appropriate Institutional Review

Boards or Office of Human Research Subjects. The tissue arrays were constructed with a

0.6 mm needles[139]. The arrays were constructed using a Beecher Instruments (Silver

Spring, MD) ATA-27 Automated Tissue Arrayer.

For sake of clarity, the arrays will be referred to by the respective patient

populations used for their construction. Specific details regarding the layout of Array P-

16, Array P-40 and Array P-80 appear in the sections below.

2.1.2 Array P-16 Design

Array P-16 was constructed using donor tissue from a population of 16 patients

with confirmed prostate adenocarcinoma. Eight unmapped 0.6 mm cores from each

patient were used for a maximum spot number of 128 spots/section. Donor core

locations were determined by examination of H&E stained sections of the donor blocks

and were chosen to provide a representative sampling of both normal prostate histology

and pathology from each patient.

2.1.3 Array P-40 Design

Array P-40 was constructed from donor tissue from a population of 40 patients that

included the set 16 patients used in the construction of Array P-16. Five unmapped 0.6

mm cores from each of the forty patients were used for a maximum spot number of 200

Page 58: Fourier-Transform Infrared Spectroscopic Imaging of

46

spots/section. Donor core locations were chosen from locations representative of both

adenocarcinoma and benign epithelium.

2.1.4 Array P-80 Design

Array P-80 was constructed of donor tissue from a population of 79 patients with

confirmed adenocarcinoma. Two mapped 0.6 mm cores were used from each patient for

a maximum spot number of 160 spots/section. H&E-stained sections of the donor tissue

blocks were used as a guide to carefully select tissue from a region of adenocarcinoma

for one core and benign epithelium for the corresponding core. Figure 2.1 below contains

an image of an H&E stained section of Array P-80 and a corresponding schematic

representation of the core layout.

Page 59: Fourier-Transform Infrared Spectroscopic Imaging of

47

2019181716151413121110987654321

2019181716151413121110987654321

20201919181817171616151514141313121211111010998877665544332211

20201919181817171616151514141313121211111010998877665544332211

4039383736353433323130292827262524232221

4039383736353433323130292827262524232221

6059585756555453525150494847464544434241

6059585756555453525150494847464544434241

8079787776757473727170696867666564636261

8079787776757473727170696867666564636261

4039383736353433323130292827262524232221

4039383736353433323130292827262524232221

6059585756555453525150494847464544434241

6059585756555453525150494847464544434241

8079787776757473727170696867666564636261

8079787776757473727170696867666564636261

Adeno-carcinoma

Cores

Benign Epithelium

Cores

Adeno-carcinoma

Cores

Benign Epithelium

Cores

600 µm

2019181716151413121110987654321

2019181716151413121110987654321

20201919181817171616151514141313121211111010998877665544332211

20201919181817171616151514141313121211111010998877665544332211

4039383736353433323130292827262524232221

4039383736353433323130292827262524232221

6059585756555453525150494847464544434241

6059585756555453525150494847464544434241

8079787776757473727170696867666564636261

8079787776757473727170696867666564636261

4039383736353433323130292827262524232221

4039383736353433323130292827262524232221

6059585756555453525150494847464544434241

6059585756555453525150494847464544434241

8079787776757473727170696867666564636261

8079787776757473727170696867666564636261

Adeno-carcinoma

Cores

Benign Epithelium

Cores

Adeno-carcinoma

Cores

Benign Epithelium

Cores

600 µm

Figure 2.1 - Array P-80 Layout. The right panel contains a visible optical image of an H&E stained section of array P-80. A schematic representation of the core layout appears on the left with patient numbers.

2.2 Tissue Array Section preparation

2.2.1 Optical Substrates for Tissue Array Sections

Standard optical materials, such as those found in microscope slides, are generally

composed of glass, quartz or fused silica. These materials all absorb radiation in the

infrared region at wavelengths longer than 2 µm. For this reason, transmission

experiments in the mid-IR require the use of alternative optical materials. Several

different halide salts are commonly used as optical materials for IR spectroscopy and

each possess different optical and physical properties[42].

Page 60: Fourier-Transform Infrared Spectroscopic Imaging of

48

Tissue array sections intended for IR imaging experiments were mounted on 3 mm-

thick, polished, barium fluoride (BaF2) optical windows. Barium fluoride is transparent

from 0.15-12.5 µm, which covers the visible and the entire spectral range of the FT-IR

instrument. Additionally, BaF2 optical elements have the lowest solubility in water (0.

17gm/100gm water at 23 °C) of materials with similar optical characteristics[2].

2.2.2 Deparaffinization

Histology grade, low melt (58-62 ºC) paraffin was removed from tissue array

sections by covering the tissue surface with hexane for 5 minutes. The samples were

rinsed with hexane several times and deparaffinization was continued by immersion in

hexane at 40ºC with continuous stirring for 48 hours. Every 3-4 hours during the

deparaffinization process, the immersion vessel was emptied, rinsed thoroughly with

acetone followed by hexane. Once dry, the vessel was filled with fresh neat hexane to

promote flow of embedded paraffin from the tissue. Thorough deparaffinization was

assured by monitoring the disappearance of the paraffin band at 1462 cm-1 at several sites

on the tissue arrays.

2.2.3 Optical imaging of H&E sections

Tissue array sections contiguous with those used for IR imaging analysis were

mounted on glass slides and stained with hematoxylin and eosin for traditional

histopathological analysis. H&E stained tissue microarray sections were optically

imaged using an Olympus BH-2 microscope equipped with a high resolution (3

megapixel), Peltier cooled, 10-bit, Q-imaging micropublisher digital camera. Tissue

array spots were imaged individually through a 4x Olympus ∞-corrected microscope

Page 61: Fourier-Transform Infrared Spectroscopic Imaging of

49

objective and 10x camera eyepiece objective. The H&E sections were reviewed with a

pathologist for diagnostic features on a two-headed teaching microscope.

2.3 Spectroscopic Imaging Instrumentation

FT-IR spectroscopic imaging was performed on a Perkin-Elmer (Shelton, CT)

Spectrum Spotlight 300 imaging microspectrometer equipped with dual mode detection

system. The imaging system is comprised of two main optical components: the

spectrometer and microscope. The spectrometer houses a ceramic globar broadband

infrared source, a continuous-scanning Michelson interferometer, and a macro sampling

area for non-microscopic single point FT-IR measurements. The modulated infrared

output beam of the spectrometer is coupled to an infrared microscope and focused onto

the sample using Cassegrain optics.

In transmission mode, the infrared beam is focused onto the sample through a

Cassegrain condenser. The condenser position can be varied along the beam axis to

correct for optical effects caused by substrate thickness and refractive index. Transmitted

radiation is collected by a Cassegrain objective and focused onto one of two mercury

cadmium telluride (MCT) detectors. Figure 2.1 shows a detailed diagram of the

microscope and the optical path in transmission mode.

A traditional single mercury cadmium telluride (MCT) detector is used in the

instrument’s point mode to take single-point spectroscopic measurements. Seamless

software control of variable-width, rotating knife-edge apertures, and a motorized

mapping stage with a precision error < 1 µm allow flexible collection of high quality

infrared spectroscopic mapping data in point detection mode. The microscope portion of

Page 62: Fourier-Transform Infrared Spectroscopic Imaging of

50

the instrument features a visible LED light source and video camera that are linked with

control software for automated collection of visible-light images as shown in figure 2.1.

Self-referenced stage position is dynamically linked with both captured visible images

and spectroscopic imaging results. This feature allows the operator to choose sample

areas for spectroscopic imaging experiments via a simple interface by selecting a

rectangular area on the displayed optical visible image of the sample, and provides

interactive registration of infrared spectroscopic imaging results with corresponding

optical visible images.

The spotlight 300’s image mode utilizes a 16-element MCT linear array detector to

build infrared spectroscopic images of any designated rectangular sample area in a line-

mapping fashion. A fixed optical zoom allows the instrument to collect image data at

two different spatial resolutions. The effective pixel size of these two resolution are 25 x

25 µm in low resolution and 6.25 x 6.25 µm at high resolution.

Page 63: Fourier-Transform Infrared Spectroscopic Imaging of

beam from spectrometer1

2

3

45

6

47

Perkin-Elmer Spectrum Spotlight 300 - Microscope Assembly Features

1) Dual-mode Mercury Cadmium Telluride (MCT) detector.

2) Visible CCD camera for optical image acquisition

3) Knife-edge apertures for single point measurements

4) Dichroic mirrors allow for a common infrared and visible beam path.

5) Z-fold allows variable pixel resolution at the sample.

6) High-precision sample stage linked directly to the interferometer allows for synchronized scanning and flexible image acquisition

7) LED visible illumination source

beam from spectrometer11

22

33

4455

66

447

Perkin-Elmer Spectrum Spotlight 300 - Microscope Assembly Features

1) Dual-mode Mercury Cadmium Telluride (MCT) detector.

2) Visible CCD camera for optical image acquisition

3) Knife-edge apertures for single point measurements

4) Dichroic mirrors allow for a common infrared and visible beam path.

5) Z-fold allows variable pixel resolution at the sample.

6) High-precision sample stage linked directly to the interferometer allows for synchronized scanning and flexible image acquisition

7) LED visible illumination source

adapted from [140]

Figure 2.2 - Spectrum Spotlight 300 Microscope Optical Configuration

The 1 GB of RAM in the controlling computer limits the size of a single line-

mapping image cube acquisition in the imaging mode. The maximum sample area size

that can be collected is thus a function of several collection parameters including spatial

resolution (high or low), spectral resolution, and spectral wavelength range. Practical

considerations such as the liquid nitrogen dewar hold time of 7 hr can also limit the

maximum size of image data collection in practice.

2.3.1 Tissue Array FT-IR Data Collection Parameters

IR Spectroscopic images of the tissue array spots were collected in transmission

configuration in image mode at the high-resolution zoom setting (pixel size of 6.25µm).

1641 data points were collected across the spectral region from 4000-720 cm-1 yielding

51

Page 64: Fourier-Transform Infrared Spectroscopic Imaging of

52

spectra with a resolution of 4 cm-1 (2 cm-1 data point interval). Four interferograms were

co-added for each individual measurement to increase data signal-to-noise ratios (SNRs).

Background spectra consisting of 190 coadded interferograms were collected from

nearby locations on the BaF2 flats between the tissue spots.

Data collection with these parameters for a typical 600 µm tissue array spot results

in a spectroscopic imaging data set with spatial dimensions of ~115 x 115 pixels and a

file size of approximately 85 MB. Acquisition time for a typical tissue array spot was

approximately 35-40 min. The average SNR for a single pixel absorbance spectrum of

tissue was >500:1.

2.3.2 Modifications and Environmental Considerations

The microscope and spectrometer assemblies were enclosed in a Plexiglas housing

to enable efficient purging with dry nitrogen gas to remove water vapor and to eliminate

air currents. The computer controlling the system was situated outside the housing and

the exhaust streams from the cooling fans of the spectrometer (source) and microscope

(detector electronics) were vented out of the housing to maintain a stable room

temperature atmosphere within the housing during data collection. Once the sample was

placed on the stage, all positioning, focusing, and experimental control could be

performed remotely by computer control without opening the housing to the atmosphere.

After opening the housing for any reason, 20 minutes were allowed for atmospheric

equilibration before spectroscopic measurements were resumed.

Page 65: Fourier-Transform Infrared Spectroscopic Imaging of

53

2.4 Data Handling and Computational Considerations

2.4.1 Data Pre-Processing

In its imaging mode, the Spectrum Spotlight 300 makes use of the dead time while

the microscope stage is stepped to a new position to perform several computational tasks.

The functions include interferogram apodization, fast Fourier transform of collected data

to single beam spectra, and ratioing of sample spectra to background spectra to provide

absorbance spectra. Spectroscopic imaging data of tissue array spots were collected

individually or in small contiguous groups, checked for spectral quality (SNR, baseline

fluctuations, etc.), and corrected for atmospheric water vapor and carbon dioxide using

Perkin Elmer proprietary software.

The resulting, atmosphere-corrected, spectroscopic images were imported into

ENVI (RSI inc., Boulder, CO) using software written in IDL by Dr. Rohit Bhargava; all

subsequent image processing was performed in this software environment. Some

downstream statistical analyses and chart plotting were performed using Microsoft Excel

and Origin. All processing was carried out computers equipped with 1.7 GHz Intel

Pentium 4 processors and a minimum of 1 GB of RAM.

Individual tissue array spots were mosaicked into one large spectroscopic image

dataset for each individual array section for further processing. For Array P-16, the final

size of the whole-array spectroscopic image was ~ 500 x 3680 pixels (or ~1.8 million

individual spectra) producing a file size of ~14 GB. Spectroscopic image datasets of the

two sections of Array P-40 were ~ 4370 x 550 pixels or (or ~2.4 million individual

Page 66: Fourier-Transform Infrared Spectroscopic Imaging of

54

spectra) with a file size of ~17 GB. Array P-80 had a final size of 2160x1250 pixels (or

~2.7 million spectra) with a file size of ~18.5 GB.

2.4.2 Spectral Baseline Correction

Every infrared absorbance spectrum in the image scene was individually baseline

corrected using custom-designed routines written in IDL by Dr. Rohit Bhargava.

Regression is used to calculate the values that lie on the line-segment intersecting each

pair of points. These values are subsequently subtracted from the spectral absorbance at

the corresponding frequency, and the process is repeated for each spectrum in the image

scene. Several hundred average spectra from different tissue regions on multiple spots

of Array P-16 were compared and frequency positions observed to be consistent local

minima were chosen as baseline points. A list of the frequency positions used as spectral

baseline points appears in Table 2.1.

Page 67: Fourier-Transform Infrared Spectroscopic Imaging of

982

1184

1144

1296

948

1328

1352

1478

1764

1984

2282

2392

2542

2644

2708

3000

3682

3774

spectral baseline points

(cm-1)

982

1184

1144

1296

948

1328

1352

1478

1764

1984

2282

2392

2542

2644

2708

3000

3682

3774

spectral baseline points

(cm-1)

Table 2.1 - Spectral frequencies used for spectroscopic baseline correction

The baseline-corrected absorbance intensity of the N-H stretching protein backbone

vibration (or Amide A) at 3290 cm-1 was used to differentiate tissue from empty space on

the array. All pixels with an absorbance less than 0.08 at 3290 cm-1 were masked to zero

for all spectral data points and disregarded during any subsequent processing.

