OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

189
OPTICAL COHERENCE TOMOGRAPHY OF HUMAN TRANSPLANT KIDNEYS A Dissertation submitted to the Faculty of the Graduate School of Arts and Sciences of Georgetown University in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biochemistry and Molecular Biology By Brandon Bright Konkel, M.S. Washington, DC July 6, 2018

Transcript of OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

Page 1: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

OPTICAL COHERENCE TOMOGRAPHY OF HUMAN TRANSPLANT KIDNEYS

A Dissertation

submitted to the Faculty of the

Graduate School of Arts and Sciences

of Georgetown University

in partial fulfillment of the requirements for the

degree of

Doctor of Philosophy

in Biochemistry and Molecular Biology

By

Brandon Bright Konkel, M.S.

Washington, DC

July 6, 2018

Page 2: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

ii

Copyright 2019 by Brandon Konkel

All Rights Reserved

Page 3: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

iii

OPTICAL COHERENCE TOMOGRAPHY OF HUMAN TRANSPLANT KIDNEYS

Brandon Bright Konkel, M.S.

Thesis Advisor: Moshe Levi, Ph.D.

ABSTRACT

Current measures for assessing the viability of donor kidneys offered for transplant are lacking.

Optical Coherence Tomography (OCT) can image subsurface tissue morphology to supplement

current measures and potentially improve prediction of post-transplant function. OCT imaging

was performed on donor kidneys before and immediately after implantation during 169 human

kidney transplant surgeries. A fully automated image analysis pipeline was developed and

validated against trained manual raters to measure structural parameters of the kidney’s proximal

convoluted tubules (PCTs) visualized in the OCT images. The association of these structural

parameters with post-transplant function was investigated. This study included kidneys from live

and deceased donors. 88 deceased donor kidneys in this study were stored by static cold storage

(SCS) and an additional 15 were preserved by hypothermic machine perfusion (HMP). A subset

of both SCS and HMP deceased donor kidneys were classified as expanded criteria donor (ECD)

kidneys, with elevated risk of poor post-transplant function. Post-transplant function was

characterized as either immediate graft function (IGF) or delayed graft function (DGF). In ECD

kidneys preserved by SCS, increased PCT lumen diameter prior to implantation was found to

predict DGF. Pre-implantation lumen diameter in the SCS-ECD group was an average of 25.5

µm in kidneys that experienced IGF, and 27.9 µm in kidneys that experienced DGF. Following

transplant and reperfusion, increased diameter continued to be predictive of DGF in SCS-ECD

kidneys. Post-reperfusion lumen diameter in the SCS-ECD group was an average of 28.1 µm in

kidneys that experienced IGF, and 32.5 µm in kidneys that experienced DGF. In standard criteria

Page 4: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

iv

donor (SCD) kidneys preserved by SCS, there were no significant differences in measurements

between IGF and DGF recovery groups. In kidneys preserved by HMP, reduced distance

between adjacent lumen following transplant and reperfusion was found to predict DGF. Post-

reperfusion inter-lumen distance in the HMP-SCD group was an average of 45.8 µm in kidneys

that experienced IGF, and 41.4 µm in kidneys that experienced DGF. Results suggest that OCT

measurements of PCTs may be useful for predicting post-transplant function in ECD kidneys and

kidneys stored by HMP, or in guiding biopsies towards pathological sites.

Page 5: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

v

This dissertation is dedicated to Joan and David Konkel for their support and to Stephanie Shuey

(who was told on our first date 2 and a half years ago that I was 3 months from graduating) for

her patience.

BRANDON KONKEL

Page 6: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

vi

TABLE OF CONTENTS

INTRODUCTION .......................................................................................................................... 1

Objectives and Specific Aims ..................................................................................................... 1

Contributions of the Thesis ......................................................................................................... 3

Structure of the Thesis................................................................................................................. 4

CHAPTER 1: CLINICAL BACKGROUND ................................................................................. 5

1.1 Introduction ...................................................................................................................... 5

1.2 The Need for Improved Markers of Viability .................................................................. 5

1.3 Proximal Tubule Morphology and Unstressed Physiology.............................................. 7

1.3.1 Na/K/ATPase in Maintenance of the Proximal Convoluted Tubule Sodium

Gradient .................................................................................................................................. 9

1.3.2 Trans-cellular Reabsorption of NaCl and Water .................................................... 10

1.3.3 Aerobic Metabolism in the Proximal Convoluted Tubules Under Normal

Physiological Conditions ....................................................................................................... 11

1.4 Disease/Insult Manifestation in Proximal Convoluted Tubules ..................................... 13

1.4.1 Warm and Cold Ischemia in Cadaver Donor Transplants ...................................... 13

1.4.2 Induction of Ischemia and Subsequent Shift to Anaerobic Metabolism................. 15

1.4.3 Cellular Edema Following Na/K/ATPase Failure in the Proximal Convoluted

Tubules ................................................................................................................................ 16

1.4.4 Ischemia-Reperfusion Injury .................................................................................. 17

1.4.5 Acute Tubular Injury............................................................................................... 19

1.4.6 Interstitial Fibrosis .................................................................................................. 20

1.4.7 Tubular Atrophy...................................................................................................... 23

1.4.8 Arterial and Arteriolar Narrowing (Arteriolar Hyalinosis/Hyaline

Arteriosclerosis)..................................................................................................................... 24

1.4.9 Glomerular Sclerosis (Glomerulosclerosis) ............................................................ 25

1.5 Visualization of Pathology ............................................................................................. 28

1.5.1 Kidney Biopsies ...................................................................................................... 28

1.5.1.1 Procurement Biopsies .......................................................................................... 30

1.5.1.2 Preimplantation “Zero-Time” Biopsies ............................................................... 30

1.5.1.3 Post-Reperfusion Protocol Biopsies .................................................................... 31

Page 7: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

vii

1.5.1.4 Remuzzi Scoring of Biopsies .............................................................................. 32

1.5.1.5 MAPI Scoring of Biopsies .................................................................................. 33

1.5.1.6 Banff Scoring of Biopsies ................................................................................... 34

1.5.2 Visible and Relevant Features in Optical Coherence Tomography Imaging of the

Human Kidney .................................................................................................................. 35

1.6 Summary ........................................................................................................................ 39

CHAPTER 2: IMAGING AND IMAGE PROCESSING ............................................................ 42

2.1 Introduction .................................................................................................................... 42

2.2 Non-Invasive Imaging Modalities Used in Kidney Transplantation ............................. 42

2.3 Optical Coherence Tomography .................................................................................... 46

2.3.1 Time-Domain Optical Coherence Tomography (TD-OCT) ................................... 50

2.3.2 Fourier-Domain Optical Coherence Tomography (FD-OCT) ................................ 51

2.4 Image Processing............................................................................................................ 53

2.4.1 Low-Pass Filter (Blurring Mask) ............................................................................ 54

2.4.1.1 Mean Filter .......................................................................................................... 55

2.4.1.2 Median Filter ....................................................................................................... 55

2.4.1.3 Gaussian Filter..................................................................................................... 56

2.4.2 High-Pass Filter (Sharpening Mask) ....................................................................... 58

2.4.2.1 Prewitt Filter ........................................................................................................ 59

2.4.2.2 Sobel Filter .......................................................................................................... 61

2.4.2.3 Laplacian Filter ................................................................................................... 63

2.4.3 Contrast Enhancement Techniques ......................................................................... 64

2.4.3.1 Contrast Stretching (Normalization) ................................................................... 65

2.4.3.2 Histogram Equalization ....................................................................................... 66

2.4.3.3 Adaptive Histogram Equalization ....................................................................... 67

2.4.4 Edge Detection ........................................................................................................ 68

2.4.4.1 First Order Derivative Edge Detection................................................................ 69

2.4.4.2 Canny Edge Detection ......................................................................................... 72

2.4.4.3 Second Order Derivative Edge Detection ........................................................... 74

2.4.5 Region Segmentation .............................................................................................. 76

2.4.5.1 Global Thresholding ............................................................................................ 76

Page 8: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

viii

2.4.5.2 Otsu’s Method ..................................................................................................... 77

2.4.5.3 Local Thresholding ............................................................................................. 78

2.4.6 Active Contour (Snakes) ......................................................................................... 79

2.4.7 Graph Cuts .............................................................................................................. 81

2.4.8 Segmentation in Optical Coherence Tomography .................................................. 83

2.5 Summary ........................................................................................................................ 85

CHAPTER 3: IMAGE CAPTURE AND ANALYSIS PIPELINE .............................................. 87

3.1 Introduction .................................................................................................................... 87

3.2 System Specifications .................................................................................................... 87

3.2.1 930 vs. 1325nm ....................................................................................................... 87

3.2.2 2D vs. 3D (OCT-B versus OCT-C cans) ................................................................ 90

3.2.3 Field of View .......................................................................................................... 90

3.2.4 Averaging ................................................................................................................ 91

3.2.5 Scale ........................................................................................................................ 91

3.3 Operating Room Imaging Protocol ................................................................................ 92

3.3.1 Basic Setup.............................................................................................................. 92

3.3.2 Timing of Pre and Post Scans, Multiple Scans ....................................................... 96

3.4 Manual Segmentation ..................................................................................................... 97

3.5 Automatic Segmentation .............................................................................................. 100

3.5.1 Automatic Analysis Pipeline................................................................................. 100

3.5.2 Empty B-Scan Detection....................................................................................... 102

3.5.3 Reflection Detection ............................................................................................. 103

3.5.4 High Adipose Detection ........................................................................................ 104

3.5.5 Segmentation of the Renal Capsule-Kidney Cortex Interface .............................. 105

3.5.6 Segmentation of Quantifiable Kidney Cortex....................................................... 106

3.5.7 Segmentation of Proximal Convoluted Tubule Lumen (Region of Interest Map for

Automatic Selection) ........................................................................................................... 110

3.6 Comparison of Automatic and Manual Segmentation ................................................. 113

3.6.1 Measurement Extraction ....................................................................................... 115

3.6.1.1 Density Measurements ...................................................................................... 115

Page 9: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

ix

3.6.1.2 Diameter Measurements .................................................................................... 118

3.6.1.3 Inter-Lumen Measurements .............................................................................. 119

3.6.1.4 Inter-Centroid Measurements ............................................................................ 120

3.6.2 B-Scan Selection and Measurement Compilation ................................................ 120

3.7 Summary ...................................................................................................................... 121

CHAPTER 4: CLINICAL RESULTS ........................................................................................ 124

4.1 Introduction .................................................................................................................. 124

4.2 Evaluating Donors ........................................................................................................ 124

4.2.1 Live and Deceased Donor Kidney Transplantation .............................................. 124

4.2.2 Static Cold Storage and Hypothermic Machine Perfusion in Kidney

Transplantation .................................................................................................................... 125

4.2.3 Standard and Expanded Criteria Donors in Kidney Transplantation.................... 129

4.2.4 Immediate and Delayed Graft Function in Kidney Transplant Recovery ............ 132

4.3 Patient Demographics .................................................................................................. 133

4.4 Density by Area Results ............................................................................................... 135

4.4.1 Density by Area Results Stratified by Transplant Group (IGF and DGF

Combined) ........................................................................................................................... 135

4.4.2 Density by Area Results Stratified by Recovery Group (IGF vs. DGF) .............. 138

4.4.3 Density Results by Association with Post-Transplant Creatinine Decline ........... 138

4.5 Diameter Results .......................................................................................................... 139

4.5.1 Diameter Results Stratified by Transplant Group (IGF and DGF Combined) ..... 139

4.5.2 Diameter Results Stratified by Recovery Group (IGF vs. DGF) .......................... 142

4.5.3 Diameter Results by Association with Post-Transplant Creatinine Decline ........ 142

4.6 Inter-Centroid Results .................................................................................................. 143

4.6.1 Inter-Centroid Results Stratified by Transplant Group (IGF and DGF

Combined) ........................................................................................................................... 143

4.6.2 Inter-Centroid Results Stratified by Recovery Group (IGF vs. DGF) .................. 144

4.6.3 Inter-Centroid Results by Association with Post-Transplant Creatinine Decline 145

4.7 Inter-Lumen Results ..................................................................................................... 145

4.7.1 Inter-Lumen Results Stratified by Transplant Group (IGF and DGF Combined) 145

4.7.2 Inter-Lumen Results Stratified by Recovery Group (IGF vs. DGF) .................... 147

Page 10: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

x

4.7.3 Inter-Lumen Results by Association with Post-Transplant Creatinine Decline ... 147

4.8 Parsimony of Image Measurements ............................................................................. 148

4.9 Summary ...................................................................................................................... 150

CONCLUSION ........................................................................................................................... 154

BIBLIOGRAPHY ....................................................................................................................... 158

Page 11: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xi

LIST OF FIGURES

Figure 1.1: Illustration of the nephron and the renal corpuscle ...................................................... 8

Figure 1.2: Illustration of the mechanisms of proximal convoluted tubule reabsorption of water

and solutes ....................................................................................................................................... 9

Figure 1.3: En-face and in vivo images captured by tandem scanning confocal microscopy on the

rabbit kidney ................................................................................................................................. 17

Figure 1.4: Histopathology of ATI ............................................................................................... 20

Figure 1.5: Histpathology of fibrosis ............................................................................................ 22

Figure 1.6: Illustration of the physical changes associated with partial EMT of tubular epithelial

cells (TECs) in the context of IF ................................................................................................... 23

Figure 1.7: Histopathology of interstitial fibrosis, tubular atrophy, and global

glomerulosclerosis ........................................................................................................................ 24

Figure 1.8: Histopathology of arteriolar hyalinosis. ..................................................................... 25

Figure 1.9: Histopathology of focal segmental glomerulosclerosis.............................................. 27

Figure 1.10: Illustration of the three forms of biopsy conducted on kidneys prior to and following

transplant ....................................................................................................................................... 29

Figure 1.11: Illustration of procurement, pre-implantation, and post-reperfusion timing in

relation to periods of warm and cold ischemic time ..................................................................... 30

Figure 1.12: Histopathology in MAPI biopsy scoring .................................................................. 34

Figure 1.13: Representative B-scan captured in the operating room of a donor kidney (pre-

implantation). ................................................................................................................................ 35

Figure 1.14: B-scan of a human kidney with capsular depressions captured ex-vivo prior to

transplant ....................................................................................................................................... 36