55

Page 68: Fourier-Transform Infrared Spectroscopic Imaging of

56

Chapter Three - Infrared Spectroscopic Histology of Prostate

3.1 Visualization of Spectral Images and Verification of Histologic Features

Infrared spectroscopic imaging datasets of prostate tissue microarray sections were

initially visualized by plotting images of the baseline-corrected absorbance at 3290 cm-1.

This wavenumber position corresponds to the N-H stretching absorbance band or Amide

A absorbance, a backbone vibration found in all proteins. Since proteins are basic

structural elements of all prostate tissue, Amide A absorbance images are useful for

verifying the presence of spots and structural correlation of features with visible optical

images of the corresponding H&E stained section. The baseline corrected Amide A

absorbance images for 4 tissue array spots from a single patient are shown in fig 3.1A

along with a corresponding H&E stained consecutive section in Fig 3.1B.

Page 69: Fourier-Transform Infrared Spectroscopic Imaging of

57

A B

0.25

0.20

0.15

0.10

0.05

AB

SO

RB

AN

CE

IN

TE

NS

ITY

A B

0.25

0.20

0.15

0.10

0.05

AB

SO

RB

AN

CE

IN

TE

NS

ITY

Figure 3.1- A) Baseline-corrected N-H stretching (3290cm-1) absorbance intensity image of four tissue array spots from a single patient on Array P-16 B) Optical images of corresponding H&E stained section.

The tissue microarray sections used for IR spectroscopic imaging experiments are

subject to harsh deparaffinization conditions of immersion in hexane at 40ºC for 4 hours.

These conditions caused artifactual damage to a handful of spots in each array sections.

Typical artifactual problems included partial or complete absence of spots, spots that

folded over onto themselves, and spots which were partially detached from the surface of

Page 70: Fourier-Transform Infrared Spectroscopic Imaging of

58

the optical flat. N-H stretching absorbance images such as those seen in figure 3.1A were

extremely useful for discovering spots that were subject to such damage so that they

could be eliminated from further analysis.

3.2 Creation of Ground Truth Data Regions of Interest

In order to analyze spectra and to train and test classification models, ground truth

data for different histological features or classes needed to be established. The name

ground truth stems from remote sensing applications where field data from various

sources on the ground are acquired and registered with image data to enable class training

and/or evaluation of classification performance[61].

A pathologist examined the matching H&E stained tissue array sections

microscopically and different histological features present in each spot were marked on

optical images of the corresponding H&E stained sections. The region of interest (ROI)

tool in ENVI allows the user to designate a collection of pixels as belonging to a set, or

ROI. ROIs can be manually generated by selecting geometric areas on the spectroscopic

images with drawing tools such as rectangles, ellipses, or polygons. Pixels may be added

to or deleted from ROIs individually, allowing the user to carefully edit such groups.

ROIs can also be generated from parameters of the data itself, which can be particularly

useful. Once created, these ROIs can be used in a variety of image analysis operations

from image subsetting and masking to statistical analyses and image classification.

In analyzing the spectroscopic datasets, specific images derived from various

absorbance band ratios provided high contrast for discerning different histologic features

in the tissue. Fig 3.2A shows the 1080 cm-1/1544 cm-1 absorbance band ratio image of

Page 71: Fourier-Transform Infrared Spectroscopic Imaging of

59

four tissue array spots from a single patient on Array P16. The 1080 cm-1 band is

attributed to a C-O stretching vibration of glycogen and the band at 1544 cm-1 to the

Amide II vibration of the protein backbone. The 1080 cm-1/1544 cm-1 image provides

high contrast between prostate epithelium and stroma. Areas of higher ratio intensity in

Fig 3.2A correspond to the basophilic-staining epithelial regions in the optical image of

the corresponding H&E stained section in panel B. The eosinophilic stromal regions of

the tissue correspond to lower intensity regions of the 1080 cm-1/1544 cm-1 ratio image

suggesting that glycogen/protein levels are higher in epithelial tissue than in stroma.

Another absorbance band ratio that produced useful images was 1206 cm-1/1544

cm-1. At the spectral resolution used of 4 cm-1, the absorbance feature at 1206 cm-1

typically appears as a shoulder off the higher intensity combination band at 1236 cm-1

attributed to both Amide III vibrational mode of proteins and the asymmetric stretch of

phosphodiester (PO2-) groups in phospholipids and nucleic acids. Fig 3.2C shows the

1206 cm-1/1544 cm-1 absorbance band ratio image of 4 tissue array spots taken from a

different patient on Array P-16. Comparison with the image of the matching H&E

stained section (Fig 3.2D) reveals poor contrast between epithelial and stromal tissues,

however, excellent contrast is seen between an area of lymphocytic infiltration, indicated

by the highest intensity area in the upper spot, and the surrounding stromal and epithelial

components.

Page 72: Fourier-Transform Infrared Spectroscopic Imaging of

DA B C

100 µm

H&E Absorbance Ratio1210/1544 cm-1 H&E

100 µm

0.25

0.20

0.15

0.10

0.05

INT

ENS

ITY

Absorbance Ratio1080/1544 cm-1

0.08

0.06

0.04

0.02

0.00

INT

ENS

ITY

INT

ENS

ITY

0.300

0.225

0.150

0.075

0.000

DA B C

100 µm

H&E Absorbance Ratio1210/1544 cm-1 H&E

100 µm

0.25

0.20

0.15

0.10

0.05

INT

ENS

ITY

Absorbance Ratio1080/1544 cm-1

0.08

0.06

0.04

0.02

0.00

0.08

0.06

0.04

0.02

0.00

INT

ENS

ITY

INT

ENS

ITY

0.300

0.225

0.150

0.075

0.000

Figure 3.2 - Absorbance Band Ratio Images of tissue array spots from Array P-16

Various absorbance band images and band ratio images were interactively overlaid

and used to assist the ROI creation process. Using the pathologist-reviewed, marked

optical images of the H&E stained sections as a guide, collections of pixels in the

spectroscopic image of each tissue spot were assigned to one of the ten histological class

ROIs listed in table 3.1. The epithelial class includes pixels from different

histopathological states, including normal benign epithelium, benign prostatic hyperplasia

(BPH), prostatic intraepithelial neoplasia (PIN), and prostatic adenocarcinoma (CaP).

60

Page 73: Fourier-Transform Infrared Spectroscopic Imaging of

Stromal histological features were separated into 3 subclasses: fibrous stroma, smooth

muscular stroma, and mixed stroma based on the H&E section images and spectral

differences noted between these three subclasses. Remaining classes included sites of

lymphocytic infiltration, vessel endothelium and muscular coat, peripheral nerve tissue,

ganglion cells, blood cells, and corpora amylacea. In making the component analysis,

much care was taken to include only those pixels that were definitively representative of

a particular class, and therefore pixels near edges or class borders were eliminated to

insure that class spectral statistics remain uncontaminated.

number of spectra in class ROI

162956

1039

628

438

2362

359

1976

2751

74609

11444

80293

16 patient array

1134lymphocytes

153554Total

828corpora amylacea

767blood cells

0ganglion cells

0peripheral nerve

54endothelium

560smooth muscle stroma

30144mixed stroma

19092fibrous stroma

1134epithelial tissue

40 patient arrayHistologic Class

number of spectra in class ROI

162956

1039

628

438

2362

359

1976

2751

74609

11444

80293

16 patient array

1134lymphocytes

153554Total

828corpora amylacea

767blood cells

0ganglion cells

0peripheral nerve

54endothelium

560smooth muscle stroma

30144mixed stroma

19092fibrous stroma

1134epithelial tissue

40 patient arrayHistologic Class

Table 3.1 - Histologic class population data

Class data were stored separately for each spot and histologic class as individual

regions of interest (ROI) in ENVI and could be operated on individually at the spot level

or merged to patient level or into a single ROI at the class level. This flexibility enables

downstream comparisons to be made at the spot-spot and patient-patient level for each

class and across classes.

61

Page 74: Fourier-Transform Infrared Spectroscopic Imaging of

62

3.3 Spectral analysis of histologic features and metric selection

The individual ROIs from each spot were merged together to form a single large

ROI for each of the ten histologic classes for each array. The total number of pixels,

where each pixel represents an individual spectrum, is shown for each histologic class in

table 3.1. The spectra from each ROI were averaged to create a mean spectrum for each

class, displayed in figure 3.3.

Page 75: Fourier-Transform Infrared Spectroscopic Imaging of

Nor

mal

ized

Abso

rban

ce

Wavenum ber (cm-1)-1)1400 1300 1200 1100 1000

D

Nor

mal

ized

Abs

orba

nce

Wavenum ber (cm-1)-1)3600 3400 3200 3000 2800

Nor

mal

ized

Abs

orba

nce

Wavenum ber (cm-1)-1)1750 1700 1650 1600 1550 1500

B

C

3500 3000 2500 2000 1500 1000Wavenum ber (cm-1)

Nor

mal

ized

Abso

rban

ce A

-1)

AEPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

B

C

D

Nor

mal

ized

Abso

rban

ce

Wavenum ber (cm-1)-1)1400 1300 1200 1100 1000

D

Nor

mal

ized

Abso

rban

ce

Wavenum ber (cm-1)-1)1400 1300 1200 1100 1000

D

Nor

mal

ized

Abs

orba

nce

Wavenum ber (cm-1)-1)3600 3400 3200 3000 2800

Nor

mal

ized

Abs

orba

nce

Wavenum ber (cm-1)-1)1750 1700 1650 1600 1550 1500

B

C

3500 3000 2500 2000 1500 1000Wavenum ber (cm-1)

Nor

mal

ized

Abso

rban

ce A

-1)

A

3500 3000 2500 2000 1500 1000Wavenum ber (cm-1)

Nor

mal

ized

Abso

rban

ce A

-1)

AEPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

EPITHELIUMFIBROUS STROM AM IXED STROM ASM OOTH M USCLENERV EGANGLION C ELLSBLOODLYM PHOCYTESCORPORA AM YLACEAENDOTHELIUM

B

C

D

Figure 3.3 - Histologic class mean spectra

The spectra were calculated from baseline corrected spectra and were normalized to amide II absorbance at 1544cm-1. Panel A contains the full spectral window collected 720-4000 cm -1. Panels B, C, and D contain enlargements of the corresponding boxes in panel A.

63

Page 76: Fourier-Transform Infrared Spectroscopic Imaging of

64

3.4 Construction of a Supervised Classification Model for Prostate Histology

3.4.1 Spectral Data Reduction

The mean spectra for each histologic class were compared and spectral features,

frequencies, and band ratios could be identified for distinguishing the various classes

from one another. A set of metrics was developed involving absorbance band ratios and

peak centers of gravity for features across the entire spectral region. Metric values were

computed using software routines written in the statistical language IDL by Dr. Rohit

Bhargava and implemented in the remote sensing software environment ENVI (RSI, inc.,

Boulder, CO).

Histograms of each training class population were plotted and compared for each

metric. Most distributions approximated a normal distribution and showed some

variation in mean and standard deviation between classes. Metrics which did not

approximate a normal distribution for most classes were discarded, since such data can

lead to poor performance with parametric classification methods, particularly with

Gaussian Maximum Likelihood classification algorithms[62] discussed below in section

3.4.2 . Metrics that showed no significant variation between classes were also discarded,

since their inclusion would likely add only noise to the classification. The spectroscopic

imaging dataset was reduced from 1641 spectral bands (wavenumber positions) to a 20-

band set of candidate spectral metrics, reducing the tissue array imaging dataset from 14

GB to a manageable 160 MB.

The construction of successful classification model is by nature an interactive,

process. Information is gained in small bits as individual problems are identified and

Page 77: Fourier-Transform Infrared Spectroscopic Imaging of

65

strategies are altered to adjust. A common problem encountered is the existence of

classes which possess bimodal distributions in several spectral bands. Such observations

typically indicate that the class is composed of two or more spectrally distinct subclasses.

In such cases, classification accuracy can often be dramatically improved by splitting the

training data for the suspect class into separate classes[63]. Similar histogram analysis

performed on several absorbance band ratio images from early FT-IR imaging studies of

non-array prostate tissue indicated that stromal tissue in the prostate was composed of

spectrally distinct subclasses. These preliminary results formed the basis for splitting

stroma into three separate subclasses: fibrous stroma, smooth muscular stroma, and

mixed fibromuscular stroma.

A listing of the parameters for each of the 20 candidate spectral metrics appears

below in Table 3.2.

Page 78: Fourier-Transform Infrared Spectroscopic Imaging of

MaxMinDenominator Band

Numerator Band

1062154412121080

13901034120615441544154415441544154415441544(cm-1)

1034145012361164

140010501210123616543292966

1034106210801114(cm-1)

Band Ratio Parameters

Center-of-GravitySpectral Region

Band Ratio20Band Ratio19Band Ratio18Band Ratio17

15601478Center of Gravity1617641572Center of Gravity1536823000Center of Gravity1412961184Center of Gravity1314201372Center of Gravity12

Band Ratio11Band Ratio10Band Ratio9Band Ratio8Band Ratio7Band Ratio6Band Ratio5Band Ratio4Band Ratio3Band Ratio2Band Ratio1

(cm-1)(cm-1)

Type of MetricMetric #MaxMinDenominator

BandNumerator

Band

1062154412121080

13901034120615441544154415441544154415441544(cm-1)

1034145012361164

140010501210123616543292966

1034106210801114(cm-1)

Band Ratio Parameters

Center-of-GravitySpectral Region

Band Ratio20Band Ratio19Band Ratio18Band Ratio17

15601478Center of Gravity1617641572Center of Gravity1536823000Center of Gravity1412961184Center of Gravity1314201372Center of Gravity12

Band Ratio11Band Ratio10Band Ratio9Band Ratio8Band Ratio7Band Ratio6Band Ratio5Band Ratio4Band Ratio3Band Ratio2Band Ratio1

(cm-1)(cm-1)

Type of MetricMetric #

Table 3.2 - Histology Spectral Metric Definitions

3.4.2 Image Classification

Several different algorithms exist for the supervised classification of multispectral

image data. Some of the more simplistic classification algorithms such as parallelpiped

or minimum-distance approaches do not consider variation that may be present within

spectral classes and do not perform well when frequency distributions from separate

classes overlap[62].