Figure 1.15: B-scan of a human kidney with superficial cysts ..................................................... 37

Figure 1.16: B-scan of human kidney with superficial glomeruli ................................................ 38

Figure 1.17: B-scan of human kidney with dense vascular features............................................. 39

Page 12: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xii

Figure 2.1: Diagram of optical coherence tomographgy penetration and resolution in relation to

other imaging modalities............................................................................................................... 43

Figure 2.2 Diffusion weighted MRI of human kidneys following transplant............................... 44

Figure 2.3: Side by side comparison of proximal convoluted tubules visualized by different

methods ......................................................................................................................................... 46

Figure 2.4: Constructive and destructive interference .................................................................. 47

Figure 2.5: Optical coherence tomography B-scans of age-related macular degeneration, diabetic

macular edema, and the healthy retina .......................................................................................... 49

Figure 2.6: System design for time and Fourier domain optical coherence tomography

systems .......................................................................................................................................... 53

Figure 2.7: Mean (box) filters of different windows sizes applied to optical coherence

tomography scan of human kidney ............................................................................................... 55

Figure 2.8: Median filters of different windows sizes applied to optical coherence tomography

scan of human kidney ................................................................................................................... 56

Figure 2.9: Gaussian filters of different sigma values applied to optical coherence tomography

scan of human kidney ................................................................................................................... 58

Figure 2.10: Prewitt operator applied to optical coherence tomography B-scan of the human

kidney ............................................................................................................................................ 60

Figure 2.11: Sobel operator applied to optical coherence tomography B-scan of the human

kidney ............................................................................................................................................ 62

Figure 2.12: Laplacian operator applied to optical coherence tomography B-scan of the human

kidney ............................................................................................................................................ 64

Figure 2.13: B-scan of human kidney before and after contrast stretching .................................. 66

Figure 2.14: B-scan of human kidney before and after adaptive histogram equalization ............ 68

Figure 2.15: Derivatives of image intensity across edges............................................................. 69

Figure 2.16: Prewitt and Sobel edge detection on B-scan of human kidney ................................ 72

Figure 2.17: Canny edge detection on B-scan of human kidney .................................................. 74

Figure 2.18: Laplacian edge detection on B-scan of human kidney ............................................. 76

Page 13: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xiii

Figure 2.19: Otsu thresholding applied to optical coherence tomography B-scan of human

kidney ............................................................................................................................................ 78

Figure 2.20: Local adaptive thresholding applied to optical coherence tomography B-scan of

human kidney ................................................................................................................................ 79

Figure 2.21: Active contour model applied to optical coherence tomography B-scan of human

kidney ............................................................................................................................................ 81

Figure 2.22: Graph cut applied to optical coherence tomography B-scan of human kidney ........ 83

Figure 3.1: 930nm and 1325nm B-scans of the human kidney .................................................... 88

Figure 3.2: Manually segmented 930nm and 1325nm B-scans of the human kidney .................. 89

Figure 3.3: Technician in sterile surgical attire operates a probe draped in a sterile sleeve to

image a kidney ex-vivo (flushed with preservation solution and resting in a bowl of ice on the

OR back-table) .............................................................................................................................. 93

Figure 3.4: Cropped portions of B-scans of donor kidneys with varying capsule and cortex

thickness ........................................................................................................................................ 93

Figure 3.5: Inter-rater segmentation overlay. Representative B-scan independently segmented by

2 manual raters .............................................................................................................................. 98

Figure 3.6: 3D and 2D representation of optical coherence tomography imaging of vessels in the

human kidney ................................................................................................................................ 99

Figure 3.7: Automated image analysis pipeline .......................................................................... 102

Figure 3.8: Empty image detection ............................................................................................. 103

Figure 3.9: Reflection detection .................................................................................................. 104

Figure 3.10: Cortex and adipose ................................................................................................. 104

Figure 3.11: Edge detection for renal capsule ............................................................................ 106

Figure 3.12: Heterogeneity of cortex appearance ....................................................................... 108

Figure 3.13: Cortex segmentation ............................................................................................... 109

Figure 3.14: Lumen segmentation .............................................................................................. 110

Figure 3.15: False regions of interest removal............................................................................ 112

Page 14: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xiv

Figure 3.16: Manual versus automatic segmentation overlay .................................................... 114

Figure 3.17: 3D lumen reconstruction with B-scan and orthogonal sectioning ......................... 117

Figure 3.18: Depiction of methodology for inter-lumen and inter-centroid measurements ....... 120

Figure 4.1: Hierarchy classification of transplant groups ........................................................... 135

Figure 4.2: Box and whisker plots of density measurements ..................................................... 137

Figure 4.3: Box and whisker plots of diameter measurements ................................................... 141

Figure 4.4: Box and whisker plots of inter-centroid measurements ........................................... 144

Figure 4.5: Box and whisker plots of inter-lumen measurements .............................................. 146

Page 15: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xv

LIST OF TABLES

Table 3.1: Intra-rater reproducibility and algorithm performance. ............................................. 115

Table 4.1: Measurements selected by lasso penalized regression modeling as the most relevant to

post-transplant function .............................................................................................................. 148

Page 16: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xvi

ABBREVIATIONS

ACD

AH

AKI

apparent diffusion coefficient

arteriolar hyalinosis

acute kidney injury

ATI acute tubular injury

ATN acute tubular necrosis

CT

CIT

computed tomography

cold ischemic time

CKD chronic kidney disease

DCD donation after cardiac death

DDKT

DGF

deceased donor kidney transplant

delayed graft function

DW-MRI

ECD

diffusion weighted magnetic resonance imaging

expanded (extended) criteria donor

ECM extracellular matrix

EMT epithelial-to-mesenchymal transition

ESRD end stage renal disease

FD-OCT

fMRI

FOV

FSGS

fourier domain optical coherence tomography

field of view

functional magnetic resonance imaging

focal segmental glomerular sclerosis

HMP hypothermic machine perfusion

HR

IF

hazard ratio

interstitial fibrosis

IGF immediate graft function

IRI ischemia-reperfusion injury

KAS kidney allocation system

KDPI kidney donor profile index

KDRI kidney donor risk index

LD luminal diameter

Page 17: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

xvii

LDKT live donor kidney transplant

LoG

MAE

MAPI

mPTP

Laplacian of Gaussian

mean absolute error

Maryland aggregate pathology index

mitochondrial transition pore

MRI

NSCC

magnetic resonance imaging

non-selective cation channel

OCT optical coherence tomography

OPTN

ORS

organ procurement and transplantation network

organ recovery system

PCT proximal convoluted tubule

RAM

RMSE

ROI

ROS

random-access memory

region root mean square error

region of interest

reactive oxygen species

SCD standard criteria donor

SCS static cold storage

SD-OCT

SLD

SS-OCT

TA

spectral domain optical coherence tomography

super luminescent diode

swept source optical coherence tomography

tubular atrophy

TD-OCT

TEC

time domain optical coherence tomography

tubular epithelial cell

UNOS

WIT

united network of organ sharing

warm ischemic time

Page 18: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

1

INTRODUCTION

Objectives and Specific Aims

The research entailed in this thesis was undertaken to assess the utility of OCT in visualizing

kidney micro-anatomy during transplant, the ability of automated image analysis to produce

measurements from the resulting imagery, and the relevance of these features to post-transplant

function. The objectives of this thesis were divided into four specific aims. These included

development of the initial protocol for capturing OCT images in the operating room during

transplant, development of a strategy for circumventing limitations and bias introduced during

imaging, development of a fully automated algorithm for segmentation of kidney features, and

finally a thorough investigation of the clinical relevance of the measurements produced in this

study. These aims are discussed here in detail here.

• Develop an operating room protocol for pre-implantation and post-transplant OCT imaging of

human kidneys during transplant.

o Define general methodology (i.e. timing during transplant process, sampling strategy)

for practical imaging of human kidneys during transplant

o Refine OCT settings (i.e. wavelength, scale, field of view, A-scan and B-scan

averaging) to optimize resolution, and minimize file size and speckle noise

• Develop a working strategy for sub-sampling of OCT imaging data to identify high-quality

images which contain the anatomical features under investigation, and to remove potential bias

and redundancy incurred by a flawed imaging protocol.

• Develop a robust system for automated segmentation and measurement of microanatomy of the

superficial kidney cortex (namely the proximal convoluted tubules) in OCT imaging data

o Construct a user-friendly system for high-speed manual segmentation of kidney

features

o Enlist multiple trained raters to segment kidney anatomy in an unbiased and

reproducible manner (to establish ground-truth segmentation labels, and to provide

measurements of inter-rater variation)

o Design a layer segmentation program for segmentation of the interface between the

renal capsule and the surface of the renal cortex

Page 19: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

2

o Design a signal/noise threshold segmentation program for segmentation of high signal

areas of kidney cortex in B-scans, where micro-anatomy should be reliably discernable

from noisy features if present

o Design a region of interest segmentation program for segmentation and identification

of cross-sections of proximal convoluted tubules

o Extract measurements of anatomical features for analysis

o Perform validation of the proposed segmentation system

• Investigate any potential clinical relevance of segmented features, as evidenced by significant

correlation with post-transplant function

o Partition heterogeneous patient pool into sub-populations to reduce the influence of

variables (e.g., storage method) which may impact the interpretation of kidney

structure morphology

o Investigate measurement trends between transplant groups and patient sub-populations

to determine potential effects of storage method or donor features on visible

anatomical features.

o Investigate recovery within each patient population to determine if correlation exists

between the produced measurements and post-transplant function

Page 20: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

3

Contributions of the Thesis

• We provide a comprehensive review of the clinical motivation for evaluating PCT morphology

in the context of kidney transplants, the suitability of OCT for this task, and the requirement

for automated processing of resulting data.

• We introduce core concepts of image processing which have motivated our segmentation

strategies.

• We propose a pipeline for classification, segmentation, and selection of images for analysis.

The proposed framework for analysis sifts through sets of redundant, biased, and

heterogeneous image sets to extract meaningful data which may have clinical relevance.

• We propose a fully automated algorithm for high-speed segmentation and measurement of

kidney microstructures in OCT image sets. The described algorithm performs well, relative to

segmentation by manual raters, despite numerous challenges presented by imaging artifacts,

image noise, heterogeneity in tissue appearance, and heterogeneity in image quality.

• We provide extensive validation of the segmentation algorithm.

• We propose several methods for quantification of PCT features in OCT image sets, including a

proposed correction for the limits imposed by a 2D imaging protocol.

• We explore the clinical relevance of quantified PCT features, address potential redundancy of

measurements, and propose several possible clinical explanations for our findings.

Page 21: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

4

Structure of the Thesis

• Chapter 1 of this Thesis provides clinical background, describing the proximal convoluted

tubules (PCTs) which comprise the bulk of the kidney cortex and are readily visible in OCT

imaging, pathology in the kidney as it pertains to graft viability, and current industry practices

for quantifying and assessing viability.

• Chapter 2 provides a technical background, including a review of optical coherence

tomography (OCT), a guide to fundamental image processing strategies, and a brief

introduction to image segmentation within the context of OCT.

• Chapter 3 introduces a transplant-setting protocol for kidney imaging by OCT, and outlines an

image classification, segmentation, and selection pipeline for analysis of OCT kidney imaging

data. Chapter 3 concludes with validation of the segmentation process by direct comparison

with manual segmentation.

• Chapter 4 introduces a framework for categorizing a heterogeneous patient population, and

investigates the clinical relevance of measurements of automatically segmented kidney

microanatomy in OCT image sets within each category.

• This Thesis concludes with a summary of the accomplishments and discoveries of the Thesis,

and proposes future directions for this line of research.

Page 22: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

5

CHAPTER 1: CLINICAL BACKGROUND

1.1 Introduction

In this chapter, we introduce the need to explore new markers of donor kidney viability.

We go on to review the proximal convoluted tubules (PCTs) in depth, both in a healthy state and

under duress from ischemic insult or pre-existing conditions like fibrosis. Next, we review

current methods for evaluation of kidney biopsies with special attention paid to the evaluation of

pathology in the PCTs. Finally, we explore the potential utility of an optical kidney biopsy

(namely, Optical Coherence Tomography (OCT)), and introduce features indicative of pathology

which can be revealed by optical biopsy. We conclude this chapter with a brief discussion on

how OCT may reveal many of the features directly assessed in traditional biopsies, but in a

global and non-invasive fashion.

1.2 The Need for Improved Markers of Viability

The number of patients in the US waiting for a kidney for transplant has essentially

doubled each decade for the last 30 years. By 2014, the number of persons in the US awaiting a

kidney for transplant had reached nearly 100,000. Around 3,000 more are added to this list each

month, roughly doubling the number of transplants performed monthly [1], [2]. An aging

population contributes to this trend and likewise contributes to the composition of the transplant

list. As the transplant list has grown, the average age, frequency of diabetes and time on dialysis

for the transplant candidates on the list has grown in tandem. In 2016, close to half of the

Page 23: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

6

candidates on the list had been on dialysis for at least 5 years; most will wait at least three and a

half years for a kidney. As they wait their turn, the health of these potential recipients

deteriorates. Nearly 4,000 candidates became too sick for transplant and nearly 5,000 died while

waiting for a kidney in 2014 [1], [3].

While the number of patients requiring a kidney transplant continues to grow, the number

of kidneys available for transplant has remained relatively steady and markedly insufficient. The

supply cannot meet the increasing demand and this disparity grows each year. To combat this,

hospitals have expanded donor criteria to include more ‘marginal’ donors. Requirements for a

kidney to be deemed acceptable for transplant have relaxed such that the donor pool now

includes older donors, donors with pre-existing conditions which may detract from the viability

of the kidney, donor kidneys with suboptimal procurement, and donors with longer periods of

warm and cold ischemia following procurement. The inclusion of these higher risk, expanded

criteria donors (ECD) in the donor pool has been successful in increasing the number of

transplants performed annually but transplant centers still ultimately discard a large portion of

kidneys procured and offered for transplant [4]–[6]. The discard rate for ECD kidneys is nearly

45% compared to just over 10% for standard criteria donor (SCD) kidneys [7].