Histogram analysis of individual metric value class distributions indicated that both

significant intraclass variation in spectral metric values exist and that significant overlap

between metric value frequency distributions of different classes was common. As

examples, individual class histograms for the three most common or populated training 66

Page 79: Fourier-Transform Infrared Spectroscopic Imaging of

classes (epithelium, mixed stroma, and fibrous stroma) are displayed for metric 02 values

(Fig. 3.4A) and for metric 11 values (Fig 3.4B).

0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5

FREQ

UEN

CY (

norm

aliz

ed to

cla

ss s

ize)

METRC 11 VALUE (band ratio 1400/1390cm-1)

Epithelium Mixed Stroma Fibrous Stroma

0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24

FREQ

UE

NC

Y (no

rmal

ized

to c

lass

siz

e)

METRIC 02 VALUE (band ratio 1080/1544cm-1)

Epithelium Mixed Stroma Fibrous Stroma

A

B

0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5

FREQ

UEN

CY (

norm

aliz

ed to

cla

ss s

ize)

METRC 11 VALUE (band ratio 1400/1390cm-1)

Epithelium Mixed Stroma Fibrous Stroma

0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24

FREQ

UE

NC

Y (no

rmal

ized

to c

lass

siz

e)

METRIC 02 VALUE (band ratio 1080/1544cm-1)

Epithelium Mixed Stroma Fibrous Stroma

A

B

Figure 3.4- Histograms of metric value class frequency distribution for the three

most populated classes (epithelium, mixed stroma, & fibrous stroma) for: A) Metric 02 (band ratio 1080/1544cm-1), and B) Metric 11 (band ratio 1400/1390 cm-1)

A parametric approach to supervised classification that is particularly well suited to

deal with such natural intraclass spectral variation and interclass overlap of metric

frequency distributions is the Gaussian Maximum Likelihood (GML) Classifier[62]. An 67

Page 80: Fourier-Transform Infrared Spectroscopic Imaging of

68

n-dimensional probability surface for each class is generated from both class mean and

variance statistics for training data consisting of n spectral bands. As classification

ensues, each pixel’s discrete spectrum can be used to calculate the corresponding

conditional probability or likelihood that the pixel belongs to each class separately from

the individual class n-dimensional probability surfaces[63]. The pixel is then assigned to

the class with the highest conditional probability. Classification sensitivity can be

adjusted by imposing minimum probability thresholds that cause pixels below a user-

supplied minimum conditional probability to be relabeled as unclassified.

A supervised Gaussian Maximum Likelihood (GML) algorithm implemented in

ENVI was used to classify the 20 metric dataset of the entire tissue array. The 10

different histologic class ROIs were used as input to train the classifier. No thresholding

was imposed during the classification forcing each pixel in the tissue array image scene

to be classified as one of the ten histologic subtypes. The 10 histologic training ROIs

were used next as a preliminary validation set to evaluate the performance of the

classification.

3.4.3 Array P-16, 20-metric, GML Self-classification results

An extremely useful tool for the evaluation of image classification results is the

expression of classification accuracy in terms of an error matrix. Such error matrices are

also commonly referred to (appropriately) as confusion matrices. Error matrices

compare, on a class-by-class basis, the relationship between known reference data

(ground truth class data) and the corresponding results of a classification attempt.

Page 81: Fourier-Transform Infrared Spectroscopic Imaging of

The same 10 histology class ROIs from Array P-16 that were used as training input

for the 20-metric, Gaussian Maximum Likelihood (GML) classification of all tissue on

Array P-16 were used as ground truth to calculate an error matrix for the same

classification. These data appear below as Table 3.3.

69

0.000.000.560.000.000.250.000.140.03BLOOD

0.001.690.280.050.000.000.070.055.68GANGLION

0.001.370.560.050.000.003.920.030.79NERVE

11.460.460.420.000.002.290.851.160.12ENDOTHELIUM

0.000.000.000.000.000.000.000.003.48LYMPHOCYTES

0.000.000.000.000.000.000.000.010.01CORPORA AMYLACEA

0.640.000.043.900.000.000.7323.180.00SMOOTH MUSCLE

0.000.003.560.000.000.000.694.820.25FIBROUS STROMA

0.160.000.000.280.000.000.400.130.00MIXED STROMA

0.000.460.170.282.070.190.000.110.18EPITHELIUM

BLO

OD

GAN

GLIO

N

NER

VE

END

OTH

ELIUM

LYMPH

OC

YTES

CO

RPO

RA

AMY

LACE

A

SMO

OTH

MU

SCLE

FIBR

OU

S STROM

A

MIXED

STROM

A

EPITHELIU

M

Ground Truth Class

Result of Classification

87.74

97.72

94.12

94.15

97.82

99.81

96.36

94.19

70.42

89.64

87.74

97.72

94.12

94.15

97.82

99.81

96.36

94.19

70.42

89.64

0.000.000.560.000.000.250.000.140.03BLOOD

0.001.690.280.050.000.000.070.055.68GANGLION

0.001.370.560.050.000.003.920.030.79NERVE

11.460.460.420.000.002.290.851.160.12ENDOTHELIUM

0.000.000.000.000.000.000.000.003.48LYMPHOCYTES

0.000.000.000.000.000.000.000.010.01CORPORA AMYLACEA

0.640.000.043.900.000.000.7323.180.00SMOOTH MUSCLE

0.000.003.560.000.000.000.694.820.25FIBROUS STROMA

0.160.000.000.280.000.000.400.130.00MIXED STROMA

0.000.460.170.282.070.190.000.110.18EPITHELIUM

BLO

OD

GAN

GLIO

N

NER

VE

END

OTH

ELIUM

LYMPH

OC

YTES

CO

RPO

RA

AMY

LACE

A

SMO

OTH

MU

SCLE

FIBR

OU

S STROM

A

MIXED

STROM

A

EPITHELIU

M

Ground Truth Class

Result of Classification

Table 3.3 - Error Matrix of supervised GML Classification results using 20 spectroscopic metrics

The classifier was implemented in ENVI and was trained on sets of reference spectra assigned to one of ten histologic classes. All matrix values are given in units of percent of ground truth class pixels.

The columns represent the ground truth or correct class designation and the rows

represent the result class as assigned by the GML classifier. The numbers at each

position are the percent of the number of total pixels in the column (ground truth class)

that were classified as the class of the row. For example, if we examine the epithelium

column, we see that 89.64% of epithelial pixels were correctly classified, 0.25 % of

epithelial pixels were misclassified as fibrous stroma, 3.48% of epithelial pixels were

Page 82: Fourier-Transform Infrared Spectroscopic Imaging of

70

misclassified as lymphocytes, etc. The values that occupy the diagonal of the confusion

matrix (shown in red in Table 3.3) are the classification accuracy for a given class. These

values show that this initial classification attempt performs above 94% for all classes

except for epithelium (89.6%), mixed stroma (70.4%), and blood (87.7%).

3.4.4 Leave-one-out metric evaluation

It was clear from the histogram analysis of the individual metrics in the original set

of 20 that certain metrics were better for discriminating certain classes than others. In

light of the significant frequency distribution overlaps seen in many cases, a given

metric’s inclusion in the classification attempt might provide little to modest increase in

classification accuracy for single class or small number of classes while causing a

significant decrease in accuracy in the remaining classes. To test for the presence of such

contaminating metrics in the original set of 20, a leave-one-out analysis was performed.

The image scene was reclassified 20 separate times using a total of 19 spectral metrics

per attempt, leaving out a different metric for each successive trial. The accuracy change

for the 3 classes with the worst 20-metric classification accuracy (epithelium, mixed

stroma, and blood) with respect to the 20 metric classification was recorded for each

successive trial and is shown below in Figure 3.5.

Page 83: Fourier-Transform Infrared Spectroscopic Imaging of

-5.5

-3.5

-1.5

0.5

2.5

4.5

6.5

8.5

10.5

12.5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Metric Left Out

Perc

ent c

hang

e in

acc

urac

y

Epithelium

Mixed Stroma

Blood

Figure 3.5 - Graphical Representation of results of the leave-one-out analysis The tissue array data was reclassified 20 separate times with a total of 19 metrics, sequentially leaving out a different metric. The percent change in classification accuracy for the three histologic classes which performed poorly in the 20 metric classification attempt (epithelium, mixed stroma, and blood) are plotted with the metric number left out varying along the x-axis.

While the results of the leave one out analysis were analyzed for every class, for

the sake of clarity, Figure 3.5 contains only data from the three classes (epithelium,

mixed stroma, and blood) which had the most classification error in the original 20-

metric classification. These three classes stand to benefit most from the removal of a

possible contaminating metric, and from the results in Fig 3.5 we see that two metrics

clearly stood out as detrimental to classification accuracy. All three of the poorly

classified classes (epithelium, mixed stroma, and blood) show a significant increase in

accuracy when metric 9 and metric 18 are left out individually.

71

Page 84: Fourier-Transform Infrared Spectroscopic Imaging of

3.4.5 Array P-16, 18-metric GML Classification Results

Reclassification with the GML algorithm in the absence of both metric 9 and

metric 18 produced the most promising self-evaluated results, which are represented as a

error matrix in Table 3.4. The training ROIs were used as ground truth input to generate

the error matrix results.

72

1.740.000.000.000.000.030.022.77GANGLION

0.001.140.560.000.000.004.100.010.48NERVE

1.270.680.550.000.002.070.900.470.07ENDOTHELIUM

0.000.000.000.000.000.000.000.000.92LYMPHOCYTES

0.000.000.000.000.000.000.000.020.01CORPORA AMYLACEA

0.640.000.044.180.000.000.945.530.00SMOOTH MUSCLE

0.000.003.680.000.000.000.691.150.19FIBROUS STROMA

0.480.000.000.280.000.002.760.790.00MIXED STROMA

0.000.910.300.283.850.190.000.210.16EPITHELIUM

BLO

OD

GAN

GLIO

N

NER

VE

END

OTH

ELIUM

LYMPH

OC

YTES

CO

RPO

RA

AMY

LACE

A

SMO

OTH

MU

SCLE

FIBR

OU

S STROM

A

MIXED

STROM

A

EPITHELIU

M

Ground Truth Class

Result of Classification

97.61

97.26

93.69

92.76

96.15

99.81

94.04

93.04

92.51

95.55

0.000.001.950.000.000.440.000.130.00BLOOD

0.00

97.61

97.26

93.69

92.76

96.15

99.81

94.04

93.04

92.51

95.55

0.000.001.950.000.000.440.000.130.00BLOOD

0.001.740.000.000.000.000.030.022.77GANGLION

0.001.140.560.000.000.004.100.010.48NERVE

1.270.680.550.000.002.070.900.470.07ENDOTHELIUM

0.000.000.000.000.000.000.000.000.92LYMPHOCYTES

0.000.000.000.000.000.000.000.020.01CORPORA AMYLACEA

0.640.000.044.180.000.000.945.530.00SMOOTH MUSCLE

0.000.003.680.000.000.000.691.150.19FIBROUS STROMA

0.480.000.000.280.000.002.760.790.00MIXED STROMA

0.000.910.300.283.850.190.000.210.16EPITHELIUM

BLO

OD

GAN

GLIO

N

NER

VE

END

OTH

ELIUM

LYMPH

OC

YTES

CO

RPO

RA

AMY

LACE

A

SMO

OTH

MU

SCLE

FIBR

OU

S STROM

A

MIXED

STROM

A

EPITHELIU

M

Ground Truth Class

Result of Classification

Table 3.4 - Confusion matrix of supervised GML Classification attempt using 18 spectroscopic metrics Metric 9 and metric 18 were left out of the original set of 20.

All 10 classes are classified at an accuracy above 92.5%. A color-coded classified

result image of four tissue array spots from a single patient are shown in figure 3.6 with

the corresponding H&E section in panel B for comparison. Classification

correspondence with the histological features observed in the H&E section is

outstanding.

Page 85: Fourier-Transform Infrared Spectroscopic Imaging of

EPITHELIUM

FIBROUS STROMA

MIXED STROMA

SMOOTH MUSCLE

NERVE

GANGLION CELLS

BLOOD

LYMPHOCYTES

CORPORA AMYLACEA

ENDOTHELIUM

EPITHELIUM

FIBROUS STROMA

MIXED STROMA

SMOOTH MUSCLE

NERVE

GANGLION CELLS

BLOOD

LYMPHOCYTES

CORPORA AMYLACEA

ENDOTHELIUM

Figure 3.6 - Classification results for 2 tissue array spots from the same patient

GML Classification was performed with a total of 18 metrics selected from the results of the leave-one-out analysis as shown in figure 3.5.

Epithelial pixels were classified correctly 95.6% of the time with the majority of

misclassification as ganglion (2.8%) and lymphocytes (0.9%). Mixed stroma pixels were

classified correctly 92.5% of the time with the majority of misclassification , not

surprisingly, as smooth muscle stroma (5.5%) and fibrous stroma (1.15%).

Fibrous stroma pixels were classified correctly 93% of the time with the major

misclassification predominately occurring as nerve (4%). Interestingly, nerve pixels were

correctly classified with an accuracy of 93.7% with the majority of misclassification as

fibrous stroma (3.7%). Upon close inspection, the mean spectra of the fibrous stroma

and nerve training ROI’s proved to have many similarities, as seen in figure 3.3. Spectral

similarities between nerve and fibrous stroma include absorbance peaks at 1034 cm-1 and 73

Page 86: Fourier-Transform Infrared Spectroscopic Imaging of

74

1206 cm-1, and a shoulder at 1280cm-1. As a result of this spectral similarity, a

substantial number of pixels at the stromal-epithelial interface were observed to be

misclassified as nerve when they probably belong to the fibrous stroma or mixed stroma

class.

Smooth muscle stroma pixels were classified with an accuracy of 94% with the

majority of misclassification as mixed stroma (2.8%) and endothelium (2%). Again of

note is that while endothelium was correctly classified 92.8% of the time, the majority of

misclassification occurred as smooth muscle stroma. While fibrous stroma and nerve

represent a pair of classes whose similarity seems likely based on a compositional

similarity, the connection between endothelium and smooth muscle stroma is probably

due to impurity in the endothelial training class. The endothelial training class by far had

the fewest number of training spectra at 359. This reflects both the paucity of discernible

endothelial tissue visible in prostate sections on H&E staining and the difficulty in

correctly identifying it in corresponding IR spectroscopic images. Endothelial cells are

typically very hard to identify as they are single-layered, and are contiguous with the

smooth muscular media which is more pronounced in arterial vessels.[7] With a single

pixel in the IR spectroscopic images representing 6.25 µm of tissue per edge, it seems

highly likely some of the endothelial training pixels are contaminated with signal from

smooth muscle tissue of the vessel media. Similarly, blood pixels were classified with an

accuracy of 97.6% with the majority of misclassification as endothelium.

Lymphocyte pixels were classified with an accuracy of 96.2% with all of the

misclassification as epithelial pixels (3.8%). A large proportion of pixels which were

incorrectly classified as lyphocytes probably represent true spectral mixtures of different

Page 87: Fourier-Transform Infrared Spectroscopic Imaging of

75

class types, since lymphocytic infiltration necessarily overlays regions of stroma and

epithelial tissue. Ganglion pixels were classified to an impressive 97.3% with the

majority of misclassification as nerve. Corpora amylacea were classified to an accuracy

of 99.8%. While this accuracy value seems aberrantly high compared with the other

classes, examination of the class mean spectrum of corpora amylacea compared with the

other class mean spectra (figure 3.3) reveals that it is quite extreme compared with every

other spectrum which probably accounts for the high self-classification accuracy.

3.5 Validation of Prostate Histology Classification Model

These impressive results with a simple set of 20 metrics hint at the promise of this

approach. One can be certain that many more metrics exist that if included would

improve classification accuracy. One of the many advantages of this approach is that we

can design our metrics to highlight the property of a spectral feature that is changing

across classes, whether it be band height relative to another band or band center of

gravity irrespective of height. Metrics which measure other spectral properties such as

absorbance band widths are other obvious choices to be tested in the future, while data

collection at higher spectral resolution and with higher single-pixel SNRs will uncover

newly resolvable spectral features which can be harnessed as metrics to improve

classification accuracy.

An important caveat mentioned prominently most remote sensing references [61-

63, 141] is that accuracy estimates made using training data regions as ground truth do

not necessarily indicate that similar results will be seen when classifying other regions of

the image scene. The pixels in the ROI sets used for classifier training and evaluation

Page 88: Fourier-Transform Infrared Spectroscopic Imaging of

76

make up only a tiny fraction of the total number of tissue pixels in the full spectroscopic

image of Array P-16. Several spots from Array P-16 were purposely avoided during the

training ROI selection process so that they could be used for qualitative validation of

promising classification results. Examination of these spots with respect to their

matching H&E stained sections gave a qualitative sense that the 18-metric classification

was performing quite well on tissue that was not included in the training sets. As an

example, the lower spot in Figure 3.6 contains no pixels used in any of the 10 training

ROIs, any the classification results agree well with the image of the matching H&E-

stained section.

3.5.1 Cross-Array Validation

As noted in table 3.2, a set of histology ground truth ROIs was constructed for the

spectroscopic imaging dataset of Array P-40 in the same manner as described in section

3.2 in reference to Array P-16.

In light of the observed classification trends seen in the 18 metric, P-16 training

data error matrix in Table 3.3 and discussed in section 3.5, adjustments were made to the

classification model class structure. The endothelial class was discarded due to

insufficient ground truth ROI pixel populations on both Array P-16 and Array P-40. The

extremely thin nature of this tissue structure on cross-section further adds to the difficulty

in both establishing ground truth information for this potential class and evaluating

results since pixels in the spectroscopic images have a size of 6.25 µm of tissue per pixel

edge. Visual analysis of the H&E-stained section of Array P-40 revealed almost no

contiguous areas of pure smooth muscle as seen frequently in Array P-16. Furthermore,

Page 89: Fourier-Transform Infrared Spectroscopic Imaging of

the 18-Metric self-classification results indicated that most of the misclassified mixed

stroma pixels were incorrectly classified as smooth muscle stroma and vice versa.

Consequently, the ground truth data smooth muscle stroma class and mixed stroma

classes were merged into a single mixed stroma class separately for both Array P-16 and

Array P-40. The spectral similarity and commission errors seen between the fibrous

stroma and nerve classes suggested they might also be better off combined as a single

fibrous-stroma class. However, no appreciable nerve or ganglion tissue was found in any

of the Array P-40 spots so both the P-16 nerve and ganglion training data were excluded

from the cross-array classification attempt. These adjustments to the histology class

structure result in a total of 6 classes. Table 3.5 contains the revised, 6-class, histology

ground truth class ROI set population data for both Array P-16 and Array P-40.

number of spectra in class ROI

162956

628

359

1039

11444

77360

80293

16 patient array

828corpora amylacea

153554Total

767blood cells

1134lymphocytes

19092fibrous stroma

30704mixed stroma

1134epithelial tissue

25 patient arrayHistologic Class

number of spectra in class ROI

162956

628

359

1039

11444

77360

80293

16 patient array

828corpora amylacea

153554Total

767blood cells

1134lymphocytes

19092fibrous stroma

30704mixed stroma

1134epithelial tissue

25 patient arrayHistologic Class

Table 3.5 - Revised 6-class histology ground truth ROIs for Array P-16 and Array P-40

The 6 ground truth ROIs for Array P-16 listed in Table 3.5 were used as training

data for supervised classification of all tissue from Array P-40. The same 18 metrics

used for classification of Array P-16 in section 3.5.5 were used for both the training data

from Array P-16 and for P-40 image data to be classified by the GML algorithm.

77

Page 90: Fourier-Transform Infrared Spectroscopic Imaging of

All pixels in the P-40 image scene were classified and the 6 ground truth class ROIs

from Array P-40 were used to construct an error matrix for the cross-array classification

result which appears below in Table 3.6.

78

0.000.000.000.040.01BLOOD

0.000.000.000.002.30LYMPHOCYTES

0.000.000.000.000.10CORPORA AMYLACEA

0.260.000.001.851.26FIBROUS STROMA

4.950.000.008.100.58MIXED STROMA

0.005.395.390.380.58EPITHELIUM

BLOO

D

LYMP

HOC

YTES

CORPO

RA

AMYLAC

EA

FIBRO

US STRO

MA

MIXED S

TRO

MA

EPITHELIUM

Ground Truth Class

Result of Classification

94.78

94.71

94.61

91.52

97.53

95.74

94.78

94.71

94.61

91.52

97.53

95.74

0.000.000.000.040.01BLOOD

0.000.000.000.002.30LYMPHOCYTES

0.000.000.000.000.10CORPORA AMYLACEA

0.260.000.001.851.26FIBROUS STROMA

4.950.000.008.100.58MIXED STROMA

0.005.395.390.380.58EPITHELIUM

BLOO

D

LYMP

HOC

YTES

CORPO

RA

AMYLAC

EA

FIBRO

US STRO

MA

MIXED S

TRO

MA

EPITHELIUM

Ground Truth Class

Result of Classification

Table 3.6 - Error Matrix for 6-Class, GML Classification Results

The classifier was trained on 6-class ground truth data from Array P-16 and applied to classify all tissue pixels in the image data from Array P-40. The same set of 18 spectral metrics used in section 3.5.5 were used for this classification.

The error matrix results indicate that classification accuracy in 5 out of 6 classes

exceeds 94.5%. Fibrous stroma was the class with the lowest classification accuracy at

91.5%, however, nearly all of such misclassified pixels were incorrectly classified as

mixed stroma. This result likely speaks more to the heterogeneity of stroma in general

than to any serious problems with the classification itself.

Page 91: Fourier-Transform Infrared Spectroscopic Imaging of

79

3.6 Conclusions and Further Directions

These results indicate that such a 6-class, supervised GML classification model can

be used to successfully segment spectroscopic images of unstained sections of prostate

tissue into useful histologic classes based on their spectral properties with respect to

spectral class information from a database of previously imaged tissue from a number of

patients. Histological class information obtained from such images is useful for image

display, however standard staining procedures are far cheaper and provide similar

information. FT-IR spectroscopic imaging data analyzed in this fashion can provide

histological image information from unstained specimens. Standard staining techniques

can interfere with other analytical techniques, such as immunohistochemistry and in situ

hybridization, as well as, nucleic acid recovery from laser capture microdissected

material[142].

The histological class information obtained could also be used to study

morphological relationships, such as epithelial/stromal density ratios in various different

states of normal prostate tissue, nodular hyperplasia (BPH), and varying grades of

prostatic adenocarcinoma[143, 144]. The supervised classification methods for providing

histological class information from IR spectroscopic imaging data developed in the above

sections are well-suited for automation, providing a means for rapid evaluation necessary

for high throughput analyses.

Furthermore, such histological classifications can be used as a tool for downstream

analysis of spectral information from epithelial tissue in an effort to further study the

infrared spectroscopic properties of benign prostate epithelial tissue and prostatic

adenocarcinoma in many patients. If reliable spectral indicators of disease presence and

Page 92: Fourier-Transform Infrared Spectroscopic Imaging of

80

progression can be found, then FTIR microspectroscopic imaging techniques can be used

as an objective tool to aid in the detection and diagnosis of prostatic adenocarcinoma.

The next section continues with some preliminary experiments using a third tissue array,

P-80, designed to investigate some of these issues.

Page 93: Fourier-Transform Infrared Spectroscopic Imaging of

81

Chapter Four - Infrared Spectroscopic Histopathology of Prostate

4.1 Classification strategy

Array P-80 is the most logical choice as a starting point for the analysis of

spectral features of populations of benign and malignant prostate epithelial tissue. Array

P-80 was constructed from formalin-fixed, paraffin-embedded tissue blocks cut from

radical prostatectomy specimens from population of 80 patients with confirmed prostatic

adenocarcinoma. The array was constructed with 2 cores from each patient, one from a

region of representative adenocarcinoma, and one from a region with only normal benign

epithelium. The intention of the array design was to provide a large patient population

and relatively even sampling of benign and malignant tissue for every patient.

The first step of the analysis will apply the histology classification developed in

section 3, using the class statistics from the P-16-Array training populations to train the

classifier. The histology classification results will be used along with the pathologist’s

interpretation of the matching H&E-stained section to designate separate ROIs for benign

and malignant epithelium for each patient. Mean spectra will be used to develop a large

set of candidate spectral metrics for distinguishing between benign epithelium and

adenocarcinoma. Spectral metrics that show a statistically significant difference between

the benign and adenocarcinoma populations will then be used in attempts to self-classify

Array-P-80 and cross-validate by classifying other arrays with training data from Array-

P-80.

Page 94: Fourier-Transform Infrared Spectroscopic Imaging of

82

4.2 Array P-80 H&E Stained Section Pathology Analysis

The H&E stained, matching section of Array P-80 was carefully reviewed with a

pathologist and each spot was evaluated for several histopathological parameters. Before

the review process, the visible optical image of each spot was printed on a separate sheet

of paper and used to record the pathologist’s comments during the review process. Each

spot was assessed initially for tissue preservation and preparation. Spots that contained

significant preparation artifact or no epithelium were removed from analysis. The

pathologist carefully characterized the remaining spots and detailed records were kept for

subsequent ROI creation and analysis. The pathological status of epithelial tissue in

each spot was considered individually and any epithelial tissue for which the pathological

or preparation status was at all questionable was marked on the optical images so that it

would not be considered in later analyses.

All regions of confirmed prostatic adenocarcinoma were individually assigned a

Gleason Grade of 1-5 indicating the predominant Gleason pattern seen in the spot. Once

the pathology analysis was complete, the results were tabulated and it was found that a

total 38 patients contained usable benign epithelial tissue 51 patients contained usable

regions of prostatic adenocarcinoma. A total of 25 patients from array P-80 contained

regions of both benign epithelium and confirmed prostatic carcinoma. Without the

corresponding benign tissue from the same patient as a control, any analysis of spectral

features of adenocarcinoma tissue would be questionable. For this reason, the full-

spectrum spectroscopic imaging datasets for the tissue array spots for these 25 patients

were mosaicked into a single spectroscopic image for faster processing during

downstream analyses.

Page 95: Fourier-Transform Infrared Spectroscopic Imaging of

4.3 Array P-80 Histology Classification Results

The 18 histology metrics used in sections 3.5.5 and 3.6.1 were calculated for the

new 25 patient image of array P-80 from the baseline-corrected full spectrum image data.

The same histology classification performed in section 3.6.1 as cross-array validation,

was applied instead to the 25-patient, 18-metric image of array P-80. The 6 histologic

class ROIs from array P-16 listed in table 3.4 were used as class training data for the

GML classification of the 18-metric image. The histology classification results for this

25-patient image are displayed in Figure 4.1. Figure 4.2 contains the corresponding

optical images from the matching H&E stained section.

ent 20ent 20PatiePatieent 20ent 20PatiePatie

EPITHELIUM

FIBROUS STROMA

MIXED STROMA

EPITHELIUM

FIBROUS STROMA

MIXED STROMA

700 µm700 µm

Benign Spot

Cancer Spot

Benign Spot

Cancer Spot

NER VE

GANGLION CELLS

BLOOD

BLOOD

LYMPHOCYTES

CORPORA AMYLAC EA

NER VE

GANGLION CELLS

BLOOD

BLOOD

LYMPHOCYTES

CORPORA AMYLAC EA

Patient 25

Patient 24

Patient 23

Patient 22

Patient 21

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

Patient 10

Patient 09

Patient 06

Patient 15

Patient 14

Patient 11 Patient 16Patient 01

Patient 20Patient 05

Patient 19Patient 04

Patient 25

Patient 24

Patient 23

Patient 22

Patient 21

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

Patient 10

Patient 09

Patient 06

Patient 15

Patient 14

Patient 11 Patient 16Patient 01

Patient 20Patient 05

Patient 19Patient 04 Example

Figure 4.1- Array P-80 histology classification results

83

Page 96: Fourier-Transform Infrared Spectroscopic Imaging of

Patient 25

Patient 24

Patient 23

Patient 22

Patient 21

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

Patient 10

Patient 09

Patient 06

Patient 15

Patient 14

Patient 11 Patient 16Patient 01

Patient 20Patient 05

Patient 19Patient 04

Patient 25

Patient 24

Patient 23

Patient 22

Patient 21

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

Patient 10

Patient 09

Patient 06

Patient 15

Patient 14

Patient 11 Patient 16Patient 01

Patient 20Patient 05

Patient 19Patient 04

Figure 4.2 - Optical images of H&E stained section of Array P-80

4.3.1 Spatial Filtering of Histology Classification Results

The collected datasets have an effective pixel size of 6.25 µm x 6.25 µm. The

spectral data are collected over the wavelength range from 4000-720 cm-1, or 2.5-13.8

µm. A given pixel will therefore contain some spectral information from tissue locations

represented in the image data by neighboring pixels. Since most cells also have a size

within or near the spectral wavelength range of radiation, some misclassification can be

attributed to spectral bleeding from neighboring pixels that contain a different class of

tissue. As expected, this phenomenon is most prevalent along borders between different

histologic classes. Additionally, though the histological GML classification performs

quite accurately, it is after all, a model and like all models has an inherent error rate.