These discards represent a largely untapped source of potentially viable kidneys which, if

properly utilized, could further widen the donor pool and narrow the gap between kidney supply

and kidney demand. Studies have demonstrated that patients who receive moderately

compromised kidneys live longer and have a higher quality of life than those who remain on

dialysis and wait for a more viable option [8], [9]. Currently there are approximately 17,000 kidney

transplants a year in the United States. It is estimated that this number could be as high as 38,000

Page 24: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

7

if more marginally compromised kidneys were considered and the donor pool properly utilized

[10].

Surgeons reference a multitude of factors which contribute to their decision to reject a

kidney. Principal among these are the results of biopsies, which are performed routinely on ECD

kidneys, and are credited as the most frequent reason for discard. The true relevance of these

factors and of biopsy results specifically is contested, with the majority appearing to have little

correlation with graft function following transplant [11]. There is a critical need to enhance

prognostic measures and to explore new ways of gaining insights into the viability of these more

at-risk kidneys.

1.3 Proximal Tubule Morphology and Unstressed Physiology

Nephrons are the functional unit of the kidney and are found in adult kidneys in numbers

of around 1 million per kidney (Figure 1.1). The two main components of the nephron are the

renal corpuscle and the renal tubule. The renal corpuscle is composed of a capillary tuft called a

glomerulus, surrounded by a capsular space (Bowman’s capsule). Filtration occurs in glomeruli

when filtrate passes from the blood through the fenestrated capillary endothelial cells, basement

membrane, and podocytes which make up the glomerulus. Glomerular filtrate enters Bowman’s

capsule and then exits the renal corpuscle via the renal tubule. The renal tubule is composed of

the proximal tubule (which can be further subdivided into the proximal convoluted tubule and

proximal straight tubule), the ascending and descending loop of Henle, the distal convoluted

tubule, and the collecting ducts. As filtrate leaves the renal corpuscle, it enters the convoluted

section of the proximal tubule first. Here, most of the water and salt are reabsorbed, entering the

Page 25: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

8

surrounding interstitial space and diffusing into the peritubular capillaries which run parallel to

the renal tubule. Next, filtrate enters the straight section of the proximal tubule, where some

Phosphate absorption occurs. As filtrate moves on to the loop of Henle, concentration or dilation

of filtrate occurs in addition to further sodium, potassium, calcium, and magnesium reabsorption

by the ascending limb of the loop. Finally, filtrate enters the distal convoluted tubule and

collecting duct where the last bit of sodium, calcium, and magnesium are reabsorbed [12].

Figure 1.1: Illustration of the nephron and the renal corpuscle. The

nephron (right) and renal corpuscle (left) are the structural and functional unit

of the kidney, and the point of filtration respectively.

The primary role of the PCTs is in reabsorption of water and solutes from the glomerular

filtrate. Around 1200 ml of blood flow through the kidneys every minute, the PCTs are

responsible for reabsorbing approximately two thirds of the glomerular filtrate. The walls of the

PCTs are composed of simple cuboidal epithelial cells interconnected by tight junctions near

their apical surface. The apical surface of these cells contains a microvillus brush border,

Page 26: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

9

designed to maximize surface area so as to optimize reabsorption from the glomerular filtrate.

Interspersed among the brush border membrane are a series of secondary-active sodium-

dependent counter-transporters and cotransporters which, driven by a low intracellular sodium

concentration relative to the luminal filtrate, facilitate the absorption of glucose, amino acids and

other organic solutes into the epithelial cells via their coupling to sodium and diffusion down the

sodium gradient. The reabsorbed solutes are actively expelled from the basolateral surface of the

tubular epithelium by sodium-independent facilitated diffusion where they enter the interstitium

and can move passively into the peritubular capillaries and reenter systemic circulation [13].

Figure 1.2: Illustration of the mechanisms of proximal convoluted tubule

reabsorption of water and solutes.

1.3.1 Na/K/ATPase in Maintenance of the Proximal Convoluted Tubule Sodium Gradient

The low intracellular sodium content (about 1/10th the extracellular environment) and

consequently the gradient which drives the reabsorption process is sustained by active extrusion

of sodium from the basolateral surface of the PCT epithelium into the interstitial fluid [14]. This

feat is accomplished by Na/K/ATPase which present on the basolateral membrane of the PCT

Page 27: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

10

endothelium. The Na/K/ATPase exchange 3 intracellular sodium ions for 2 extracellular

potassium ions. The positively charged sodium ions are drawn to the electronegative intracellular

environment and similarly tend to diffuse down their concentration towards the comparatively

low intracellular concentration of sodium. The extrusion process therefore occurs against both an

electrical and chemical gradient [15].

The PCT cell membrane is permeable to sodium and potassium, allowing these ions to

readily diffuse down their electrochemical gradient into and out of the cell respectively.

Maintenance of the low intracellular sodium concentration and high intracellular potassium

concentration therefore must occur not just against the electrochemical gradient but must do so at

a rate that exceeds the passive diffusion of sodium back into the cell. This primary-active process

which in turn drives the secondary-active sodium-dependent apical absorption from the

glomerular filtrate is the largest consumer of energy in the PCTs and as such-the majority of the

cell’s mitochondria aggregate along the basolateral surface to provide ATP for catalyzed

hydrolysis to power this extrusion process. This process is dependent upon a healthy supply of

metabolites and fresh supply of oxygen [16].

1.3.2 Trans-cellular Reabsorption of NaCl and Water

Chloride is primarily reabsorbed into active circulation through leaky tight junctions via

the paracellular route. Chloride can move through the PCT cell as well, via apical Na/Cl and

Na/K/2Cl transporters, piggybacking on sodium’s diffusion down its’ electrochemical gradient.

Chloride also enters the cell through a chloride base-exchanger, driven by the pH gradient

established as hydrogen ions are excreted apically in exchange for Na ions by the NHE3

Page 28: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

11

exchanger [17], [18]. In essence, chloride follows sodium into the cell either by direct coupling

or indirectly through gradients established by intake of sodium. Influx of NaCl into the cell is

followed immediately by an influx of water, drawn osmotically through apical aquaporins.

Chloride’s basolateral export from the PCT is similarly occurs through exchangers and

transporters driven by the movement of sodium. Again, basolateral aquaporins permit water to

follow the ionic shift unencumbered. In this fashion, quantities of water up to four times the

volume of the PCT cell pass through it every minute; combined with the passive diffusion of

sodium, chloride and water across leaky tight junctions, this mechanism promotes the

reabsorption of 60-70% of the NaCl and water from the filtrate. Over 25,000 mmoles of sodium

pass through the lumen of the PCTs each day, and of this only 0.4% makes it to urine, meaning

the PCTs are responsible for absorption of roughly 2/3rds of 99.6% of the sodium load or 15,000

to 17,500 mmoles of sodium a day [16].

1.3.3 Aerobic Metabolism in the Proximal Convoluted Tubules Under Normal Physiological

Conditions

Respiration in the PCTs is predominantly aerobic. Metabolites including glucose, lactate,

Krebs cycle intermediates and amino acids arrive via the filtrate and enter the cells of the PCT

via transporters in the microvillus brush border on the apical surface. Absorbed metabolic

substrates enter the Krebs cycle directly through a specific route or are interconverted to another

substrate which then enters the Krebs cycle. Some evidence suggests transporters for some of

these, and subsequently the potential for their absorption, are present to a lesser to degree on the

basolateral surface of the PCT epithelial cells-suggesting they may absorb some portion of these

Page 29: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

12

metabolites from the peritubular capillaries. Similarly, fatty acids, one of the main suppliers of

energy to the PCTs, are primarily bound to albumin which does not pass into the filtrate.

Consequently, fatty acids must be drawn into the PCTs from the peritubular capillaries through

the basolateral surface [16].

The PCTs succeed in recapturing 99.8% of the metabolites that pass through their lumen,

and utilize only a minority of this to power Na/K/ATPase in the maintenance of the sodium

electrochemical gradient which drives reabsorption. The cells of the PCT are adaptable in that

they can use any of a variety of metabolic substrates provided in the filtrate or drawn from the

peritubular capillaries. While they exhibit some preference under physiologically normal

conditions, their choice of metabolic substrate is also influenced by their metabolic state,

temperature, pH, the mix of metabolites of which they are provided in the filtrate or peritubular

capillaries.

While the aerobic PCT metabolism is flexible and can utilize whichever metabolic

substrates are most readily available; there are preferences when multiple metabolites are present

simultaneously. Fatty acids are the preferred metabolic substrate for the aerobic PCT. In the

presence of other substrates, fatty acids are selectively oxidated while the other metabolites are

reabsorbed into the bloodstream or utilized as gluconeogenic substrates (the PCTs are an

important site for gluconeogenesis). Oxidation of lactate and ketones provides another

substantial source of energy for the PCT-these substrates outcompete amino acids as targets for

metabolism. Amino acids that enter the glomerular filtrate are almost entirely reabsorbed by the

PCTs; the kidney is able to metabolize glutamine, alanine, glycine, serine and arginine, although

glutamine is the preferred substrate among these. Krebs cycle intermediates are readily utilized

Page 30: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

13

for respiration, either by direct entry into the Krebs cycle or entry following interconversion to

another intermediate; chief among these is citrate which under normal conditions provides 10-

15% of the PCTs required energy [19].

Filtered glucose is almost completely reabsorbed in the PCTs but plays very little role in

the aerobic PCT metabolism. While the PCT epithelial cells contain the necessary enzymes to

metabolize glucose, glucose is outcompeted by most other substrates when offered together [20].

Finally, endogenous lipids and glycogen are also readily metabolized by the cells of the PCT.

Glycogen stores in the PCT however are limited and so are quickly exhausted. Even prior to their

exhaustion, endogenous substrates alone can only provide the PCT with a portion of its required

energy.

1.4 Disease/Insult Manifestation in Proximal Convoluted Tubules

1.4.1 Warm and Cold Ischemia in Cadaver Donor Transplants

Kidneys destined for transplant are often procured from non-heart-beating donors

(donation after cardiac death (DCD)). In non-heart-beating donors, there is often an extended

period of warm ischemia prior to any intervention, likely followed by a prolonged period of cold

ischemia as the kidney is matched and transported to the recipient [21]. Multiple studies have

suggested that progressively longer periods of warm ischemia time (WIT) correspond with an

increasing risk of poor graft function and graft failure [6], [22]. This correlation is echoed by

discard rates which increase in parallel with increasing WIT [23].

Following procurement of the kidney, the graft is generally perfused with preservative

and transitioned to cold storage. Preservation solutions first came to use in the late 1960’s and

Page 31: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

14

have become indispensable in kidney preservation. University of Wisconsin solution [24], Euro-

Collins and HTK are the three solutions most often in use today. These contain varying amounts

of impermeant osmotic agents, pH and ion buffers, free radical scavengers, calcium antagonists,

colloids, complement regulators and antiplatelet agents. Chief among these in terms of

contribution to extending the viability of the kidney for transplant is the impermeant osmotic

agent [25], [26].

As their name suggests, impermeant osmotic agents cannot permeate the cell membrane

and so remain extracellular where they provide an osmotic force. The osmotic properties of the

extracellular impermeant agents balance the contributions intracellularly of the influx of sodium

caused by hypothermic or ATP-starvation-induced disabling of the Na/K/ATPase [27], [28]. This

prevents water from osmotically entering the cell or draws water out of the cell if swelling has

already occurred [29]. If water is exiting the cell, the rising intracellular potassium concentration

relative to decreasing cell volume will promote potassium’s passive diffusion out of the cell,

further contributing to the extracellular colligative properties osmotic force. Similarly, water

exiting the cell can physically carry potassium with it and achieve the same outcome [30].

Cold ischemic time (CIT) is unavoidable in kidney transplants, and can range from under

an hour (in some live donor transplants) to more than 30 hours. Each hour of cold storage further

reduces the viability of the graft; a kidney stored for 30 hours has a 40% higher change of graft

failure than a kidney stored for 6 hours [31]. Prolonged cold ischemia can also have a synergistic

effect when coupled with periods of WIT, compromising viability of the graft further [23], [32].

Page 32: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

15

1.4.2 Induction of Ischemia and Subsequent Shift to Anaerobic Metabolism

The high energy expenditure/requirements of the PCTs make them especially susceptible

to ischemic conditions [33]–[35] . If blood flow to the kidney is restricted, supply of oxygen to

the kidney’s cells is abruptly cutoff; delivery of metabolites ceases and metabolic waste products

from remaining sources begin to accumulate and act to further inhibit energy production. Under

normal physiological conditions, oxygen levels are highest in the kidney’s cortex where

epithelial cells employ an oxygen dependent metabolism. Following induction of ischemia, the

cortex’s relatively large oxygen reserve will sustain aerobic metabolism for a very brief period.

When oxygen reserves are depleted, the cells revert to an anaerobic metabolic state [36], [37].

In the anaerobic state, the aerobically insignificant glucose becomes the primary source

of energy for the PCT cells. The anaerobic metabolism of glucose generates energy by substrate

level phosphorylation: alpha-ketoglutarate dehydrogenase’s conversion of alpha-ketoglutarate to

succinyl-CoA. Similarly, fumarate reductase’s conversion of fumarate to succinate coupled to

the oxidation of quinol to quinone drives electron transport in complexes 1 and 2. These

pathways are markedly less efficient than aerobic glycolysis, producing a fraction of the ATP

and producing lactate as a harmful byproduct. The modest ATP production is insufficient to

support PCT homeostasis, and is further reduced as metabolic substrates, un-replenished by an

active circulatory system, are depleted [35], [38]. In addition to the dramatic reduction in ATP

generation, the hypoxic environment will lead to an increase in ATP consumption. Inhibition of

the electron transfer chain will compromise mitochondrial membrane potential. ATP synthase

will reverse direction in an effort to preserve the membrane potential and will being hydrolyzing

ATP instead of synthesizing it [39].

Page 33: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

16

1.4.3 Cellular Edema Following Na/K/ATPase Failure in the Proximal Convoluted Tubules

As osmotic pressure is a colligative property, the number of solutes within the cell

dictates the volume of water within the cell. Potassium ions, which are generally at very high

concentrations intracellularly, are the principal osmotic solute within cells. Macromolecules and

impermeable metabolites which accumulate within the cell contribute little in number to the

osmolality of the cell but their high charge draws a high number of counter-ions which make

more substantial contributions to the cell’s osmolarity. The relative high concentration of sodium

outside the cell, formed by action of the Na/K/ATPase pump which disproportionally excludes 3

cations for every 2 it accumulates, serves to balance the colligative contributions of the

impermeable macromolecules and metabolites and their accompanying counter-ions within the

cell [15], [40].

The high levels of energy required to power Na/K/ATPase and maintain the absorption

process are what make the PCTs so sensitive to ischemic insult [41]. Following induction of

ischemia, the rapid drop in available ATP produces, accordingly, a parallel drop in Na/K/ATPase

performance. Sodium continues to enter the cell apically, diffusing down its electrochemical

gradient, trailed by chloride, through various transporters or the sodium-permeable membrane.

The paralyzed Na/K/ATPase no longer extrudes sodium at a rate matching its entry and so

sodium and chloride accumulate within the cell. With the colligative contributions of the

impermeant intracellular molecules no longer balanced by the active extrusion of sodium, water

osmotically pours in apically and basolaterally through the aquaporins and cellular edema ensues

[27]. Under normothermic conditions (warm ischemia), ischemia leads to edema of the PCTs

Page 34: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

17

within seconds (~30s) [42]. Under hypothermic conditions and perfused with preservation

solution, swelling is a much more gradual process.

Figure 1.3: En-face and in vivo images captured by tandem scanning

confocal microscopy on the rabbit kidney. (a) The PCTs of the kidney

cortex prior to ischemic insult. The hyper-reflective lining of the tubular

lumen is the microvillus brush border. (b) The PCTs following roughly 40

seconds of normothermic ischemia. Swelling of the cuboidal epithelium has

fully occluded the luminal space. (Snapshots acquired from video provided by

Peter Andrews, PhD).

1.4.4 Ischemia-Reperfusion Injury

Prolonged periods of cold ischemia, and to a greater extent, combined warm and cold

ischemic insult (DCD donors) to the kidney produce a host of deleterious responses which are

further activated upon reperfusion of the transplanted kidney. This effect is referred to as

ischemia-reperfusion injury (IRI). The reintroduction of oxygen into the ischemic kidney

produces a boom in reactive oxygen species (ROS), an inflammatory response, an increase in

intracellular calcium concentrations, mitochondrial dysfunction and various other effects which

contribute to apoptosis and the general dysfunction of the kidney. Upon reperfusion, sub-lethally

damaged cells which may be displaying necrotic symptoms due to prolonged ischemia become

apoptotic [43]–[47].

Page 35: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

18

The most dangerous actors during reperfusion are the ROS. While ROS are generated

throughout normothermic and hypothermic ischemia, they are produced in far greater quantities

upon reperfusion and reintroduction of oxygen to the ischemically damaged system. Nitric oxide

synthase and xanthine oxidase are activated under ischemic conditions. Xanthine oxidase acts as

the principal contributor to ROS generation upon reperfusion. Following reperfusion, oxygen

reacts with nitric oxide produced by the ischemically activated nitric oxide synthase and xanthine

oxidase, forming peroxynitrite [43], [48]. Peroxynitrite acts as a highly volatile oxidizing agent,

promoting oxidative stress and further damaging the cell [49].

Under ischemic conditions, cytosolic calcium accumulates at a relatively modest rate.

Free cytosolic calcium doubles within 60 minutes of ischemia. The degree to which calcium is

allowed to accumulate intracellularly is thought to parallel the cell’s movement towards an

irreversible extent of damage [50]. As calcium concentrations rise in the cytosol, the cell’s

mitochondria begin to take up some of this excess. Calcium influx into the mitochondria would

typically lead to formation of a mitochondrial transition pore (mPTP), however the acidic

intracellular environment produced by the byproducts of an anaerobic metabolism serves to

inhibit this process. As pH levels stabilize upon reperfusion, inhibition of mPTP formation

dissipates and the high intra-mitochondrial calcium concentrations induce mPTP opening,

leading to apoptosis [51].

The inflammatory response following reperfusion contributes to the deleterious effects of

IRI. The PCTs of the kidney are stocked with a host of inflammatory mediators, intended for

deployment into the urinary tract if a urinary infection presents. Once triggered, the PCTs

generate and release cytokines and chemokines [46]. Disruptions in the cell surface may pose a

Page 36: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

19

target to these mediators of inflammation. Cellular components released during ischemic injury

or upon reperfusion may similarly pose a target for inflammation. In addition to producing

targets for an inflammatory response, the hypoxic state may also impair anti-inflammatory

mechanisms, exacerbating the inflammatory effect by dampening the cell’s response [47], [52].

1.4.5 Acute Tubular Injury

Prolonged periods of warm and cold ischemia are believed to be the main factors

contributing to acute tubular injury (ATI) in the context of renal transplant [53]. The degree of

ATI may be evident prior to transplantation and is considered a valuable tool in assessment and

allocation of the kidney. During transplant, ATI leaves the kidney more susceptible to IRI.

Similarly, the effects of IRI may exacerbate the degree of ATI in the transplanted kidney and

further contribute to acute injury [54], [55]. Consequently, ATI is often identified in kidney

grafts following transplant, and is believed to be an early predictor of poor post-transplant

function [56], [57]

Ischemic ATI manifests morphologically as cell swelling (Figure 1.4a), shedding of the

microvillus brush border, and sloughing of viable and necrotic tubular epithelial cells (TECs)

into the tubular lumen (Figure 1.4b). Under ischemic conditions, cell polarity is disrupted, low

ATP leads to a disruption of tight junction and adherent junction integrity, integrins are

redistributed and cell-cell adhesion as well as cellular adhesion to the extracellular membrane is

compromised. Epithelial cells detach and slough into the PCT where they aggregate, forming

casts, and obstruct filtration (Figure 1.4c) [58], [59]. Similarly, Na/K/ATPase dissociates from its

actin cytoskeletal anchors following ischemic insult; dissolution of membrane polarity facilitates

Page 37: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

20

the free moving Na/K/ATPase’s redistribution to the apical surface of the PCT epithelium. This

in turn disrupts the PCT’s ability to effectively reabsorb sodium from the filtrate following

reperfusion [60], [61].

Figure 1.4: Histopathology of ATI. (a) 400x HPS stain showing cell

swelling and apical blebbing (blue arrow) (b) 400x HPS stain showing

epithelial sloughing (black arrow). (c) acid-Schiff stain showing tubular casts

(yellow arrow) (d) 400x HPS stain showing Tubular flattened epithelial cells

(red arrow). [62]

1.4.6 Interstitial Fibrosis

Interstitial fibrosis (IF), or the accumulation of collagen and other molecules which

compose the extracellular matrix (ECM) in the renal interstitium, can serve to provide structural

integrity for surrounding tubules. This can stabilize tubule morphology around locations of

injury, and in doing so help maintain functional structure [63]. This can be advantageous in

Page 38: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

21

repairing short-term insults, but with chronic activation, this process can gradually replace

functional renal parenchyma with scar tissue and impact kidney function. IF can accumulate at

the site of severe focal injuries, or more frequently is diffusely represented surrounding regions

of glomerular, tubular, and vascular disease [64]. IF correlates negatively with kidney function

and similarly correlates negatively with graft function and graft survival if a fibrotic kidney is

transplanted [65], [66].

Page 39: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

22

Figure 1.5: Histpathology of fibrosis. (a,b) Trichrome stain with

corresponding fibrosis quantification markup. (c,d) Collagen III stain with

corresponding fibrosis quantification markup. (e,f) Sirius red stain with

corresponding fibrosis quantification markup. [67]

Myofibroblasts, tubular epithelial cells, endothelial cells, and immune cells are all

believed to play a role in the pathogenesis of IF. In IF, TECs are believed to undergo at least

partial epithelial-to-mesenchymal transition (EMT). Consistent with the early stages of EMT,

TECs lose their epithelial markers and acquire mesenchymal markers [63], [68]. It remains

Page 40: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

23

unclear, however, whether the acquired motility and migratory behavior associated with the later

stages of EMT presents in TECs. In IF, TECs that have undergone partial EMT differentiate into

fibroblasts and contribute to the deposition of collagen and other molecules into the interstitium

[69], [70].

Figure 1.6: Illustration of the physical changes associated with partial

EMT of tubular epithelial cells (TECs) in the context of IF. [69]

1.4.7 Tubular Atrophy

Tubular atrophy (TA) generally occurs in conjunction with IF. TA can refer either to the

loss of individual TECs or entire tubules. In the early stages of TA, TECs lose their brush border

and apical mitochondria. The TECs then undergo tubular simplification (flattening of epithelium

accompanied by dilation of tubular lumen) [71]. Finally, the basement membrane supporting the

TECs experiences wrinkling, inflammatory cells and macrophages invade the cell, and the cell

ultimately scars over [67], [72]. Following TA, cells experience a loss of metabolic activity and

transport functions. TA is generally accompanied by hypertrophy of remaining nephrons to

accommodate the resulting increase in workload [73].

Page 41: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

24

In addition to its association with chronic kidney disease (CKD), TA has been associated

with poor graft function in kidneys following transplant [74]. Bunnag et al. found TA (together

with IF) to be the most predictive, out of a dozen features investigated, of poor graft function

[75]. Seron et al. similarly found TA to correlate negatively with graft survival [76]. Similarly,

Cravedi et al. demonstrated a strong association between TA and several specific graft-

threatening reactions, including acute cellular rejection, antibody mediated injury, and chronic

rejection [77]. With respect to long term graft survival, TA and IF are so commonplace and such

definitive characteristics of graft failure that in recent years “IFTA” has replaced the terminology

“chronic allograft nephropathy”.

Figure 1.7: Histopathology of interstitial fibrosis, tubular atrophy, and

global glomerulosclerosis. (a) 200x PAS stain showing

glomerulosclerosis. (b) 400x PAS stain showing atrophic tubules.

1.4.8 Arterial and Arteriolar Narrowing (Arteriolar Hyalinosis/Hyaline Arteriosclerosis)

Vasculopathy has long been considered predictive of poor post-transplant function and

poor graft survival rates. In a study of 280 patients, Seron et al. reported 95% survival in kidneys

where biopsy histology revealed no pathology, 82% graft survival when biopsies showed IFTA

with no vasculopathy, and 41% graft survival when IFTA was identified in tandem with

vasculopathy [67]. Arteriolar Hyalinosis (AH) is defined by the deposition of hyaline into the

Page 42: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

25

vascular endothelium or sub-endothelial space, leading to a thickening of the vascular walls and

reduction in luminal space. The reduction in luminal space limits blood flow to the functional

units of the kidney, and in doing so compromises kidney function. AH is found more with

increasing age, and is routinely found in conjunction with other kidney pathology (hypertension,

diabetes, glomerular disease) [78], [79]. While some studies contest the link between AH and

graft function, most assert that it correlates strongly with delayed graft function (DGF) and low

graft survival rates [80]–[83]. In addition, AH may manifest before symptoms of IF, TA, and

glomerulosclerosis, indicating AH as an early predictor of graft viability and valuable feature in

biopsy evaluation.

Figure 1.8: Histopathology of arteriolar hyalinosis. 400x PAS stain of

grade 0 (a), 1 (b), 2 (c), and grade 3 (d) hyalinosis. [84]

1.4.9 Glomerular Sclerosis (Glomerulosclerosis)

Glomerulosclerosis is generally associated with the deposition of matrix into the

glomerular capillary lumen, leading to a loss of function. Non-specific global glomerulosclerosis

(sclerosis of the entire glomerular tuft in non-specific glomeruli) is a common occurrence that

develops with aging. While glomerulosclerosis is considered a hallmark of renal aging, its effects

vary widely between individuals. Kaplan et al. reported a range in the percentage of sclerotic

glomeruli of between 0.2 and 16.7% in kidneys from individuals who were at least 55 years old,

Page 43: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

26

and a range of between 1.5 and 23% in kidneys from individuals who were at least 75 [85].

Glomerulosclerosis does not affect whole kidney function to the point of progressing to CKD or

compromising graft viability until the extent of glomerulosclerosis exceeds what would be

expected for the age of the kidney [86].

In 1995, Gaber et al. released their landmark study on the link between

glomerulosclerosis and graft function and survival following transplant. The study reported DGF

in 22% of patients whose biopsy demonstrated no evidence of glomerulosclerosis, DGF in 33%

of patients with less than 20% of biopsied glomeruli sclerosed, and DGF in 87% of patients with

greater than 20% of biopsied glomeruli sclerosed. Incidence of graft loss echoed this trend, with

7% graft loss in patients with less than 20% of biopsied glomeruli sclerosed, and 38% graft loss

in patients with more than 20% of biopsied glomeruli sclerosed [87]. While a handful of studies

contest the correlation between glomerulosclerosis and graft failure, glomerulosclerosis has

remained the most studied feature of the donor biopsy and the most heavily weighted feature in

many biopsy scoring methods [88]. Similarly, discard rates increase in parallel with increasing

degrees of glomerulosclerosis, reflecting surgeons’ confidence in the link between

glomerulosclerosis and graft viability [89].

Focal segmental glomerulosclerosis (FSGS) affects specific glomeruli and only a portion

of the glomerular tuft. FSGS stems from injury or loss of podocytes due to a litany of causes

(genetic mutations, drug abuse, various infections, etc.) [90]. Morphologically, FSGS may

present in several different forms, even within single biopsies, with each variant suggesting a

different prognosis. D’Agati et al proposed a system of classification for these variants in 2004.

This system classified FSGS as either Collapsing FSGS, Cellular FSGS, Tip Lesion, Perihilar

Page 44: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

27

Variant or FSGS not otherwise specified. Instances of Collapsing FSGS were associated with the

worst prognosis [79], [91], [92].

Figure 1.9: Histopathology of focal segmental glomerulosclerosis. (a)

Not otherwise specified type with obliteration of segmental areas of the

glomerular capillary tuft by increased matrix. (b) Collapsing type, with

proliferation of visceral epithelial cells and collapse of the tuft. (c) Tip

lesion with adhesion and/or sclerosis at the proximal tubular pole (right).

(d) Cellular, with increased endocapillary cells. (e) Hilar, with sclerosis

with or without hyalinosis at the vascular pole. Stains: part a, periodic acid

Schiff; parts b–e, Jones’ silver. [79]

Page 45: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

28

1.5 Visualization of Pathology

1.5.1 Kidney Biopsies

The focus on biopsies as a tool for assessing organ quality has drawn more focus as the

donor pool has expanded to include more compromised kidneys. Increased risk of graft failure in

ECD kidneys necessitates a more comprehensive evaluation of kidneys offered for transplant.