84

Page 97: Fourier-Transform Infrared Spectroscopic Imaging of

85

While many types of spatial filtering techniques exist for digital image processing, the

nature of the classification results suggest a particular method is most applicable for

removing randomly distributed misclassified pixels. Spectroscopic image data and

spectral metric data both span continuous ranges of data values within a single image

plane. Most commonly applied spatial image filtering techniques work well with such

data and involve some type of spatially-dependent averaging of pixel values within a

defined local neighborhood of pixels.

The GML classifier assigns each pixel in the image scene to one of 6 discrete

classes. Each class is represented in the image results data by a unique integer values. In

such a case, the data values associated with each pixel do not form any sort of continuous

scale, thus any spatial filtering techniques that rely on averaging would produce

meaningless results. Several useful spatial filtering techniques have been developed for

image classification results. Some complex methods utilize the conditional probability

statistics developed during the application of the GML classification algorithm to analyze

individual pixels with respect to a defined local neighborhood of pixels.[62]

Two more simple operations, which also produce satisfactory spatial filtering

results, are the sieve operation and majority analysis. A sieve operation considers the

neighborhood of pixels around a center pixel of class X, and applies a group minimum

threshold. If the number of pixels in the neighborhood classified as X is less than the

group minimum, then the pixel is relabeled as unclassified. The process is repeated for

every pixel in the image. An alternative spatial filtering technique that can be applied to

improve the appearance of image classification results is a majority analysis. This

technique also considers a kernel, or set of neighboring pixels, which is rastered across

Page 98: Fourier-Transform Infrared Spectroscopic Imaging of

the image pixel-by-pixel. As the kernel moves, the center pixel is changed to the class

that occupies the majority of the kernel positions that do not contain unclassified pixels.

The weight of the center pixel can be changed in integer increments to alter the amount of

filtering applied. This technique provides effective filtering of randomly misclassified

pixels and changes them to the class that dominates the neighborhood. For this reason,

majority-filtered images appear smoother than sieve-filtered results, which contain more

unclassified pixels. Figure 4.2 contains results of these two different filtering strategies

on a small example region of a classified prostate histology image representing a typical

border between epithelium and mixed stroma.

111111111

111111111

Sieve Operation8 Nearest Neighbor

Group Min = 5

Majority Analysis3x3 Kernel

Even weighting

Raw Classification

Result

Epithelium

Mixed Stroma

Epithelium

Mixed Stroma

Figure 4.3 - Spatial filtering techniques for classified image results

The majority analysis produces results that are extremely smooth and preferable for

general classification image display, however, it is important to point out that the

86

Page 99: Fourier-Transform Infrared Spectroscopic Imaging of

87

majority analysis changes the class designation of pixels based solely on spatial

information without any spectral information whatsoever. At this stage in the data

analysis, the sieve method is much more appropriate precisely because it is subtractive.

Since the histology classification results will be used to construct epithelial ROIs for

downstream spectroscopic analyses, it is important that such populations be as spectrally

pure as possible.

The histology classification result image displayed in figure 4.1 was spatially

filtered using a sieve operation implemented in ENVI using a neighborhood of eight

pixels, and a group minimum threshold of five. The results were qualitatively compared

with the matching H&E section and found to provide satisfactory removal of randomly

misclassified pixels, while also removing questionable pixels near class boundaries.

Figure 4.3 contains images of the raw histology classification and post-sieve operation

classification image for patient 2 from array P-80.

Page 100: Fourier-Transform Infrared Spectroscopic Imaging of

Raw Histology

Classification

Sieve Results

8 Nearest Neighbor

Group Min – 5

EPITHELIUM

FIBROUS STROMA

MIXED STROMA

EPITHELIUM

FIBROUS STROMA

MIXED STROMANERVE

GANGLION CELLS

BLOOD

BLOOD

LYMPHOCYTES

CORPORA AMYLACEA

NERVE

GANGLION CELLS

BLOOD

BLOOD

LYMPHOCYTES

CORPORA AMYLACEA

200 µm 200 µm

Figure 4.4 - Sieve operation spatial filtering of histology classfication results for patient 2 from array P-80

4.4 Construction of a Supervised Classification Model for Prostate Pathology

4.4.1 Creation of pathology ground truth ROIs

The sieved histology classification results produced in section 4.3 were used as the

starting point for the designation of pathology ground truth ROIs for array P-80. First,

the sieved histology classification result for a given spot was compared with the

annotated optical image of the matching H&E stained section. The epithelial

classification result pixels that corresponded to epithelial tissue selected for use in the 88

Page 101: Fourier-Transform Infrared Spectroscopic Imaging of

marked optical H&E stained imaged were grouped into separate ROIs for benign

epithelium and prostatic adenocarcinoma for each patient for a total of 50 ROIs. An

image of the pathology ground truth ROIs and the corresponding number of pixels in

each ROI is shown in Figure 4.4.

AdenocarcinomaTotal = 42,239 Pixels

AdenocarcinomaTotal = 42,239 Pixels

Benign Epithelium Total = 19,492 Pixels

700 µm700 µm

Array P80 – Pathology Regions-of-Interest (ROI’s)

Individual ROI size in

pixels

885698510312696 3883662114320372001

10732224048031219 224830230382464690

289985350518532 400154983317741113

13737862314761479 2934260414343941464

Patient 25

Patient 24

Patient 23

Patient 22

Patient 211812 406

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

19

Patient 10

Patient 09

Patient 062751 1832

Patient 15

Patient 14

Patient 112276

Patient 16Patient 01849

Patient 20Patient 05

Patient 19Patient 04

58310991334

885698510312696 3883662114320372001

10732224048031219 224830230382464690

289985350518532 400154983317741113

13737862314761479 2934260414343941464

Patient 25

Patient 24

Patient 23

Patient 22

Patient 211812 406

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

19

Patient 10

Patient 09

Patient 062751 1832

Patient 15

Patient 14

Patient 112276

Patient 16Patient 01849

Patient 20Patient 05

Patient 19Patient 04

58310991334

Figure 4.5 - Array P-80 pathology ground truth ROIs

4.4.2 Pathology Spectral Data Reduction

The mean infrared absorbance spectrum for each ROI was created and normalized

to Amide II protein backbone absorbance at 1544 cm-1. These mean spectra were

compared and spectral features, frequencies, and band ratios could be identified for

distinguishing benign epithelial tissue from prostatic adenocarcinoma. A set of 54

candidate metrics was developed involving absorbance band ratios and peak centers of

gravity for features across the entire spectral region. A listing of the parameters for each

of the 54 candidate spectral metrics appears below in table 4.1.

89

Page 102: Fourier-Transform Infrared Spectroscopic Imaging of

32903060Band Ratio3832903064Band Ratio3932903078Band Ratio4032903084Band Ratio4132903180Band Ratio4232903192Band Ratio4332903202Band Ratio4432903214Band Ratio4532903226Band Ratio4632903232Band Ratio47

9821144Center of Gravity4811441182Center of Gravity4911821296Center of Gravity5013521426Center of Gravity5114781578Center of Gravity5215851718Center of Gravity53

14501390Band Ratio2414501400Band Ratio2510801032Band Ratio2610801016Band Ratio2712361208Band Ratio28

12361278Band Ratio3012361262Band Ratio29

14501080Band Ratio3316521080Band Ratio3432901080Band Ratio3532903044Band Ratio3632903050Band Ratio37

12361516Band Ratio3112361080Band Ratio32

15443082Band Ratio2015443290Band Ratio2115443450Band Ratio2214501426Band Ratio23

MaxMinDenominator Band

Numerator Band

1544154415441544154415441544154415441544154415441544154415441544154415441544(cm-1)

165215881562153615161502145013121278123612061170115811161080106210401012966

(cm-1)

Band Ratio ParametersCenter-of-GravitySpectral Region

30003682Center of Gravity54

Band Ratio19Band Ratio18Band Ratio17Band Ratio16Band Ratio15Band Ratio14Band Ratio13Band Ratio12Band Ratio11Band Ratio10Band Ratio9Band Ratio8Band Ratio7Band Ratio6Band Ratio5Band Ratio4Band Ratio3Band Ratio2Band Ratio1

(cm-1)(cm-1)

Type of MetricMetric #

32903060Band Ratio3832903064Band Ratio3932903078Band Ratio4032903084Band Ratio4132903180Band Ratio4232903192Band Ratio4332903202Band Ratio4432903214Band Ratio4532903226Band Ratio4632903232Band Ratio47

9821144Center of Gravity4811441182Center of Gravity4911821296Center of Gravity5013521426Center of Gravity5114781578Center of Gravity5215851718Center of Gravity53

14501390Band Ratio2414501400Band Ratio2510801032Band Ratio2610801016Band Ratio2712361208Band Ratio28

12361278Band Ratio3012361262Band Ratio29

14501080Band Ratio3316521080Band Ratio3432901080Band Ratio3532903044Band Ratio3632903050Band Ratio37

12361516Band Ratio3112361080Band Ratio32

15443082Band Ratio2015443290Band Ratio2115443450Band Ratio2214501426Band Ratio23

MaxMinDenominator Band

Numerator Band

1544154415441544154415441544154415441544154415441544154415441544154415441544(cm-1)

165215881562153615161502145013121278123612061170115811161080106210401012966

(cm-1)

Band Ratio ParametersCenter-of-GravitySpectral Region

30003682Center of Gravity54

Band Ratio19Band Ratio18Band Ratio17Band Ratio16Band Ratio15Band Ratio14Band Ratio13Band Ratio12Band Ratio11Band Ratio10Band Ratio9Band Ratio8Band Ratio7Band Ratio6Band Ratio5Band Ratio4Band Ratio3Band Ratio2Band Ratio1

(cm-1)(cm-1)

Type of MetricMetric #

Table 4.1 - Pathology spectral metric parameters

90

Page 103: Fourier-Transform Infrared Spectroscopic Imaging of

4.4.3 Histogram analysis of Spectral Metric Data

Initial metric evaluation was conducted plotting histograms of different pathology

ground truth ROIs for individual metrics. Histograms analyzed on a patient-to-patient

basis revealed that for many metrics, a similar directional shift in the means of frequency

distributions between benign and adenocarcinoma populations was present. For many of

these metrics, while the direction of the shift was consistent from patient-to-patient, the

absolute values of the respective distributions varied quite significantly among patients.

This situation is depicted schematically in figure 4.5.

Benign EpitheliumAdenocarcinomaBenign EpitheliumAdenocarcinoma

0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

Metric Value0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

Patient 1 Patient 2 Patient 3

Benign EpitheliumAdenocarcinomaBenign EpitheliumAdenocarcinomaBenign EpitheliumAdenocarcinomaBenign EpitheliumAdenocarcinoma

0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

Metric Value0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

Metric Value0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

0.02 0.04 0.06 0.08 0.10 0.12 0.14

Freq

uenc

y (N

orm

aliz

ed)

Patient 1 Patient 2 Patient 3

Figure 4.6 - Patient-to-patient metric variation

It was clear that many of these metrics were providing information regarding real

spectral differences between benign and cancerous prostate tissue, however, the

significant patient-to-patient variation rendered these metrics ineffective for use in

parametric classification attempts.

91

Page 104: Fourier-Transform Infrared Spectroscopic Imaging of

92

4.4.4 Mean-centering of epithelial metric data.

The data were mean-centered in order to make use of the spectroscopic information

contained in the metrics affected by significant patient-to-patient variation and to

simplify the process of metric evaluation. The mean metric spectrum for each patient’s

benign ground truth ROI was calculated by averaging the individual metric-spectra

within each ROI. The discrete 54-metric spectrum of each individual epithelial pixel was

divided by the mean benign metric spectrum from the corresponding patient. This

calculation has the effect of normalizing the benign population distributions for all

patients individually for each metric. Thus, all patient-patient variation among benign

metric distributions is effectively collapsed such that recalculation of the benign 54-

metric spectrum for any patient would yield a value of one at every metric.

4.4.5 Metric Statistical Analysis

A major advantage of mean-centering the metric data is that it simplifies the task of

identifying which metrics provide statistically significant discrimination between benign

and adenocarcinoma patient populations. The mean 54-metric spectrum for each

patient’s adenocarcinoma ground truth ROI population was recalculated from the benign

mean-centered metric data. A one population t-test was applied to each of 54 sets of 25

patient-mean metric values to determine if the 25 patient population was significantly

different from the constant 1.0 at the 0.05 level. The results of the t-test for each metric

and the associated p-values are listed in table 4.2.