Biopsies offer insight into pre-existing pathologies which may exist in a kidney but not

necessarily present in non-invasive assessment of the donor profile. In 1995, Gaber et al.

demonstrated an association between the degree of glomerulosclerosis in biopsies and the

survival of the graft. Gaber reported that sclerosis in 20% or more of biopsied glomeruli

correlated with a reduction in graft life [87]. Biopsies began to gain prominence following this

and are now conducted on nearly 50% of kidneys procured from cadaver donors. The kidney

allocation system (KAS) implemented in 2014 by the Organ Procurement and Transplantation

Network (OPTN) recommended procurement biopsies for all kidneys classified as ECD or at the

request of the receiving transplant surgeon [93]. In kidneys where histologic abnormalities are

expected or which qualify as ECD, biopsies are conducted on roughly 85% of kidneys offered

for transplant [94]. The scoring of biopsies is a continually evolving process, complicated by

heterogeneity in the kind of biopsy (wedge vs. core), and preparation method (frozen vs.

paraffin-embedded). Assessing scoring strategies is similarly complicated by variance and poor

reproducibility between pathologists in applying each strategy [95].

Page 46: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

29

Figure 1.10: Illustration of the three forms of biopsy conducted on

kidneys prior to and following transplant. Advantages and

disadvantages of each method are listed. Dotted lines indicate approximate

position and depth of biopsies, highlighting differences in procured

anatomy from each method. [96]

Transplantation biopsies can be divided in 3 categories: procurement biopsies,

preimplantation biopsies and protocol biopsies (Figure 1.11). Procurement biopsies (also referred

to as “harvest biopsies”) occur immediately following extraction of the kidney from the donor.

This follows any time period associated with extraction of the kidney, and for DCD kidneys also

follows a variable period of warm ischemia. Preimplantation biopsies (also referred to as “zero-

time biopsies”) occur just prior to transplant and following a variable period of cold storage and

cold ischemia. Protocol biopsies occur following reperfusion of the transplanted graft. These can

occur shortly following reperfusion (post-reperfusion biopsy) or months to years later.

Page 47: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

30

Figure 1.11: Illustration of procurement, pre-implantation, and post-

reperfusion timing in relation to periods of warm and cold ischemic

time. [96]

1.5.1.1 Procurement Biopsies

Procurement biopsies are the most frequently performed transplant-associated kidney

biopsies. The inherent advantage of the procurement biopsy over preimplantation or post-

reperfusion biopsies is the timing; early evaluation of biopsy pathohistology can influence the

allocation of the organ prior to prolonged periods of cold ischemia incurred during transit [96]. A

donor kidney biopsy can demonstrate chronic changes, as well as the degree of pathologic

deterioration, and help with the assessment of the kidney’s suitability for transplantation.

1.5.1.2 Preimplantation “Zero-Time” Biopsies

Preimplantation biopsies taken at the time of transplant (“zero-time” biopsies) similarly

inform clinicians and are a standard of care in some transplant centers as they reveal early

Page 48: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

31

symptoms of graft-threatening pathology [97]–[99]. Little histologic difference is expected

between procurement and preimplantation biopsies. However, if molecular markers of ischemic

damage are evaluated, preimplantation biopsies would reflect the full scope of both the warm

and cold ischemic damage together while procurement biopsies would reflect only periods of

warm ischemia. Preimplantation biopsies occur following transit and so occur too late to affect

allocation. They can however influence last-minute decisions to accept or reject a kidney of

borderline quality for transplant. Similarly, preimplantation biopsies can improve prediction of

post-transplant outcome and so may influence clinical management of the graft. Preimplantation

biopsies may also be used as a baseline for interpreting progression of pathology in later protocol

biopsies [5].

1.5.1.3 Post-Reperfusion Protocol Biopsies

Protocol biopsies are routinely performed at 6 and 12 months following transplant to

assess the health of the graft or identify incidence and form of rejection. Post-reperfusion

protocol biopsies are performed during transplant, following re-anastomosis and prior to skin

closure of the recipient (i.e. while the kidney is still accessible). While post-reperfusion biopsies

occur too late to affect the decision of whether to accept or discard a kidney, they are

advantageous over pre-implantation biopsies in that they reveal the extent of IRI. Koo et al.

demonstrated IRI in post-reperfusion biopsies, evidenced by infiltration of recipient-derived

neutrophils [100]. Haas et al. confirmed this observation and similarly demonstrated donor-

derived neutrophil infiltration [101]. Kanellis et al. similarly demonstrated IRI in post-

reperfusion biopsies, evidenced by acute c-Jun N-terminal kinase JNK activation in tubular

epithelial cells [102]. In each study, cadaver donor transplants were included as models of IRI

Page 49: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

32

while live donor kidney transplants were included as controls which did not experience IRI (live

donor kidneys are not subjected to the prolonged cold ischemia which precedes IRI). The time

between reperfusion and the post-reperfusion biopsy is not an established duration and so is

variable between patients. This is a potentially serious limitation of the post-reperfusion biopsy

as this variability may affect the stage of IRI suffered by the graft.

1.5.1.4 Remuzzi Scoring of Biopsies

The Remuzzi scoring system was introduced in 1999. Remuzzi et al. assigned scores for

each kidney from 0 to 3 for each of 4 features. Glomerular sclerosis, TA, and IF were rated from

0 to 3 where 0 indicated absent pathology, 1 indicated <20% of biopsied physiology affected, 2

indicated 20 to 50% affected, and 3 indicated more than 50% affected. Arterial and arteriolar

narrowing were similarly rated from 0 to 3 where 0 indicated absent narrowing, and 1 to 3

indicated varying degrees of increased wall thickness relative to the diameter of the lumen. The

final score ranged from 0 to 12, with cumulative scores from 0 to 3 considered mild, scores from

4 to 6 considered moderate, and scores from 7 to 12 considered severe pathology. Kidneys with

mild, moderate, and severe ratings were recommended for single, double, or no transplant

respectively. This scoring system was suggested to promote efficient use of an expanded donor

pool as evidenced by graft survival, however sampling in the study was limited (24 recipients of

2 marginal kidneys and 48 recipients of kidneys from standard donors) and the score was not

validated on an independent population [103].

Page 50: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

33

1.5.1.5 MAPI Scoring of Biopsies

In 2008, Munivenkatappa et al. proposed an alternative scoring system, the Maryland

Aggregate Pathology Index (MAPI). Munivenkatappa et al. utilized cox proportional hazard

methods to identify histological features which correlated with graft loss. Variables found to

demonstrate significant correlation included percentage of glomerulosclerosis, presence of

periglomerular fibrosis (thickening, wrinkling and reduplication of Bowman’s capsule), vascular

pathology (ratio of arterial wall to lumen), presence of AH (amorphous, homogenous

eosinophilic deposits in the wall of arterioles), and presence of scarring (sclerosis and renal

parenchymal fibrosis and atrophy of 10 tubules or more). The hazard ratio (HR) of each feature

was rounded to the nearest integer, and all rounded ratios summed to produce a cumulative score

of 0 to 15 “MAPI points”. A MAPI score of 0 to 7 was designated as representative of a kidney

with a low risk of graft failure. A score of 8 to 11 was designated as intermediate risk, and a

score of 12 to 15 was designated as high risk. Low, intermediate, and high risk kidneys were

associated with 10%, 37% and 47% graft failure at 5 years post-transplant [104].

Page 51: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

34

Figure 1.12: Histopathology in MAPI biopsy scoring. (a) Periglomerular

fibrosis, (b) arteriolar hyalinosis, (c) scar including features of interstitial

fibrosis, tubular atrophy and glomerulosclerosis, and (d) measurements for

arterial wall-to-lumen ratio (WLR) calculation, including the thickness of two

opposing walls (T1 and T2) and the luminal diameter (LD). WLR=(T1

+T2)/LD. [104]

1.5.1.6 Banff Scoring of Biopsies

In an effort to standardize biopsy scoring and interpretation, a coalition of renal

pathologists, nephrologists, and transplant surgeons met in Banff, Canada in 1991. The Banff

group initially worked to refine criteria for assessment of post-transplant biopsies in diagnosis of

instances of rejection. In 2010, a Banff working groups was established to similarly refine

criteria for assessment of procurement biopsies [99], [105]. They continue to meet at regular

intervals to consolidate terminology, scoring, and classification criteria [106]. The Banff working

group on procurement biopsies initially conducted a mass survey in 2011 of pathologists to

establish areas of focus. In 2017, a follow-up Banff histopathological consensus criteria for

assessing procurement biopsies was published. This scoring system graded IF, TA, interstitial

inflammation, arterial intimal fibrosis, AH, glomerular thrombi, and ATI as non, mild, moderate,

or severe based on the percentage of relevant anatomy effected [107].

Page 52: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

35

1.5.2 Visible and Relevant Features in Optical Coherence Tomography Imaging of the Human

Kidney

OCT imaging (which will be discussed in detail in the next chapter) is an interferometry-

based imaging modality, capable of visualizing the superficial portions of the kidney at very high

resolutions (Figure 1.13). OCT allows us the opportunity to visualize many of the structures

which are assessed and established in traditional biopsies as predictors of poor graft viability.

OCT has the added benefit of being non-invasive, meaning cortex structures can be seen without

the tearing and artifacts introduced by biopsies. Banff criteria for biopsies stipulate that at least 2

areas should be sampled in the kidney to provide an accurate assessment of viability; OCT

allows for global imaging of the kidney so heterogeneous distributions of pathology can be

properly realized [105].

Figure 1.13: Representative B-scan captured in the operating room of a

donor kidney (pre-implantation). (a) The original greyscale B-scan. (b) The

Tegaderm film highlighted in red, the renal capsule highlighted in blue and

the kidney cortex highlighted in green.

The renal capsule is readily visible in OCT imaging. While the thickness of the renal

capsule is not known to coincide with kidney function or graft viability, its structure may offer

insight into pathology. For example, fibrotic scarring in the kidney cortex can induce contraction

and produce depressions or divots in the renal capsule near the point of scarring (red arrow in

Page 53: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

36

Figure 1.14) [108]. Assessment of this in kidney grafts or even in-vivo (through laparoscopic

OCT) in patients with chronic kidney disease (CKD) may help in identification and

quantification of fibrosis and the progression of fibrotic symptoms respectively.

Figure 1.14: B-scan of a human kidney with capsular depressions

captured ex-vivo prior to transplant. The red arrow indicates a point of

depression (divot) in the renal capsule which may represent a symptom of

fibrotic scarring. The green arrow indicates a superficial glomerulus.

OCT is also proficient in identifying sub-capsular cysts (red arrows in Figure 1.15). Cysts

are common in aging, especially in persons over 50, and are not generally directly associated

with kidney function [109]–[111]. In some instances, however, the presence of cysts may suggest

polycystic kidney disease, or may promote incidence of renal cell carcinoma in the recipient if

transplanted [86], [112].

Page 54: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

37

Figure 1.15: B-scan of a human kidney with superficial cysts.

Particularly in kidneys from older patients, cysts often appeared in the

superficial cortex and were often numerous within a single kidney. Cysts

ranged in size but were routinely flush with the capsule-cortex interface.

In the study described in this Thesis, Glomeruli appeared in the OCT imagery in less than

half of the kidneys imaged. The number of visible glomeruli varied widely between kidneys, and

the depth of the glomeruli greatly affected distinguishable features (Bowman’s space, and the

capillary tuft) (Figure 1.16). While glomerulosclerosis is accepted as an important feature relative

to graft function, little information is available on the relevance of only superficial glomeruli to

graft function.

In more superficial glomeruli, where resolution was improved, collapsing FSGS seemed

to be apparent in certain kidneys (red arrow in Figure 1.16). In these instances, the glomerular

tuft appeared shrunken while the capsular space of Bowman remained fully dilated. In contrast,

many glomeruli appeared with a large capillary tuft which nearly occupied the full capsular

space of Bowman (green arrow in Figure 1.17). A 3D protocol would be advantageous in

Page 55: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

38

evaluation of this pathology; 3D imaging of glomeruli would provide accurate measurement of

the size of capillary tufts relative to the surrounding capsular space of Bowman.

Figure 1.16: B-scan of human kidney with superficial glomeruli. The

green arrow indicates superficial glomeruli, the red arrow indicates what

is likely a collapsed capillary tuft, indicative of collapsing FSGS.

In some cadaver donor kidney transplants imaged, bright sub-capsular features were

noted. These were diffuse throughout the cortex, but represented globally throughout each of the

3 kidneys. While it is challenging to tell from 2D image sets, the bright features appeared to be

vascular; the features appeared somewhat linear relative to the convoluted nature of the proximal

tubules, and the diameter appeared larger than what would be expected in PCTs. This may

represent some vasculopathy, potentially arteriosclerosis. Similarly, diameter measurements of

vascular lumen present in OCT imaging may provide a non-invasive method of quantifying

arteriolar narrowing.

Page 56: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

39

Figure 1.17: B-scan of human kidney with dense vascular features. B-

scan from one of 3 kidneys which exhibited dense vascular features (red).

The green arrow indicates a glomerulus.

Capsular depressions, subcapsular cysts, glomeruli and vasculopathy evident in OCT may

provide valuable insight into pre-existing pathology within a donor kidney. These features,

however, are not visible in every kidney (with the exception of the capsule), and were not the

direct objective of this thesis. The PCTs occupy the majority of the superficial kidney cortex, and

so are readily available for imaging by OCT. Abnormalities in PCT morphology may also

represent symptoms of ischemic insult in addition to pre-existing pathology, making them an

ideal candidate for analysis.

1.6 Summary

OCT imaging of the superficial cortex reveals primarily the lumen of the PCTs and some

other features (distal convoluted tubules, glomeruli, vasculature, and superficial cysts) in less

frequency. Visualization of PCT lumen is sufficient for evaluation of much of the pathology

assessed in traditional biopsies; reduction in the area of tubular lumen may suggest PCT swelling

while dilation of the tubular lumen could suggest tubular simplification, partial EMT,

hypertrophy, TA, or epithelial sloughing. Similarly, distances between adjacent tubular lumen

Page 57: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

40

may offer clues as to accumulation of interstitial pathology. Areas where no lumen are apparent

may reflect atrophied or fibrotic regions of the kidney, although this is purely speculative.

Identification of pre-existing pathology (e.g., IFTA), made visible by OCT at the time of

procurement, may guide procurement biopsies towards regions of pathology. Similarly, OCT

may help in interpretation of traditional procurement biopsies by providing pathologists with a

global distribution of pathology which they have identified in biopsies captured locally. Used in

tandem with traditional procurement biopsies, or potentially independently, OCT can provide

surgeons and pathologists with a better assessment of pre-existing pathology in kidneys at the

time of retrieval and so can aid in kidney allocation.

While pre-implantation (zero-time) biopsies are infrequent, OCT may similarly aid in

their interpretation. As with procurement biopsies, OCT may be able to guide pre-implantation

biopsies towards sites of pathology (e.g., ATI) and/or offer a global view of the distribution of

pathology seen in pre-implantation biopsies. Pre-implantation OCT could similarly be used

independently, offering surgeons a non-invasive indication of pre-existing pathology and

ischemic damage. While this would all occur too late to effect allocation of the kidney, this

information could factor into last-minute decisions to accept or reject a kidney and could inform

post-operative diagnoses and care.

If the degree of ischemic damage could be accurately determined prior to transplant from

symptoms of ATI (swelling, simplification, etc.), the degree of IRI which would ensue following

reperfusion could be ascertained. This measure could provide a predictive value for DGF.

Predicting IRI-induced DGF in this fashion could prove useful in diagnosis following transplant,

Page 58: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

41

allowing physicians to distinguish between delays in recovery due to ischemic damage from

delays in recovery due to an immune response. This distinction could support the use of less

nephrotoxic immunosuppressant in patients experiencing a delay in function (i.e. alternatives to

calcineurin inhibitors)[113]. More importantly, an accurate measure and quantification of the

accumulation of factors from ischemic damage that would contribute to IRI could enable a

widening of the donor pool through inclusion of kidneys which based purely on the duration of

cold and warm ischemia would normally be discarded for an assumed degree of ischemic

damage but which may still be viable due to an increased resistance to ischemic insult relative to

the mean kidney or due to some undetermined variable.

OCT imaging performed following transplant may similarly highlight pre-existing

pathology, and ischemic damage. Reduction in ischemic swelling following reperfusion may

unveil pathology hidden prior to reperfusion, allowing more accurate assessment of the graft. In

addition, the dissipation of swelling would provide a benchmark of native lumen structure from

which to gauge ischemia-induced swelling visualized in pre-implantation OCT scans. In

addition, OCT imaging post-transplant may reveal inflammatory effects associated with IRI,

which again may inform post-operative diagnostics and care.

Page 59: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

42

CHAPTER 2: IMAGING AND IMAGE PROCESSING

2.1 Introduction

We open this chapter with a brief introduction to non-invasive imaging modalities used in

the context of kidney transplantation. Next, we provide an introduction to OCT: offering an

overview of the guiding principles behind interferometry, discussing the different forms of OCT,

and highlighting medical fields where OCT has found use. We then provide an introduction to

image processing: we summarize several common pre-processing steps, including image

smoothing and contrast enhancement, we go on to summarize several common strategies for

edge and region detection, and then explore their use in the segmentation of OCT images. We

conclude this chapter with a brief discussion on the advantages of OCT over other imaging

modalities for assessing donor kidney viability, and a brief discussion on the motivation behind

selection of specific image processing tools for segmentation of OCT images of human kidneys.

2.2 Non-Invasive Imaging Modalities Used in Kidney Transplantation

Non-invasive imaging of kidneys prior to and/or following transplant may provide a

valuable supplement to current measures of viability. Distinct advantages of non-invasive

imaging modalities over biopsy are: reduced risk of dangerous complications, global assessment

to remove biasing of patchy pathology, and removal of artifacts and perturbations to anatomy

incurred by sectioning and slide preparation. Several non-invasive imaging modalities, each with

their own benefits and drawbacks, have been introduced into the kidney transplant arena.

Page 60: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

43

Magnetic resonance imaging (MRI), high-resolution computed tomography (CT), ultrasound,

OCT, and confocal microscopy have all been used in some capacity to aid in the assessment of

transplant kidneys. Some of these have been used, with varying success, in attempts to visualize

pathology at a macro level. Others have been used and are now commonplace in guidance of

protocol biopsies into the kidney cortex of a transplant recipient. Each modality carries with it a

penetration and resolution trade off. Consequently, modalities which can image more of the

kidney tend to lack the resolving power to discriminate finer features. Conversely, modalities

which can reveal fine kidney microstructure tend to be limited in the area that can be imaged

(Figure 2.1).

Figure 2.1: Diagram of optical coherence tomographgy penetration

and resolution in relation to other imaging modalities. [114]

MRI has been used in kidney transplantation primarily to assess graft function in

recipients following transplant. Hueper et al. used functional magnetic imaging (fMRI), with T2

Page 61: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

44

mapping and diffusion weighted imaging (which measures the displacement of water in tissue) to

evaluate edema in the preserved mouse kidney. Using T2 mapping they were able to characterize

the water content within various sections of the kidney and did observe significant swelling in

the outer and inner medulla following 45 minutes of normothermic ischemia [115]. Steiger et al.

similarly investigated diffusion weighted MRI (DW-MRI) in the assessment of kidneys. Steiger

et al. employed a combination of qualitative and quantitative assessment of transplanted kidneys,

using the apparent diffusion coefficient (ADC) as a marker and mapping of potential pathology

(ADC reveals the magnitude of diffusion of water in the sample) in the transplanted kidney

(Figure 2.2) [116].

Figure 2.2 Diffusion weighted MRI of human kidneys following

transplant. Upper row: Patient in the “normal or mild histopathologic

changes” group with mild histopathologic changes showing a homogenous

ADC and fmap. Lower row: Patient in the “severe histopathologic changes”

group with an acute tubular necrosis and a BK virus nephropathy showing a

heterogeneous ADC and f map. On morphologic T1w and T2w images the

kidneys of patients in the two groups cannot be distinguished. [116]

CT and ultrasound have both been used in the kidney transplant arena. Both modalities

have been employed to guide protocol biopsies, although ultrasound has emerged as the industry

Page 62: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

45

standard. Standards for adequacy of biopsy specimen set forth by the Banff ’97 conference

required that biopsies contain 10 or more glomeruli and at least 2 arteries [105]. The most

common strategy to achieve this standard is with the cortical tangential approach. In the cortical

tangential approach, CT or ultrasound are used to guide placement of the biopsy needle into the

kidney cortex, parallel to the renal capsule. This is intended to maximize cortical tissue sampled,

and to avoid puncture of the renal collecting system and large vascular structures. This method

does not direct the biopsy towards any site of presumed pathology and does not specify a

standardized location on the kidney where the cortical tangential biopsy should be captured

[117], [118].

Tandem scanning confocal microscopy has been used previously by Andrews et al. to

diagnose animal models of ischemia-induced acute tubular necrosis (ATN) by visualization and

quantification of swelling of the PCTs. The tandem scanning system however is costly, highly

impractical in an OR setting and lacks the penetration depth to visualize past the renal capsule

which is present in humans and absent in the animal models used [42].

Using OCT, Andrews et al. was able to acquire high-resolution 3D images of the human

kidney cortex and PCTs [119]. While the depth of imaging penetration for OCT is relatively

shallow, the technique proved sufficient to penetrate into the renal cortex. OCT was able, with

8um axial resolution, to accurately visualize the lumen of the proximal convoluted tubules,

permitting accurate measurement of PCT anatomy. This system has the added benefit of being

relatively affordable, portable, and conducive to an OR setting.

Page 63: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

46

Figure 2.3: Side by side comparison of proximal convoluted tubules

visualized by different methods. (a) OCT (b) tandem scanning confocal

microscopy (c) frozen biopsy sections and (d) paraffin embedded sections.

2.3 Optical Coherence Tomography

Optical coherence tomography (OCT) is a relatively new imaging modality which can

produce high speed, high resolution depictions of biological structures from the light scattering

characteristics of the target tissue. OCT occupies a niche in the penetration:resolution spectrum

of non-invasive imaging modalities between confocal microscopy and ultrasound. Penetration of

OCT is typically limited to a few millimeters, an order of magnitude less than conventional

ultrasound and further still from the full-body penetrating power of MRI, and X-ray CT. OCT

penetration, however, surpasses that of confocal microscopy (typically limited to a few hundred

microns) by an order of magnitude. OCT can reach micron-level resolution, moderately reduced

compared to confocal microscopy but significantly improved over the resolving power of

ultrasound, CT, and MRI. Like ultrasound, the high speed of OCT permits video rate capture of

tomographs without concern of significant contribution from parasitic motion (e.g., eye

movement in retinal OCT scans). Like MRI and CT, OCT does not require direct contact with

the imaged sample (which is a significant benefit when imaging, for example, the eye of a non-

sedated patient) [120], [121].

Page 64: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

47

OCT is an interferometry based modality, analogous to ultrasound, which utilizes light

waves as opposed to sound waves to generate depth-wise profiles of the light scattering

properties of an imaged sample [121].

Interference patterns occur when waves combine. Two waves of the same frequency and

phase will combine in an additive manner with the amplitude of resulting peaks and troughs

being doubled (constructive interference). Conversely, two waves of the same frequency with a

half-wavelength phase shift (i.e., the peaks of one wave match the troughs of the second wave)

will combine in a subtractive manner, essentially cancelling each other out (destructive

interference) (Figure 2.4). Waves with a phase shift of less than or more than a half-wavelength

will produce an interference pattern with a combination of constructive and destructive

interference. The degree of phase-shift can be extrapolated from these interference patterns

[122].

Figure 2.4: Constructive and destructive interference. [123]

OCT takes advantage of this effect by splitting a beam in two, directing some portion of

the wave to a mirror (reference arm) and the remainder to the tissue of interest (sample arm), and

Page 65: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

48

combining the waves reflected from each. Assuming the path length of each arm is the same, the

resulting interference pattern should provide information about the light scattering characteristics

of the sample. If the sample was a second mirror and path lengths were equal, reflected waves

would combine constructively. If the sample was tissue which backscatter the wave with a half-

wavelength time delay, reflected waves would combine destructively.

Strong reflection occurs in tissue where the scattering coefficient is high (e.g., collagen),

and at transitions between tissues which slow light at different rates (tissue with different

refractive indices). Weak reflection occurs in media with a low scattering coefficient (e.g., clear

fluids), and over regions with homogenous refractive indices. Collection of scattering properties

at different depths (an axial scan, or A-scan) is accomplished by either Time-Domain or Fourier-

Domain OCT methods. The resulting axial scans are 1-dimensional, and are captured

sequentially in the transverse direction by rotation of the sample arm mirror to generate 2-

dimensional images (B-scan). For 3D tomographs, B-scans are captured sequentially and

combined [121], [124].

Since OCT’s introduction in the early 1990’s, it has found prominent use particularly in

ophthalmology. OCT is able to penetrate through the cornea, lens, and vitreous humor to reach

the retina with little dissipation of signal. Once at the retina, OCT can visualize each retinal layer

with higher resolution than any other non-invasive imaging modality, and can penetrate beyond

the choroidal-scleral junction. OCT enables ophthalmologists to visualize retinal layer thickness,

and fluid filled pathology like cyst formation in macular edema. In addition, OCT can aid in the

diagnosis and progression monitoring of macular degeneration (age and non-age related),

Page 66: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

49

diabetic retinopathy, idiopathic serous chorioretinopathy, and a litany of other retinal pathologies

by revealing irregularities in retinal layer thickness (Figure 2.5) [105], [125]–[128].

Figure 2.5: Optical coherence tomography B-scans of age-related

macular degeneration, diabetic macular edema, and the healthy

retina. [129]

OCT has similarly emerged in other fields where desired clinical features are more

superficial (e.g., dentistry, and dermatology). In dentistry, OCT reveals morphology of soft and

hard dental features, and can be used in diagnosis of caries (tooth decay), periodontal disease and

oral cancer. In dermatology, OCT is used to visualize superficial skin layers, facilitating

evaluation and quantification of skin lesions in non-melanoma skin cancers and inflammatory

disease [130], [131]. OCT can also be implemented in an endoscopic or catheter form, enabling

its introduction into interventional cardiology, gastrointestinal endoscopy, laryngology, and

gynecology. Endoscopic OCT is used not just in pathology detection, but in biopsy site selection,

and stent placement [132].

Page 67: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

50

2.3.1 Time-Domain Optical Coherence Tomography (TD-OCT)

The simplest and most intuitive to understand form of OCT is Time-Domain OCT. In

TD-OCT, the reference arm is moved along the propagation direction of light, increasing and

decreasing the path length from the beam splitter. Waves from each arm are reflected and return

to combine at the beam splitter. Interference patterns are then generated for waves which

travelled equal path lengths. As the reference arm path length increases, the equivalent optical

distance in the sample arm penetrates deeper into the sampled tissue. Interference patterns of

equivalent path lengths combine at each movement of the reference to form a depth-wise axial

scan.

The complicated part of TD-OCT is in isolating the point where path lengths are equal.

As the path length of the reference arm is increased, the corresponding path length in the sample

moves from superficial to deep. Light is reflected from the deeper location in the sample, but

also reflected from all superficial points which the light passes through. TD-OCT uses properties

of low-coherent light to circumvent this problem. Low coherence occurs when the phase

difference changes over time, a byproduct of waves with different frequencies. Overlain,

sinusoidal waves of different frequencies will align moderately well (low phase shift) near their

source but will line up less and less well (higher phase shifts) at increasing distance and from the

source (until the lower frequency starts to lap the higher frequency). The distance along a set of

waves where the phase shift is minimal can be considered the coherence length. The more

frequencies that are included (the broader the spectral bandwidth), the shorter the coherence

length will be.

Page 68: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

51

TD-OCT generally employs a low-coherence semiconductor super-luminescent diode

(SLD) as its light source. When a low coherence light source is split and recombined,

interference patterns are only discernable when the coherence lengths of the two beams are of the

same pathlength. When detecting all returning light from the reference and sample arms

(including light reflected from all depths of the sample), coherence lengths will be in sync for the

reference arm and the light reflected from the equivalent optical distance in the sample. Scattered

light from other depths will be effectively filtered out, allowing TD-OCT to generate a depth-

wise scattering profile.