Page 105: Fourier-Transform Infrared Spectroscopic Imaging of

Significantly different from 1.0?p-valuet-valueVarianceMeanMetric #

NO0.5010.6840.0031.00738NO0.1551.4680.0021.01439NO0.1761.3930.0031.01540NO0.1781.3880.0031.01441NO0.979-0.0270.0021.00042NO0.907-0.1190.0020.99943NO0.835-0.2100.0010.99844NO0.775-0.2890.0010.99845NO0.678-0.4210.0010.99746NO0.634-0.4820.0010.99747YES0.0013.8063.016x10-061.00148YES0.0152.6308.991x10-071.00049NO0.271-1.1273.754x10-071.00050NO0.0522.0492.437x10-071.00051YES0.0102.7861.663x10-071.00052NO0.5780.5643.885x10-071.00053

NO0.5670.5810.0091.01124NO0.2831.0980.0081.02025YES0.003-3.2640.0070.94626YES9.7x10-6-5.5780.0130.87027NO0.5090.6710.0041.00828

NO0.4390.7870.0141.01930NO0.135-1.5480.0030.98429

NO0.1451.5050.0271.04933NO0.2451.1920.0111.02534YES0.0312.2980.0111.04735NO0.3320.9910.0031.01036NO0.8080.2460.0031.00337

YES0.001-3.6530.0100.92831NO0.343-0.9670.0130.97832

NO0.425-0.8120.0030.99120YES0.013-2.6820.0020.97721NO0.419-0.8220.0170.97822YES0.0013.8730.0091.07323

-0.444

-0.085-4.0870.559-3.179-3.597-2.681-0.727-1.1692.0953.0992.715-3.023-3.0921.5041.0861.175-0.277-5.0992.463

3.350x10-06

0.0014.4x10-4

0.0011.8x10-4

0.0020.0060.0070.0290.0320.0070.0160.0120.0380.0350.0140.0150.0110.0170.078

NO0.6611.00054

NO0.9331.00019YES4.2x10-40.98318NO0.5821.00317YES0.0040.99116YES0.0010.96915YES0.0130.96014NO0.4740.98813NO0.2540.96112YES0.0471.07511YES0.0051.05210YES0.0121.0699YES0.0060.9348YES0.0050.8797NO0.1461.0566NO0.2881.0265NO0.2511.0294NO0.7830.9943YES0.0030.8652YES0.0211.1371

Significantly different from 1.0?p-valuet-valueVarianceMeanMetric #

NO0.5010.6840.0031.00738NO0.1551.4680.0021.01439NO0.1761.3930.0031.01540NO0.1781.3880.0031.01441NO0.979-0.0270.0021.00042NO0.907-0.1190.0020.99943NO0.835-0.2100.0010.99844NO0.775-0.2890.0010.99845NO0.678-0.4210.0010.99746NO0.634-0.4820.0010.99747YES0.0013.8063.016x10-061.00148YES0.0152.6308.991x10-071.00049NO0.271-1.1273.754x10-071.00050NO0.0522.0492.437x10-071.00051YES0.0102.7861.663x10-071.00052NO0.5780.5643.885x10-071.00053

NO0.5670.5810.0091.01124NO0.2831.0980.0081.02025YES0.003-3.2640.0070.94626YES9.7x10-6-5.5780.0130.87027NO0.5090.6710.0041.00828

NO0.4390.7870.0141.01930NO0.135-1.5480.0030.98429

NO0.1451.5050.0271.04933NO0.2451.1920.0111.02534YES0.0312.2980.0111.04735NO0.3320.9910.0031.01036NO0.8080.2460.0031.00337

YES0.001-3.6530.0100.92831NO0.343-0.9670.0130.97832

NO0.425-0.8120.0030.99120YES0.013-2.6820.0020.97721NO0.419-0.8220.0170.97822YES0.0013.8730.0091.07323

-0.444

-0.085-4.0870.559-3.179-3.597-2.681-0.727-1.1692.0953.0992.715-3.023-3.0921.5041.0861.175-0.277-5.0992.463

3.350x10-06

0.0014.4x10-4

0.0011.8x10-4

0.0020.0060.0070.0290.0320.0070.0160.0120.0380.0350.0140.0150.0110.0170.078

NO0.6611.00054

NO0.9331.00019YES4.2x10-40.98318NO0.5821.00317YES0.0040.99116YES0.0010.96915YES0.0130.96014NO0.4740.98813NO0.2540.96112YES0.0471.07511YES0.0051.05210YES0.0121.0699YES0.0060.9348YES0.0050.8797NO0.1461.0566NO0.2881.0265NO0.2511.0294NO0.7830.9943YES0.0030.8652YES0.0211.1371

93

Table 4.2 - Results of t-test on mean adenocarcinoma metric values from population of 25 patients on array P-80 for 54 candidate pathology metrics

Page 106: Fourier-Transform Infrared Spectroscopic Imaging of

94

The t-test results indicate that 20 metrics from the candidate set of 54 pathology

metrics show statistically significant deviation between their respective populations of

adenocarcinoma and patient-matched benign epithelial pixels.

4.4.6 GML Pathology Classification of Array P-80

The set of 20 pathology metrics identified in section 4.2.8 for discriminating

between benign and malignant prostate tissue were used as data for GML classification of

all epithelial pixels from P-80 ground truth epithelial ROIs. The benign ground truth ROI

sets for all 25 patients were merged into one large benign training ROI comprised of

19,492 total pixels. Likewise, the adenocarcinoma ground truth ROI sets for all 25

patients were merged into one large adenocarcinoma training ROI comprised of 42,239

total pixels. A supervised 2-class (benign epithelium & adenocarcinoma) classification

was implemented in ENVI using the 20 metrics identified in section 4.2.8.

The classification image results for all 25 patients appear below in Figure 4.6.

Page 107: Fourier-Transform Infrared Spectroscopic Imaging of

Patient 25

Patient 24

Patient 23

Patient 22

Patient 21

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

Patient 10

Patient 09

Patient 06

Patient 15

Patient 14

Patient 11 Patient 16Patient 01

Patient 20Patient 05

Patient 19Patient 04

Patient 25

Patient 24

Patient 23

Patient 22

Patient 21

Patient 17Patient 12Patient 07Patient 02

Patient 18Patient 13Patient 08Patient 03

Patient 10

Patient 09

Patient 06

Patient 15

Patient 14

Patient 11 Patient 16Patient 01

Patient 20Patient 05

Patient 19Patient 04

Epithelium 700 µmAdenocarcinoma

Figure 4.7 - Array P-80 pathology classification results

The ground truth training ROIs were used to construct an error matrix to evaluate

the classification results on a whole-array basis. The error matrix appears below in

Figure 4.3.

95

Page 108: Fourier-Transform Infrared Spectroscopic Imaging of

96

BENIGN EPITHELIUM

ADENO

CAR

CINOM

A

BEN

IGN EPITH

ELIUM

Ground Truth Class

Result of Classification

74.50

89.59

10.41ADENOCARCINOMA

25.50

74.50

89.59

10.41ADENOCARCINOMA

25.50BENIGN EPITHELIUM

ADENO

CAR

CINOM

A

BEN

IGN EPITH

ELIUM

Ground Truth Class

Result of Classification

Table 4.3 - Error matrix for 20-metric pathology GML classification of

epithelial tissue on array P-80

These classification results give a sense that in general, the classifier is performing

adequately for distinguishing benign epithelium from regions of adenocarcinoma. While

very little misclassification of ground truth benign pixels is seen, there are a handful of

adenocarcinoma spots in Figure 4.8 which seem to be classified with less certainty than

the remainder of the patients.

4.5 Individual Patient Evaluation of P-80 Pathology Classification

The 20-metric pathology classification results were analyzed next on an individual

patient basis. For each of the fifty pathology ground truth ROIs (25 benign + 25

adenocarcinoma), the percentage of ROI pixels classified as adenocarcinoma for each

ROI was plotted as a bar chart in Figure 4.7 below.

Page 109: Fourier-Transform Infrared Spectroscopic Imaging of

Individual patient analysis of Pathology Classification Results

0

10

20

30

40

50

60

70

80

90

100

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26

Patient Number

Perc

ent o

f RO

I pix

els

clas

sifie

d as

A

deno

carc

inom

a

Benign Epithelium

Adenocarcinoma

Figure 4.8 - Individual patient analysis of 20-metric GML pathology classification

The data reveals that the pixels from the benign ROIs of all 25 patients were

classified with an accuracy > 80%. Imposing a minimum threshold for adenocarcinoma

classification of 20% on the data in figure 4.7 provides 100% discrimination between foci

benign and malignant epithelial tissue across the entire population of 25 patients.

4.6 Cross-Array Validation

Again it must be noted that such self-evaluation of training data ROIs represents the

best possible scenario for producing accurate supervised results. To examine the cross-

array performance of the pathology classification model, training data from array p-80

was used to classify mean-centered 20-metric data from other arrays.

Upon the pathologist’s review of the H&E stained section, arrays P-16 and P-40

were each found to contain 5 patients with usuable regions of both benign epithelium.

Initial cross-array classification attempts did not yield consistent results. While some

97

Page 110: Fourier-Transform Infrared Spectroscopic Imaging of

98

individual patients yielded results similar to those seen with Array P-80, the limited

population sizes of five patients on each array made it impossible to draw any substantive

conclusions regarding the cross-array performance of the developed pathology

classification model.

4.7 Conclusions and Further Directions

These results indicate that spectral features from FTIR spectroscopic imaging data

can be used to differentiate between regions of healthly benign prostate epithelial tissue

and regions harboring prostatic adenocarcinoma. The results presented in this section

represent an initial attempt to probe the infrared spectroscopic characteristics of prostate

histopathology and serve to highlight the promise that such vibrational spectroscopic

imaging techniques hold for the objective analysis of sectioned tissue.

Such methods provide simple, readily interpretable image-based results that convey

histological and pathological information provided by referencing spectral database

information. Some of the most useful of these preliminary results are those from section

4.8 from the t-test analyses of individual metrics.

The t-test results displayed in table 4.2 show the 20 metrics for which the patient

populations of patient-mean adenocarcinoma metric values differed significantly from

corresponding benign metric values. Close examination of the spectral parameters of

each of these 20 metrics listed in table 4.1 reveals that most successful metrics involve

spectral information from a handful of spectral regions corresponding to specific

vibrational modes. Several metrics involve spectral information from the spectral region

between 1200-1000 cm-1, a region with prominent absorbances due to vibrational modes

Page 111: Fourier-Transform Infrared Spectroscopic Imaging of

99

of glycogen, as well as symmetric stretching of phosphodiester (PO2-) groups of nucleic

acids. The region between 1300-1200 cm-1 also contributed to several metrics; this

region contains spectral absorbances due to protein Amide III modes and antisymmetric

stretching modes of nucleic acid PO2- groups. Finally, many metrics involved features

from the spectral region between 1590-1500 cm-1, a region whose main absorbance is the

Amide II mode of proteins arising from N-H bending modes coupled to C-N stretching

on the protein backbone.

Future spectroscopic imaging of prostate tissue at higher spectral resolutions will

allow more information to be extracted from these spectral regions. Alternative

classification methods, such as spectral-angle mapping and hierarchical cluster analysis

more readily make use of continuous spectral information, and can be employed using

data from these isolated regions of the spectrum. If performed on substantially larger

patient populations, it is likely that such approaches will lead to more specific

information regarding, spectrally similar subgroups of related cancers and correlations

with histologic grade and/or disease progression

Clearly, studies conducted with larger tissue microarrays and patient populations

will advance our understanding of the spectroscopic properties of prostate pathology.

The technology utilized to collect vibrational spectroscopic imaging data is advancing at

a rapid pace. Faster collection times, better SNRs, and higher data collection at higher

spatial and spectral resolutions will all add to the power of this technique in the future.

Among the most promising future analytical approaches will be to create techniques

to register spectroscopic image data with results from other analytical techniques, such as

immunohistochemical staining and in-situ hybridization conducted after IR data

Page 112: Fourier-Transform Infrared Spectroscopic Imaging of

100

collection, or performed on serial tissue array section. Such combinatorial approaches

should enable calibrations to be constructed that could tentatively predict staining

patterns for multiple panels of antibodies or other probes via spectral pattern recognition

from spectroscopic image data of unstained tissue.

Page 113: Fourier-Transform Infrared Spectroscopic Imaging of

101

References

1. Salzer, R., et al., Infrared and Raman imaging of biological and biomimetic samples.

Fresenius Journal of Analytical Chemistry, 2000. 366(6-7): p. 712-726.

2. Ingle, J.D. and S.R. Crouch, Spectrochemical analysis. 1988, Englewood Cliffs, N.J.:

Prentice Hall. v, 590.

3. Meloan, C.E., Elementary infrared spectroscopy. 1963, New York,: Macmillan. vii,

193 p.

4. Günzler, H. and H.M. Heise, IR spectroscopy : an introduction. 2000, Weinheim:

Wiley-VCH. xiii, 361 p.

5. Struve, W.S., Fundamentals of molecular spectroscopy. 1989, New York: Wiley. xii,

379 p.

6. Levine, I.N., Quantum chemistry. 4th ed. 1991, Englewood Cliffs, N.J.: Prentice Hall.

x, 629 p.

7. Stevens, A. and J.S. Lowe, Histology. 1992, London,New York, Philadelphia: Gower

Medical Pub.(Distributed in the USA and Canada by J.B. Lippincott Co.). 378 p.

8. Stryer, L., Biochemistry. 4th ed. 1995, New York: W.H. Freeman. xxxiv, 1064 p.

9. Lehninger, A.L., D.L. Nelson, and M.M. Cox, Principles of biochemistry. 2nd ed.

1993, New York, NY: Worth Publishers. xli, 1013, [77] p.

10. Solomons, T.W.G., Organic chemistry. 5th ed. 1992, New York: Wiley. 1 v. (various

pagings).

11. Jackson, M. and H.H. Mantsch, Biomedical Infrared Spectroscopy, in Infrared

spectroscopy of biomolecules, D. Chapman, Editor. 1996, Wiley-Liss: New York.

p. 311-340.

Page 114: Fourier-Transform Infrared Spectroscopic Imaging of

102

12. Jackson, M. and H.H. Mantsch, Valinomycin and Its Interaction with Ions in

Organic-Solvents, Detergents, and Lipids Studied by Fourier-Transform Ir

Spectroscopy. Biopolymers, 1991. 31(10): p. 1205-1212.

13. Kubelka, J. and T.A. Keiderling, Differentiation of beta-sheet-forming structures: Ab

initio- based simulations of IR absorption and vibrational CD for model peptide

and protein beta-sheets. Journal of the American Chemical Society, 2001.

123(48): p. 12048-12058.

14. Haris, P.I. and D. Chapman, The Conformational-Analysis of Peptides Using Fourier-

Transform Ir Spectroscopy. Biopolymers, 1995. 37(4): p. 251-263.

15. Silva, R., et al., Discriminating 3(10)- from alpha helices: Vibrational and electronic

CD and IR absorption study of related Aib-containing oligopeptides.

Biopolymers, 2002. 65(4): p. 229-243.