One important takeaway is that the coherence length of the broadband source determines

the axial resolution; a shorter coherence length means a smaller section of the wave combination

can be isolated, increasing resolving power. Transverse (or lateral) resolution, however, is

independent of axial resolution. Lateral resolution is dependent on the spot size of the focused

interrogating beam, which is in turn dependent on the numerical aperture (ability to gather light

and resolve fine detail) of the focusing lens [124], [133].

2.3.2 Fourier-Domain Optical Coherence Tomography (FD-OCT)

In contrast to TD-OCT, the reference arm of Fourier-Domain OCT systems is immobile

and the light scattering profiles are captured at all axial depths simultaneously. This method is

considerably faster than TD-OCT (50-100x), as it is not limited by movement of the reference

mirror. FD-OCT can be further subdivided into spectral-domain OCT (SD-OCT) and swept-

source OCT (SS-OCT), but both methods rely on the same principal of utilizing Fourier

transforms of frequency spectrums to extract interference patterns at different depths (spectral

Page 69: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

52

interference). Different wavelengths respond differently to the refractive indices of imaged

samples; shorter and higher wavelengths move faster and slower, respectively, through tissue (an

effect termed “dispersion”). FD-OCT systems emit a spectrum of frequencies, which at any one

point in time will have high frequencies (lower wavelengths) which travelled further through the

sample than low frequencies over the same amount of time. By emitting a broad range of

frequencies simultaneously or rapidly in succession, FD-OCT produces an interference pattern

for each frequency at varying depths in the sample. Fourier transformation of backscattered light

moves the time domain wave combinations into the frequency domain, and allows interference

patterns to be mapped axially.

In SD-OCT, this range of frequencies is generated immediately prior to detection with a

diffraction grating (splits and diffracts light) and detected by a high-speed CCD (charged

coupled device) line camera (a high resolution spectrometer, or array-type detector which detects

each frequency independently at different locations). In SS-OCT, a tunable narrow linewidth

scanning laser sweeps through a range of wavelengths sequentially, and accumulates the spectral

interference one wavelength at a time with a single detector. SS-OCT generally provides deeper

penetration depth than SD-OCT as well as faster imaging speeds (SD-OCT is limited in scanning

speed by the read-out rate of its CCD line camera). SD-OCT, however, generally provides higher

axial resolution than SS-OCT [121], [124], [126].

Page 70: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

53

Figure 2.6: System design for time and Fourier domain optical

coherence tomography systems. [120]

2.4 Image Processing

In signal processing, “filtering” is the act of accepting or rejecting components of certain

frequencies. In image processing, this can be used to accentuate or suppress image features to

facilitate further processing steps or to aid in interpretation of the data. This process cab be

achieved by convolution of a specified filter kernel over the original image, with the product of

the applied filter at each pixel point yielding the value of the corresponding pixel location in an

output (filtered) image. Image filtering can be utilized to smooth (remove noise) or sharpen

(accentuate) edges an image. Smoothing and sharpening operations can be accomplished by

convolution with low-pass and high-pass filters respectively. Low-pass filters iterate over an

image, removing high-frequency while retaining low-frequency features, effectively smoothing

the image. Conversely, high-pass filters iterate over an image, removing low-frequency while

maintaining high-frequency features, effectively highlighting contours in an image.

Page 71: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

54

2.4.1 Low-Pass Filter (Blurring Mask)

Medical images are routinely corrupted by variations in intensity values which occur

independently of signal representing anatomical structure. This effect is deemed “noise,” and can

be further subdivided into salt and pepper noise, impulse noise, and Gaussian noise. Salt and

pepper noise produces random instances of both high and low intensity values (i.e. black and

white), while impulse noise contains only instances of high intensity values. Gaussian noise

produces variations in intensity drawn from a normal distribution and are often present as sensor

noise. The presence of noise greatly reduces the efficiency of most edge and region detection

methods. Noise reduction steps are therefore a common prerequisite to any image processing

task. Noise reduction, or smoothing, involves decreasing the dissimilarity between nearby pixels

by averaging values over a localized window.

Low pass filters, or “blurring-masks”, are the foundation of most smoothing strategies.

Low pass filter kernels generally contain only positive values, with all values summing to one.

Kernels are most often odd and symmetric in window size (i.e. 3x3, 5x5, 7x7, etc.). With most

low-pass filters, the larger the window size, the greater the degree of smoothing and noise

reduction in the filtered image. Smoothing operations, however, produce the unwanted side-

effect of edge blurring; this must be considered when choosing the size or type of filter.

Reduction of noise must be balanced with preservation of edge strength to ensure that regions of

interest are discriminable following smoothing [134].

Page 72: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

55

2.4.1.1 Mean Filter

The simplest and most intuitive smoothing filter is the mean filter, or box filter. The

mean filter is an example of a linear operator, in that the output is a linear combination of inputs

from the original image. Mean filters convolve a kernel of a specified window size across the

original image, calculating the mean pixel value of the local neighborhood (all pixels in the

specified window size, centered on the target pixel). The corresponding pixel in the output image

receives the averaged value, producing a “smoothed” copy of the original input. While mean

filters are effective in noise reduction, they are particularly insensitive to sharp discontinuities

(i.e., they tend to blur edges more than other filters) [134].

Figure 2.7: Mean (box) filters of different windows sizes applied to optical

coherence tomography scan of human kidney. (a) original unfiltered image

(b) smoothed by a 3x3 mean filter (c) smoothed by a 5x5 mean filter (d)

smoothed by a 7x7 mean filter (e) smoothed by a 9x9 mean filter.

2.4.1.2 Median Filter

Median filters work similarly to mean filters in that they convolve over the image with an

evenly weighted kernel of a specified window size, although they are nonlinear operators. As the

name suggests, median filters then assign median values of local neighborhoods from the input

image to the output image. Median filters are proficient in removing salt and pepper, as well as

Page 73: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

56

impulse noise. Median filters remove outlier pixel values but tend to retain image details better

than mean filters, and so produce less blurring and edge deterioration than is seen in mean filters

of identical window sizes [134].

Figure 2.8: Median filters of different windows sizes applied to optical

coherence tomography scan of human kidney. (a) original unfiltered image

(b) smoothed by a 3x3 median filter (c) smoothed by a 5x5 median filter (d)

smoothed by a 7x7 median filter (e) smoothed by a 9x9 median filter.

2.4.1.3 Gaussian Filter

While the choice of filter is unique to the task at hand, Gaussian smoothing is by far the

most frequently used smoothing technique in image processing. Gaussian smoothing is another

linear filter but deviates from mean or median filters in that it performs a weighted average.

Whereas in a mean or median filter, each pixel location from the local neighborhood in the input

image receives an equal weight (e.g., a 3x3 mean filter would produce an output with the sum of

the values of the 9-pixel neighborhood divided by 9), Gaussian filters apply weights to each

pixel’s value based on a Gaussian function centered around the center pixel. The pixel-wise

output of the Gaussian filter is therefore an average of the corresponding local neighborhoods of

the input image, with more weight attributed to the center pixel in each neighborhood and weight

decreasing monotonically with distance away from the center pixel. This method is particularly

Page 74: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

57

effective in reducing noise from a normal distribution and tends to retain important features due

to its’ preferential weighting of centered pixels. Finally, Gaussian filters are advantageous in that

they are isotropic, or rotationally symmetric (i.e., weights are applied circularly emanating out

from the center pixel), while mean or median filters generally draw evenly from values spread

across a square neighborhood. This has the effect of not biasing the output image in any one

direction, which is especially important when subsequent steps involve edge detection. These

characteristics make Gaussian filters a very effective and reliable low-pass filter for pre-

processing.

In Gaussian smoothing, the sigma parameter is tantamount to window size in mean or

median filters; sigma determines the width and spread of the Gaussian distribution of weights. A

smaller sigma gives greater weight to the center pixel with more rapidly decreasing weights as

you move from the center pixel. A larger sigma flattens out the distribution of weights; weights

are still greater towards the center pixel but the decrease in weights moving away from the center

pixel is less steep. A smaller Gaussian sigma essentially produces a smaller filter which does less

smoothing but retains more edge structure, while a larger Gaussian sigma essentially produces a

wider filter which does greater degrees of smoothing but leads to loss of fine structure [134],

[135]

Page 75: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

58

Figure 2.9: Gaussian filters of different sigma values applied to optical

coherence tomography scan of human kidney. (a) original unfiltered image

(b) smoothed by a gaussian filter with a sigma of 1 (c) smoothed by a

gaussian filter with a sigma of 3(d) smoothed by a gaussian filter with a sigma

of 5 (e) smoothed by a gaussian filter with a sigma of 7.

2.4.2 High-Pass Filter (Sharpening Mask)

High-pass filters, or “derivative masks,” are the foundation for most image sharpening

and edge enhancement strategies. Similar to low-pass filters, high-pass filter kernels are most

often odd and symmetric in window size (i.e. 3x3, 5x5, 7x7, etc.). High-pass kernels, however,

generally contain both positive and negative values, with all values summing to zero. By

including both positive and negative weights which sum to zero in high-pass filters, differences

can be highlighted between pixels which are assigned positive values and pixels which are

assigned negative values (i.e., the output value is greater when there is a stronger disagreement

between pixel values with positive and negative weights). A kernel with a positive central weight

surrounded by negative neighboring pixel weights can accentuate fine structure by highlighting

differences between a pixel with its’ surrounding pixels (e.g., Laplacian operator). Similarly, a

kernel with positive weights on one side of the central pixel and negative weights on the opposite

side can highlight abrupt linear changes in intensity (edges) in an image (e.g., Sobel operator).

Sharpening operations, it should be noted, tend to amplify noise. With most high-pass filters, the

larger the window size, the less the amplification of noisy features. However, larger window

Page 76: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

59

sizes tend to reduce the ability of high-pass filters to localize edges. The amplification of noise

can similarly be diminished by first applying a low-pass filter to the original image as a

preliminary step. Application of smoothing filters prior to edge enhancement can, however, also

reduce the ability of high-pass filters to localize edges. The appropriate preliminary smoothing

operations and high-pass filter window size must therefore be carefully considered, as there is a

notable tradeoff between the dampening of noise amplification and edge localization.

There are a wide array of edge enhancing filters which can be generally subdivided into

two categories: first derivative and second derivative filters. Edges occur in an image when there

is a steep gradient in intensity values in a particular direction between adjacent pixels or groups

of pixels. First derivative filters quantify this change in values. This is often accomplished by the

combination of partial derivatives; gradient is calculated in the x-direction (with positive kernel

weights positioned above or below the central kernel pixel, and negative weights positioned

opposite) and y-direction (with positive kernel weights positioned to the left or right of the

central kernel pixel, and negative weights positions opposite) separately and summed [134]. The

Prewitt and Sobel operators are a commonly used examples of this method.

2.4.2.1 Prewitt Filter

The Prewitt filter attempts to estimate the gradient magnitude by assigning equal weights

to pixels on each side of the central pixel. For a 3x3 x-direction operator, this yields a kernel with

central 3-pixel column weights of 0, weights of positive 1 in the column of three vertical pixels

to the left or right of the central pixel, and weights of negative one in the opposite column. This

kernel convolves over the image and produces high output values when there is a strong gradient

Page 77: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

60

or large difference between values on the right and left of the central pixel (vertical edge). The

kernel is then transposed and convolved over the image to produce the partial derivative for the

y-direction. The partial derivatives are summed and provide a fair estimate of edge strength in

the original image. Prewitt filters are simple and easy to implement but are highly sensitive to

noise [136].

Figure 2.10: Prewitt operator applied to optical coherence tomography B-

scan of the human kidney. (a) kernels for horizontal and vertical Prewitt

operators (b) gradient output of horizontal Prewitt filter (c) gradient output of

vertical Prewitt filter

Page 78: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

61

2.4.2.2 Sobel Filter

The Sobel filter is similarly a first derivative operator which convolves over the original

image producing two partial derivative outputs which sum to produce gradient magnitude

estimates. The Sobel filter deviates from the Prewitt filter by assigning greater weight to pixels at

a lesser Euclidean distance from the central pixel. Using a 3x3 x-direction operator again as an

example, this yields a kernel with central 3-pixel column weights of 0, weights of positive 1,2,

and 1 from top to bottom in the column of three vertical pixels to the left or right of the central

pixel, and weights of negative 1, 2 and 1 in the opposite column. This moderate adjustment to the

weighting scheme has the effect of approximating a Gaussian distribution of weights. As a

product of this effect, the Sobel filter essentially contains a built in Gaussian smoothing filter and

so is less sensitive to noise than, for example, the Prewitt filter.

In addition to estimates of gradient magnitude, both the Prewitt and Sobel filter provide

information about the gradient direction. For example, a 3x3 x-direction Prewitt filter with

weights of negative 1 in the column to the left of the central pixel, weights of zero in the column

of the central pixel and weights of positive 1 in the column to the right of the central pixel

provides an output value with both direction and magnitude of gradient. If the 3x3 neighborhood

in the original image contains intensity values of 10 to the left of the central pixel and intensity

values of 1 to the right of the central pixel, the output of the applied kernel will equal -27 ((-

1x10)+(-1x10)+(-1x10)+0+0+0+(1x1)+(1x1)+(1x1)). The absolute value, 27, provides the

magnitude of the gradient, while the negative sign provides the direction (leftward). Conversely,

if the 3x3 neighborhood in the original image were reversed with intensity values of 10 to the

right of the central pixel and intensity values of 1 to the left of the central pixel, the output from

Page 79: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

62

the same kernel would yield 27 ((-1x1)+(-1x1)+(-1x1)+0+0+0+(1x10)+(1x10)+(1x10)). The

absolute value, 27, provides the magnitude of the gradient, while the positive sign provides the

direction (rightward). Gradient direction can be useful when performing edge detection, allowing

users to unify adjacent high gradient magnitude pixels with similar gradient direction [137].

Figure 2.11: Sobel operator applied to optical coherence tomography B-

scan of the human kidney. (a) kernels for horizontal and vertical Sobel

operators (b) gradient output of horizontal Sobel filter (c) gradient output of

vertical Sobel filter.

Page 80: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

63

2.4.2.3 Laplacian Filter

Second derivative filters measure the rate of change of first derivatives (gradients).