16. Barth, A. and C. Zscherp, What vibrations tell us about proteins. Quarterly Reviews

of Biophysics, 2002. 35(4): p. 369-430.

17. Jackson, M., P.I. Haris, and D. Chapman, Fourier-Transform Infrared Spectroscopic

Studies of Lipids, Polypeptides and Proteins. Journal of Molecular Structure,

1989. 214: p. 329-355.

18. Prestrelski, S.J., D.M. Byler, and M.N. Liebman, Comparison of Various Molecular-

Forms of Bovine Trypsin - Correlation of Infrared-Spectra with X-Ray Crystal-

Structures. Biochemistry, 1991. 30(1): p. 133-143.

19. Surewicz, W.K., H.H. Mantsch, and D. Chapman, Determination of Protein

Secondary Structure by Fourier- Transform Infrared-Spectroscopy - a Critical-

Assessment. Biochemistry, 1993. 32(2): p. 389-394.

20. Torii, H. and M. Tasumi, Theoretical Analyses of the Amide I Infrared Bands of

Globular Proteins, in Infrared spectroscopy of biomolecules, D. Chapman, Editor.

1996, Wiley-Liss: New York. p. 1-18.

21. Byler, D.M. and H. Susi, Examination of the Secondary Structure of Proteins by

Deconvolved Ftir Spectra. Biopolymers, 1986. 25(3): p. 469-487.

Page 115: Fourier-Transform Infrared Spectroscopic Imaging of

103

22. Dong, A., P. Huang, and W.S. Caughey, Protein Secondary Structures in Water from

2nd-Derivative Amide-I Infrared-Spectra. Biochemistry, 1990. 29(13): p. 3303-

3308.

23. Jackson, M., P.I. Haris, and D. Chapman, Fourier-Transform Infrared Spectroscopic

Studies of Ca2+- Binding Proteins. Biochemistry, 1991. 30(40): p. 9681-9686.

24. Trewhella, J., et al., Calmodulin and Troponin-C Structures Studied by Fourier-

Transform Infrared-Spectroscopy - Effects of Ca-2+ and Mg-2+ Binding.

Biochemistry, 1989. 28(3): p. 1294-1301.

25. Lis, H. and N. Sharon, Lectins: Carbohydrate-specific proteins that mediate cellular

recognition. Chemical Reviews, 1998. 98(2): p. 637-674.

26. Schnaar, R.L., et al., Adhesion of Eukaryotic Cells to Immobilized Carbohydrates.

Methods in Enzymology, 1989. 179: p. 542-558.

27. Lewis, R., et al., Physical-Properties of Glycosyldiacylglycerols - an Infrared

Spectroscopic Study of the Gel-Phase Polymorphism of 1,2-Di-O- Acyl-3-O-

(Beta-D-Glucopyranosyl)-Sn-Glycerols. Biochemistry, 1990. 29(38): p. 8933-

8943.

28. Jackson, M., D.S. Johnston, and D. Chapman, Differential Scanning Calorimetric and

Fourier-Transform Infrared Spectroscopic Investigations of Cerebroside

Polymorphism. Biochimica Et Biophysica Acta, 1988. 944(3): p. 497-506.

29. Lee, D.C., I.R. Miller, and D. Chapman, An Infrared Spectroscopic Study of

Metastable and Stable Forms of Hydrated Cerebroside Bilayers. Biochimica Et

Biophysica Acta, 1986. 859(2): p. 266-270.

30. Mueller, E., et al., Oriented 1,2-dimyristoyl-sn-glycero-3-

phosphorylcholine/ganglioside membranes: A Fourier transform infrared

attenuated total reflection spectroscopic study. Band assignments; Orientational,

hydrational, and phase behavior; And effects of Ca2+ binding. Biophysical

Journal, 1996. 71(3): p. 1400-1421.

Page 116: Fourier-Transform Infrared Spectroscopic Imaging of

104

31. Mueller, E. and A. Blume, Ftir Spectroscopic Analysis of the Amide and Acid Bands

of Ganglioside Gm1, in Pure Form and in Mixtures with Dmpc. Biochimica Et

Biophysica Acta, 1993. 1146(1): p. 45-51.

32. Brandenburg, K., S. Kusumoto, and U. Seydel, Conformational studies of synthetic

lipid A analogues and partial structures by infrared spectroscopy. Biochimica Et

Biophysica Acta-Biomembranes, 1997. 1329(1): p. 183-201.

33. Brandenburg, K., Fourier-Transform Infrared-Spectroscopy Characterization of the

Lamellar and Nonlamellar Structures of Free Lipid-a and Re

Lipopolysaccharides from Salmonella-Minnesota and Escherichia- Coli.

Biophysical Journal, 1993. 64(4): p. 1215-1231.

34. Naumann, D., et al., Investigations into the Polymorphism of Lipid-a from

Lipopolysaccharides of Escherichia-Coli and Salmonella- Minnesota by Fourier-

Transform Infrared-Spectroscopy. European Journal of Biochemistry, 1987.

164(1): p. 159-169.

35. Barbucci, R., et al., Physicochemical Surface Characterization of Hyaluronic-Acid

Derivatives as a New Class of Biomaterials. Journal of Biomaterials Science-

Polymer Edition, 1993. 4(3): p. 245-273.

36. Lewis, R.N.A.H. and R.N. McElhaney, Fourier Transform Infrared Spectroscopy in

the Study of Hydrated Lipids and Lipid Bilayer Memebranes, in Infrared

spectroscopy of biomolecules, D. Chapman, Editor. 1996, Wiley-Liss: New York.

p. 159-202.

37. Clark, G. and Biological Stain Commission., Staining procedures. 4th ed. 1981,

Baltimore: Published for the Biological Stain Commission by Williams &

Wilkins. xi, 512 p.

38. Presnell, J.K., M.P. Schreibman, and G.L. Humason, Humason's Animal tissue

techniques. 5th ed. 1997, Baltimore: Johns Hopkins University Press. xix, 572 p.

39. Parker, F.S., Applications of infrared, raman, and resonance raman spectroscopy in

biochemistry. 1983, New York: Plenum Press. xiv, 550 p.

Page 117: Fourier-Transform Infrared Spectroscopic Imaging of

105

40. Liquier, J. and E. Taillandier, Infrared Spectroscopy of Nucleic Acids, in Infrared

spectroscopy of biomolecules, D. Chapman, Editor. 1996, Wiley-Liss: New York.

p. 131-158.

41. Diem, M., S. Boydston-White, and L. Chiriboga, Infrared spectroscopy of cells and

tissues: Shining light onto a novel subject. Applied Spectroscopy, 1999. 53(4): p.

148A-161A.

42. Griffiths, P.R. and J.A. De Haseth, Fourier transform infrared spectrometry.

Chemical analysis ; v. 83. 1986, New York: Wiley. xv, 656 p.

43. Christy, A.A., Y. Ozaki, and V.G. Gregoriou, Modern fourier transform infrared

spectroscopy. Comprehensive analytical chemistry, v. 35. 2001, Amsterdam ;

New York: Elsevier. xx, 356 p.

44. Schaeberle, M.D., I.W. Levin, and E.N. Lewis, Infrared and raman spectroscopy of

biological materials, in Practical spectroscopy, B. Yan, Editor. 2001, M. Dekker:

New York. p. 231-258.

45. Treado, P.J. and M.D. Morris, Infrared and Raman Spectroscopic Imaging, in

Microscopic and spectroscopic imaging of the chemical state, M.D. Morris,

Editor. 1993, M. Dekker: New York. p. 71-108.

46. Perkin-Elmer (Shelton, CT) Spectrum Spotlight 300 FTIR Imaging System

47. Lewis, E.N., et al., Fourier-Transform Spectroscopic Imaging Using an Infrared

Focal-Plane Array Detector. Analytical Chemistry, 1995. 67(19): p. 3377-3381.

48. Colarusso, P., et al., Infrared spectroscopic imaging: From planetary to cellular

systems. Applied Spectroscopy, 1998. 52(3): p. 106A-120A.

49. Bhargava, R. and I.W. Levin, Noninvasive imaging of molecular dynamics in

heterogeneous materials. Macromolecules, 2003. 36(1): p. 92-96.

50. Bhargava, R. and I.W. Levin, Fourier transform infrared imaging: Theory and

practice. Analytical Chemistry, 2001. 73(21): p. 5157-5167.

Page 118: Fourier-Transform Infrared Spectroscopic Imaging of

106

51. Bhargava, R., et al., Novel route to faster Fourier transform infrared spectroscopic

imaging. Applied Spectroscopy, 2001. 55(8): p. 1079-1084.

52. Snively, C.M., et al., Fourier-transform infrared imaging using a rapid-scan

spectrometer. Optics Letters, 1999. 24(24): p. 1841-1843.

53. Huffman, S.W., R. Bhargava, and I.W. Levin, Generalized implementation of rapid-

scan Fourier transform infrared spectroscopic imaging. Applied Spectroscopy,

2002. 56(8): p. 965-969.

54. Lewis, E.N., et al., High-fidelity Fourier transform infrared spectroscopic imaging of

primate brain tissue. Applied Spectroscopy, 1996. 50(2): p. 263-269.

55. Lewis, E.N., et al., Applications of Fourier transform infrared imaging microscopy in

neurotoxicity, in Imaging Brain Structure and Function. 1997, NEW YORK

ACAD SCIENCES: New York. p. 234-247.

56. Lester, D.S., et al., Infrared microspectroscopic imaging of the cerebellum of normal

and cytarabine treated rats. Cellular and Molecular Biology, 1998. 44(1): p. 29-

38.

57. Kidder, L.H., et al., Infrared spectroscopic imaging of the biochemical modifications

induced in the cerebellum of the Niemann-Pick type C mouse. Journal of

Biomedical Optics, 1999. 4(1): p. 7-13.

58. Mendelsohn, R., et al., IR microscopic imaging of pathological states and fracture

healing of bone. Applied Spectroscopy, 2000. 54(8): p. 1183-1191.

59. Marcott, C., et al., Infrared microspectroscopic imaging of biomineralized tissues

using a Mercury-Cadmium-Telluride focal-plane array detector. Phosphorus

Sulfur and Silicon and the Related Elements, 1999. 146: p. 417-420.

60. Mendelsohn, R., E.P. Paschalis, and A.L. Boskey, Infrared spectroscopy, microscopy,

and microscopic imaging of mineralizing tissues: Spectra-structure correlations

from human iliac crest biopsies. Journal of Biomedical Optics, 1999. 4(1): p. 14-

21.

Page 119: Fourier-Transform Infrared Spectroscopic Imaging of

107

61. Campbell, J.B., Introduction to remote sensing. 3rd ed. 2002, New York: Guilford

Press. xxxi, 621 p., [16] p. of plates.

62. Richards, J.A. and X. Jia, Remote sensing digital image analysis : an introduction.

3rd ed. 1999, Berlin ; New York: Springer. xxi, 363 p.

63. Lillesand, T.M. and R.W. Kiefer, Remote sensing and image interpretation. 4th ed.

2000, New York: John Wiley & Sons. xii, 724 p.

64. Cotran, R.S., et al., Robbins pathologic basis of disease. 6th ed. 1999, Philadelphia:

Saunders. xv, 1424 p.

65. McNeal, J.E., Prostate, in Histology for pathologists, S.S. Sternberg, Editor. 1997,

Lippincott-Raven: Philadelphia. p. 997-1017.

66. Jemal, A., et al., Cancer statistics, 2003. Ca-a Cancer Journal for Clinicians, 2003.

53(1): p. 5-26.

67. Boring, C.C., T.S. Squires, and T. Tong, Cancer statistics, 1993. CA Cancer J Clin,

1993. 43(1): p. 7-26.

68. Mettlin, C.J. and G.P. Murphy, Why is the prostate cancer death rate declining in the

United States? Cancer, 1998. 82(2): p. 249-51.

69. Kirby, R.S., M.K. Brawer, and T.J. Christmas, Prostate cancer. 2nd ed. 2001,

London: Mosby. xiii, 230.

70. Franks, L.M., Latent Carcinoma of the Prostate. Journal of Pathology and

Bacteriology, 1954. 68(2): p. 603-&.

71. Breslow, N., et al., Latent Carcinoma of Prostate at Autopsy in 7 Areas -

Collaborative Study Organized by International-Agency-for- Research-on-

Cancer, Lyons, France. International Journal of Cancer, 1977. 20(5): p. 680-688.

72. Sakr, W.A., et al., The Frequency of Carcinoma and Intraepithelial Neoplasia of the

Prostate in Young Male-Patients. Journal of Urology, 1993. 150(2): p. 379-385.

Page 120: Fourier-Transform Infrared Spectroscopic Imaging of

108

73. Sheldon, C.A., R.D. Williams, and E.E. Fraley, Incidental Carcinoma of the Prostate

- a Review of the Literature and Critical Reappraisal of Classification. Journal of

Urology, 1980. 124(5): p. 626-631.

74. Silverberg, E. and J.A. Lubera, Cancer statistics, 1989. CA Cancer J Clin, 1989.

39(1): p. 3-20.

75. Woolf, C.M., An Investigation of the Familial Aspects of Carcinoma of the Prostate.

Cancer, 1960. 13(4): p. 739-744.

76. Gronberg, H., F. Wiklund, and J.E. Damber, Age specific risks of familial prostate

carcinoma: a basis for screening recommendations in high risk populations.

Cancer, 1999. 86(3): p. 477-83.

77. Cannon, L., et al., Genetic epidemiology of prostate cancer in the Utah Mormon

genealogy. Cancer Surv, 1982. 1: p. 47-69.

78. Steinberg, G.D., et al., Family History and the Risk of Prostate-Cancer. Prostate,

1990. 17(4): p. 337-347.

79. Matikainen, M.P., et al., Detection of subclinical cancers by prostate-specific antigen

screening in asymptomatic men from high-risk prostate cancer families. Clinical

Cancer Research, 1999. 5(6): p. 1275-1279.

80. Carter, B.S., et al., Mendelian Inheritance of Familial Prostate-Cancer. Proceedings

of the National Academy of Sciences of the United States of America, 1992.

89(8): p. 3367-3371.

81. Smith, J.R., et al., Major susceptibility locus for prostate cancer on chromosome 1

suggested by a genome-wide search. Science, 1996. 274(5291): p. 1371-1374.

82. Xu, J.F., et al., Evidence for a prostate cancer susceptibility locus on the X

chromosome. Nature Genetics, 1998. 20(2): p. 175-179.

83. Rosen, E.M., S. Fan, and I.D. Goldberg, BRCA1 and prostate cancer. Cancer Invest,

2001. 19(4): p. 396-412.

Page 121: Fourier-Transform Infrared Spectroscopic Imaging of

109

84. Gronberg, H., et al., BRCA2 mutation in a family with hereditary prostate cancer.

Genes Chromosomes Cancer, 2001. 30(3): p. 299-301.

85. Edwards, S.M., et al., Two percent of men with early-onset prostate cancer harbor

germline mutations in the BRCA2 gene. Am J Hum Genet, 2003. 72(1): p. 1-12.

86. Bonn, D., Prostate-cancer screening targets men with BRCA mutations. Lancet

Oncol, 2002. 3(12): p. 714.

87. Pienta, K.J. and P.S. Esper, Risk-Factors for Prostate-Cancer. Annals of Internal

Medicine, 1993. 118(10): p. 793-803.

88. Moul, J.W., et al., Racial differences in tumor volume and prostate specific antigen

among radical prostatectomy patients. Journal of Urology, 1999. 162(2): p. 394-

397.

89. Steele, R., et al., Sexual Factors in Epidemiology of Cancer of Prostate. Journal of

Chronic Diseases, 1971. 24(1): p. 29-&.

90. Kipling, M.D. and Waterhou.Ja, Cadmium and Prostatic Carcinoma. Lancet, 1967.

1(7492): p. 730-&.

91. Rooney, C., et al., Case-control study of prostatic cancer in employees of the United

Kingdom Atomic Energy Authority. Bmj, 1993. 307(6916): p. 1391-7.

92. Rosenberg, L., et al., Vasectomy and the Risk of Prostate-Cancer. American Journal

of Epidemiology, 1990. 132(6): p. 1051-1055.

93. Giovannucci, E., et al., A Retrospective Cohort Study of Vasectomy and Prostate-

Cancer in United-States Men. Jama-Journal of the American Medical

Association, 1993. 269(7): p. 878-882.

94. Giovannucci, E., et al., A Prospective Cohort Study of Vasectomy and Prostate-

Cancer in United-States Men. Jama-Journal of the American Medical

Association, 1993. 269(7): p. 873-877.

Page 122: Fourier-Transform Infrared Spectroscopic Imaging of

110

95. Howards, S.S. and H.B. Peterson, Vasectomy and Prostate-Cancer - Chance, Bias, or

a Causal Relationship. Jama-Journal of the American Medical Association, 1993.

269(7): p. 913-914.

96. Stanford, J.L., et al., Vasectomy and risk of prostate cancer. Cancer Epidemiology

Biomarkers & Prevention, 1999. 8(10): p. 881-886.

97. Scher, H.I., Hyperplastic and Malignant Diseases of the Prostate, in Harrison's

principles of internal medicine, A.S. Fauci, Editor. 1998, McGraw-Hill, Health

Professions Division: New York.

98. Chodak, G.W. and H.W. Schoenberg, Early Detection of Prostate-Cancer by Routine

Screening. Jama-Journal of the American Medical Association, 1984. 252(23): p.

3261-3264.

99. Chodak, G.W., P. Keller, and H.W. Schoenberg, Assessment of Screening for

Prostate-Cancer Using the Digital Rectal Examination. Journal of Urology, 1989.

141(5): p. 1136-1138.

100. Wajsman Z., C.T., Detection and Diagnosis of Prostate Cancer, in Prostatic

cancer, G.P. Murphy, Editor. 1979, PSG Pub. Co.: Littleton, Mass. p. 94-99.

101. Jacobsen, S.J., et al., Screening digital rectal examination and prostate cancer

mortality: A population-based case-control study. Urology, 1998. 52(2): p. 173-

179.

102. Richert-Boe, K.E., et al., Screening digital rectal examination and prostate

cancer mortality: a case-control study. Journal of Medical Screening, 1998. 5(2):

p. 99-103.

103. Friedman, G.D., et al., Case-control study of screening for prostatic cancer by

digital rectal examinations. Lancet, 1991. 337(8756): p. 1526-9.

104. Stenman, U.H., et al., A complex between prostate-specific antigen and alpha 1-

antichymotrypsin is the major form of prostate-specific antigen in serum of

patients with prostatic cancer: assay of the complex improves clinical sensitivity

for cancer. Cancer Res, 1991. 51(1): p. 222-6.

Page 123: Fourier-Transform Infrared Spectroscopic Imaging of

111

105. Christensson, A., et al., Serum Prostate-Specific Antigen Complexed to Alpha-1-

Antichymotrypsin as an Indicator of Prostate-Cancer. Journal of Urology, 1993.

150(1): p. 100-105.

106. Higashihara, E., et al., Significance of serum free prostate specific antigen in the

screening of prostate cancer. Journal of Urology, 1996. 156(6): p. 1964-1968.

107. Luderer, A.A., et al., Measurement of the Proportion of Free to Total Prostate-

Specific Antigen Improves Diagnostic Performance of Prostate- Specific Antigen

in the Diagnostic Gray Zone of Total Prostate- Specific Antigen. Urology, 1995.

46(2): p. 187-194.

108. Zhang, W.M., et al., Characterization and immunological determination of the

complex between prostate-specific antigen and alpha(2)-macroglobulin. Clinical

Chemistry, 1998. 44(12): p. 2471-2479.

109. Brawer, M.K., et al., Complexed prostate specific antigen provides significant

enhancement of specificity compared with total prostate specific antigen for

detecting prostate cancer. Journal of Urology, 2000. 163(5): p. 1476-1480.

110. Benson, M.C., et al., Prostate Specific Antigen Density - a Means of

Distinguishing Benign Prostatic Hypertrophy and Prostate-Cancer. Journal of

Urology, 1992. 147(3): p. 815-816.

111. Babaian, R.J., et al., Comparative analysis of prostate specific antigen and its

indexes in the detection of prostate cancer. Journal of Urology, 1996. 156(2): p.

432-437.

112. Horninger, W., et al., Improvement of specificity in PSA-based screening by using

PSA- transition zone density and percent free PSA in addition to total PSA levels.

Prostate, 1998. 37(3): p. 133-137.

113. Carter, H.B., et al., Longitudinal Evaluation of Prostate-Specific Antigen Levels in

Men with and without Prostate Disease. Jama-Journal of the American Medical

Association, 1992. 267(16): p. 2215-2220.

Page 124: Fourier-Transform Infrared Spectroscopic Imaging of

112

114. Carter, H.B., et al., Prostate-Specific Antigen Variability in Men without Prostate-

Cancer - Effect of Sampling Interval on Prostate-Specific Antigen Velocity.

Urology, 1995. 45(4): p. 591-596.

115. Etzioni, R., R. Cha, and M.E. Cowen, Serial prostate specific antigen screening

for prostate cancer: A computer model evaluates competing strategies. Journal of

Urology, 1999. 162(3): p. 741-748.

116. Ferguson, J.K., et al., Prostate-specific antigen detected prostate cancer:

pathological characteristics of ultrasound visible versus ultrasound invisible

tumors. Eur Urol, 1995. 27(1): p. 8-12.

117. Bree, R.L., The role of color Doppler and staging biopsies in prostate cancer

detection. Urology, 1997. 49(3A): p. 31-34.

118. Yu, K.K. and H. Hricak, Imaging prostate cancer. Radiologic Clinics of North

America, 2000. 38(1): p. 59-+.

119. Harris, R.D., A.R. Schned, and J.A. Heaney, Cancer with Endorectal Mr-Imaging

- Lessons from a Learning- Curve. Radiographics, 1995. 15(4): p. 813-829.

120. Jager, G.J., et al., Dynamic TurboFLASH subtraction technique for contrast-

enhanced MR imaging of the prostate: Correlation with histopathologic results.

Radiology, 1997. 203(3): p. 645-652.

121. Bostwick, D.G. and P.A. Dundore, Biopsy pathology of prostate. 1st ed. Biopsy

pathology series ; 20. 1997, London ; New York: Chapman & Hall. xii, 267 p.

122. Gleason, D.F., Histologic grading and clinical staging of prostatic

adenocarcinoma, in Urologic pathology : the prostate, M.P. Tannenbaum, Editor.

1977, Lea & Febiger: Philadelphia. p. 171-197.

123. Iczkowski, K.A. and D.G. Bostwick, Prostate biopsy interpretation - Current

concepts, 1999. Urologic Clinics of North America, 1999. 26(3): p. 435-+.

124. Montironi, R., Prognostic factors in prostate cancer - Pathologists glean a wealth

of clinical detail from the smallest piece of tissue. British Medical Journal, 2001.

322(7283): p. 378-379.

Page 125: Fourier-Transform Infrared Spectroscopic Imaging of

113

125. McNeal, J.E., et al., Cribriform adenocarcinoma of the prostate. Cancer, 1986.

58(8): p. 1714-9.

126. Nielsen, K., et al., Histological grade, DNA ploidy and mean nuclear volume as

prognostic factors in prostatic cancer. Apmis, 1993. 101(8): p. 614-20.

127. Epstein, J.I., G. Pizov, and P.C. Walsh, Correlation of pathologic findings with

progression after radical retropubic prostatectomy. Cancer, 1993. 71(11): p.

3582-93.

128. Chodak, G.W., et al., Results of Conservative Management of Clinically Localized

Prostate-Cancer. New England Journal of Medicine, 1994. 330(4): p. 242-248.

129. Egawa, S., et al., Long-Term Impact of Conservative Management on Localized

Prostate-Cancer - a 20-Year Experience in Japan. Urology, 1993. 42(5): p. 520-

526.

130. Albertsen, P.C., et al., Competing risk analysis of men aged 55 to 74 years at

diagnosis managed conservatively for clinically localized prostate cancer. Jama,

1998. 280(11): p. 975-80.

131. Gaffney, E.F., S.N. Osullivan, and A. Obrien, A Major Solid Undifferentiated

Carcinoma Pattern Correlates with Tumor Progression in Locally Advanced

Prostatic-Carcinoma. Histopathology, 1992. 21(3): p. 249-255.

132. Blackwell, K.L., et al., Combining Prostate-Specific Antigen with Cancer and

Gland Volume to Predict More Reliably Pathological Stage - the Influence of

Prostate-Specific Antigen Cancer Density. Journal of Urology, 1994. 151(6): p.

1565-1570.

133. Claudio, P.P., et al., Expression of cell-cycle-regulated proteins pRb2/p130, p107,

p27(kip1), p53, mdm-2, and Ki-67 (MIB-1) in prostatic gland adenocarcinoma.

Clinical Cancer Research, 2002. 8(6): p. 1808-1815.

134. Cowen, D., et al., Ki-67 staining is an independent correlate of biochemical

failure in prostate cancer treated with radiotherapy. Clinical Cancer Research,

2002. 8(5): p. 1148-1154.

Page 126: Fourier-Transform Infrared Spectroscopic Imaging of

114

135. Sebo, T.J., et al., Perineural invasion and MIB-1 positivity in addition to gleason

score are significant preoperative predictors of progression after radical

retropubic prostatectomy for prostate cancer. American Journal of Surgical

Pathology, 2002. 26(4): p. 431-439.

136. Bryden, A.A.G., et al., Ki-67 index in metastatic prostate cancer. European

Urology, 2001. 40(6): p. 673-676.

137. Bubendorf, L., et al., Tissue microarray (TMA) technology: miniaturized

pathology archives for high-throughput in situ studies. Journal of Pathology,

2001. 195(1): p. 72-79.

138. Bubendorf, L., et al., Hormone therapy failure in human prostate cancer:

Analysis by complementary DNA and issue microarrays. Journal of the National

Cancer Institute, 1999. 91(20): p. 1758-1764.

139. Kononen, J., et al., Tissue microarrays for high-throughput molecular profiling of

tumor specimens. Nat Med, 1998. 4(7): p. 844-7.

140. Brochure, Perkin Elmer Spectrum Spotlight 300 - FT-IR imaging system. 2001,

Perkin-Elmer Instruments, LLC.

141. Schowengerdt, R.A., Remote sensing, models, and methods for image processing.

2nd ed. 1997, San Diego: Academic Press. xlv, 522 p.

142. Fend, F. and M. Raffeld, Laser capture microdissection in pathology. J Clin

Pathol, 2000. 53(9): p. 666-72.

143. Marks, L.S., et al., Morphometry of the prostate: I. Distribution of tissue

components in hyperplastic glands. Urology, 1994. 44(4): p. 486-92.

144. Svindland, A., L.M. Eri, and K.J. Tveter, Morphometry of benign prostatic

hyperplasia during androgen suppressive therapy. Relationships among epithelial

content, PSA density, and clinical outcome. Scand J Urol Nephrol Suppl, 1996.

179: p. 113-7.

Page 127: Fourier-Transform Infrared Spectroscopic Imaging of

115

About the Author

Daniel Celestino Fernandez received his A.B. in Chemistry in 1997 from Amherst

College in Amherst, MA. During his senior year he undertook an honors research project

where he was exposed to vibrational spectroscopy for the first time and wrote a thesis

entitled “Spectroscopic Characterization of the Iron-Sulfur Cluster of Pyruvate Formate-

Lyase Activating Enzyme.” After college he returned home to Tampa, FL where he

began medical school at the University of South Florida. After finishing his first two

years of study toward an M.D. degree, he accepted a Howard Hughes Medical Institute

Research Scholar Fellowship at the National Institutes of Health in Bethesda, MD. In the

Section of Molecular Biophysics of the Laboratory of Chemical Physics of the National

Institute of Diabetes, Digestive, and Kidney Diseases, he joined a team working on

biological applications of spectroscopic imaging techniques with Dr. Ira W. Levin. After

two years as an HHMI-NIH Research Scholar, with the help of the NIH Graduate

Partnerships Program, he enrolled in the Medical Sciences Ph.D. program in the

Department of Pathology and Laboratory Medicine at the College of Medicine of the

University of South Florida and was able to stay at the NIH to complete two additional

years of doctoral research. He has now transferred to the Mount Sinai School of

Medicine in New York City where he is finishing his last year of study toward the M.D.

degree. After graduation he plans to complete a residency in diagnostic radiology.