Instead of reflecting change in intensity values between adjacent pixels or groups of pixels, the

output of second derivative filters reflects the abruptness of the change (i.e. a sudden change in

intensity as opposed to a progressive constant increase or decrease in intensity). The most

frequently used second derivative filter is the Laplacian operator. The Laplacian kernel is

composed of a positively weighted central pixel, surrounded by negatively or zero weighted

negative weights. This generally presents, in for example a 3x3 kernel, as either a central weight

of 8, surrounded by 8 pixels weights of -1, or more often a central weight of 4, with vertically

and horizontally adjacent pixel weights of -1 and corner pixel weights of 0. The Laplacian filter

is isotropic, and so does not require the combination of outputs from transposed kernels.

Consequently the Laplacian filter is computationally less expensive (allows for faster processing

speeds) than first derivative methods. The isotropic nature of Laplacian filters, however, means

that information regarding gradient direction is lost. Second derivative operators, such as the

Laplacian filter, are also highly sensitive to noise relative to first derivative operators. As is the

case with most high-pass filters, a preliminary smoothing with a low-pass filter can greatly

reduce the amplification of noise. Laplacian filters are routinely applied in tandem with Gaussian

smoothing, a method titled Laplacian-of-Gaussian (LoG) [138], [139].

Page 81: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

64

Figure 2.12: Laplacian operator applied to optical coherence tomography

B-scan of the human kidney. (a) kernels for diagonal sensitive (left) and

diagonal insensitive (right) Laplacian operators (b) gradient output of the

diagonally sensitive Laplacian filter (c) gradient output of the diagonally

insensitive Laplacian filter.

2.4.3 Contrast Enhancement Techniques

“Contrast” refers to the difference in intensity values between an object and its

surroundings, whereby higher contrast correlates with an increased ability to distinguish objects

or features in an image. Greyscale images generally include an intensity range of 0 to 255 (with 0

indicating absolute black and 255 indicating absolute white). Histograms displaying each

potential intensity value as a bin with each pixel value as an instance provide a visual

Page 82: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

65

representation of contrast in an image. Low contrast images exhibit a narrow range of values,

tightly clustered in a histogram. A dark image, for example, will have a narrow range of values

towards the 0 end of the histogram. Conversely a saturated image will have a narrow range of

values towards the 255 end of the histogram. High contrast images exhibit a wide range of values

and span the full width of their histograms. Interpreting an image visually is more manageable

when contrast of the image is optimized. Several techniques exist for enhancing contrast of an

image.

2.4.3.1 Contrast Stretching (Normalization)

Contrast stretching, or “normalization,” is a simple method of enhancing contrast in an

image by stretching a narrow range of values to fit the full range of possible values. In its

simplest form, contrast stretching takes the minimum and maximum intensity values in an image

and stretches them to values of 0 and 255 respectively. Pixel values between the minimum and

maximum are increased linearly to accommodate the stretch. Simply taking the minimum and

maximum as the lower and upper limits for a stretching operation, however, leaves the operation

susceptible to noise. A single aberrant low or high intensity pixel could greatly reduce the

stretch. To guard against this vulnerability, contrast stretch operations routinely plot intensity

values from the original image to a histogram and take the 1st or 5th percentile as the lower limit

and the 99th or 95th percentile as the upper limit [134].

Page 83: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

66

Figure 2.13: B-scan of human kidney before and after contrast

stretching. (a) Original B-scan (b) histogram of initial intensity values plotted

across 0-255 intensity range (c) histogram of intensity values after contrast

stretching, plotted across 0-255 intensity range (d) B-scan following contrast

stretching

2.4.3.2 Histogram Equalization

Histogram equalization is a more sophisticated form of contrast enhancement than

contrast stretching. Similar to contrast stretching, the intensity range of the image is widened to

span the full 0 to 255 spectrum. Histogram equalization, however, reassigns value in a

continuous non-linear fashion. Rather than stretching all values equally to fit the full dynamic

range, values are reassigned based on a desired distribution. Histogram equalization attempts to

flatten histograms, stretching values at sharp histogram peaks further apart. This generally

Page 84: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

67

provides higher quality enhancement than is seen in contrast stretching, but can in some

instances enhance contrast of image noise while suppressing the contrast of signal [140].

2.4.3.3 Adaptive Histogram Equalization

Adaptive histogram equalization is a localized application of histogram equalization. In

adaptive histogram equalization, an image is partitioned into tiles, and standard histogram

equalization is performed on each tile independently. It is effective in improving local contrast

and edge strength, highlighting detail throughout homogenous portions of the image. Global

reference, however, is lost in this strategy; bright or dark portions of the image are treated as

independent dark or light images, and so contrast enhanced outputs of these tiles will have a

wide dynamic range similar to other tiles even if they were brighter or darker than other tiles in

the original image [141].

Page 85: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

68

Figure 2.14: B-scan of human kidney before and after adaptive histogram

equalization. (a) original B-scan (b) histogram of initial intensity values

plotted across 0-255 intensity range (c) histogram of intensity values after

histogram adjustment, plotted across 0-255 intensity range (d) B-scan

following histogram equalization

2.4.4 Edge Detection

Edges in images are represented by abrupt changes in intensity between adjacent pixels.

In medical imaging, edges often reflect anatomical boundaries which may be of clinical

importance. Consequently, edge detection has proven to be incredibly important in many subsets

of medical image analysis. Maps of gradient magnitude and rate of gradient change, generated by

first and second derivative high-pass filters respectively, provide the input for most edge

Page 86: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

69

detection strategies. Smoothing prior to first or second order edge detection is required to

alleviate false labeling of noise as edges.

Figure 2.15: Derivatives of image intensity across edges. [142]

2.4.4.1 First Order Derivative Edge Detection

Edge detection using gradient magnitude involves the detection of local maxima or

minima of first derivatives. In a vector that passes perpendicularly over a bright line, intensity

values of the original image increase as the vector approaches the line, and then decrease as the

Page 87: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

70

vector moves past the line. The first derivative of this transition in intensity is represented by an

incline where the slope of the rate of change of intensity values is increasing, and a decline

where the slope of the rate of change of intensity values is decreasing.

Edge detection using the rate of change of the first derivative involves the detection of

zero crossings of second derivatives. The second derivative is represented by an incline where

the slope of the rate of change of the first derivative is increasing, and a decline where the slope

of the rate of change of the first derivative is decreasing. The maxima of the second derivative

indicate the point at which the slope of the rate of change of the first derivative begins to

decrease, while the minima of the second derivative indicate the point at which the slope of the

rate of change of the first derivative begins to increase. Zero crossings of second derivatives

occur when the slope of the rate of change of the first derivative is zero and coincide with the

location of the first derivative maxima and minima).

The first derivative indicates the gradient of the original image or the slope of the rate of

change in intensity values along any one point in the vector. At zero, this indicates no change in

intensity values (corresponding to homogenous areas (non-edges) where adjacent pixels are

roughly equal in value). First derivative values move away from zero (either positively or

negatively) when the difference in intensity value between adjacent pixels increases. The greater

the increase in intensity values between adjacent pixels, the higher the first derivative. The

greater the decrease in intensity values between adjacent pixels, the more negative the first

derivative. The maxima of the first derivative indicate the highest gradient magnitude along a

transition from low to high intensity values (i.e., the point along a series of pixels increasing in

intensity where the greatest increase in intensity between a pixel and subsequent pixel occurs).

Page 88: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

71

The minima indicate the highest gradient magnitude in a transition from high to low values (i.e.,

the point along a series of pixels decreasing in intensity where the greatest decrease in intensity

between a pixel and subsequent pixel occurs). We can classify these points as edges; gradient

maxima as localization of the sharpest increase in a point of transition between low and high

intensity, and minima as the localization of the sharpest decrease in a point of transition between

high and low intensity.

The gradient extrema (maxima and minima) are extracted from greyscale outputs of first

derivative filters by gradient thresholding. This method transforms a greyscale image with

intensity values corresponding to gradient magnitudes into a binary image--with gradient

magnitudes exceeding the threshold producing an output of one, and all other pixel values

producing an output of zero. This provides a tool for customizing the strength of edge tolerated.

If only strong edges are desired in the binary output, a high gradient magnitude threshold will

only identify gradient magnitudes which extend far above or below zero (i.e., which represent a

strong increase or decrease between adjacent pixel intensity values). Conversely, a lower

gradient magnitude threshold may identify gradient magnitudes close to zero (i.e., which

represent a small increase or decrease between adjacent pixel intensity values). The problem with

gradient thresholding is that it does not specifically identify the single pixel location

corresponding to the maxima or minima. In a single transition from low to high intensity values

in the original image, for example, all resulting gradient magnitudes along the transition which

exceed the threshold will produce a 1 in the binary image. If the maxima exceed the threshold,

the edge will be identified, but all additional pixels along the transition with a gradient

magnitude above the threshold will also be identified. Consequently, gradient thresholding can

Page 89: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

72

produce edge lines which are multiple pixels in thickness and require further thinning steps

[134]. This method is routinely applied to outputs of Sobel and Prewitt operators, 2 of the most

commonly used edge detection methods.

Figure 2.16: Prewitt and Sobel edge detection on B-scan of human

kidney. 3x3 (a) Prewitt and (b) Sobel methods applied to a smoothed

(Gaussian, sigma=3) B-scan of human kidney. The output of each edge

detection method is overlaid in green over the original image.

2.4.4.2 Canny Edge Detection

Canny edge detection is the most widely used and arguably most versatile edge detector.

Canny edge detection begins with application of a Gaussian smoothing filter. Window size of the

Gaussian filter is customizable but carries with it the tradeoff of noise suppression for

localization. Larger Gaussian filters will remove more noise, reducing false detection of spurious

edges. Smaller Gaussian filters will leave more noise but reduce the ability of the edge detector

to precisely locate edge points.

Page 90: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

73

Following smoothing operations, the Canny edge detector maps gradient magnitudes and

orientations using a first derivative operator (like Sobel or Prewitt). 4 partial first derivative

kernels (horizontal, vertical, and diagonal in each direction) are convolved over the smoothed

image, calculating gradient magnitude and determining edge direction at each pixel location. To

avoid the multi-pixel edge effect of gradient thresholding, the Canny method then employs non-

maxima suppression (i.e., setting to zero all gradient magnitudes that are not local maxima (or

minima)) for edge thinning. Identification of these maxima involves comparison of the gradient

magnitude of a specified pixel to the gradient magnitudes of the previous and subsequent pixel

along the direction of the edge (perpendicular to the length of the edge). If the target pixel does

not contain a gradient magnitude greater than the previous or subsequent pixel, it will be

suppressed.

To remove gradient extrema of low gradient magnitudes (spurious edges), the Canny

edge detector employs hysteresis thresholding. Hysteresis thresholding involves the use of two

threshold limits-upper and lower. Any identified extrema with a gradient magnitude above the

upper threshold is considered a definite edge, and any extrema with a gradient magnitude below

the lower threshold is immediately excluded from consideration. Extrema with gradient

magnitudes between the lower and upper thresholds are identified as edges only if they bridge

the gap between two edges above the upper threshold (a process known as edge-linking). This

process preserves edge continuity along lengths of edge which may vary in edge strength [143].

Page 91: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

74

Figure 2.17: Canny edge detection on B-scan of human kidney. 3x3

Canny method applied to a smoothed (Gaussian, sigma=3) B-scan of human

kidney. The output is overlaid in green over the original image.

2.4.4.3 Second Order Derivative Edge Detection

The second order derivative reflects the rate of change of the gradient. At zero, this

indicates no change in the gradient rate (i.e., intensity values are either constant, increasing at a

constant rate, or decreasing at a constant rate). Second order derivatives move away from zero

(either positively or negatively) as the rate of change of the gradient increases or decreases

respectively. The greater the increase or decrease in intensity values between adjacent pixels

relative to the increase or decrease in the previous set of pixels, the higher the second order

derivative. The smaller the increase or decrease in intensity values between adjacent pixels

relative to the increase or decrease in the previous set of pixels, the more negative the second

order derivative. The zero crossing of a second order derivative occurs at transitions from

increasing gradient rate to decreasing gradient rate. In the case of a zero-crossing moving from

maxima towards minima: the zero crossing reflects a point in the original image where there is a

transition from intensity values which are increasing in each successive pixel by larger and larger

increments, to where pixel values are increasing by smaller and smaller increments. In the case

of a zero crossing moving from minima towards maxima, the zero crossing reflects a point in the

original image where there is a transition from intensity values which are decreasing in each

Page 92: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

75

successive pixel by smaller and smaller increments, to where pixel values are decreasing by

larger and larger increments. We can similarly classify these points as edges; zero-crossings as

localization of the points of tapering off of gradient rate increase.

Second order zero-crossings correspond in location to first derivative minima and

maxima, but offer the benefit of precision and circumvention of superfluous edge thickness by

gradient thresholding. Zero-crossings can be extracted from greyscale outputs of second

derivative filters by isolation of zero-values flanked by positive and negative values. This

method produces a binary image with zero-crossings producing an output of one, and all other

pixel values producing an output of zero. It is advantageous over gradient thresholding of first

derivatives in that it produces single pixel edges and no edge thinning steps are required. Second

order edge detection, however, is indiscriminate in edge selection. The zero-crossing of the

second derivative offers a location of an edge without quantification of edge strength, prohibiting

thresholding steps to eliminate weak edges. To circumvent this dilemma, first and second order

derivative edge detectors are often used in tandem. First order derivatives provide gradient

magnitude and gradient orientation, while the zero-crossings of second derivatives offer precise

locations of gradient extrema [134].

Page 93: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

76

Figure 2.18: Laplacian edge detection on B-scan of human kidney. 3x3

Laplacian method applied to a smoothed (Gaussian, sigma=3) B-scan of

human kidney. The output is overlaid in green over the original image.

2.4.5 Region Segmentation

In image processing, segmentation involves the subdivision of an image into discrete

parts based on similarity or discontinuity in pixel values or local patterns. In medical imaging,

region segmentation can be used to highlight pathology (e.g., tumor or lesion), or to study

normal anatomical structure. Segmentation of pathology may inform diagnosis, help in tracking

of disease progression, and aid in treatment planning and dosing. The wide array of imaging

modalities in use, variations within each modality depending on manufacturer, litany of

structures imaged, and natural and pathological heterogeneity of anatomical structures makes the

choice of region segmentation technique uniquely dependent on the task at hand.

2.4.5.1 Global Thresholding

The simplest and most intuitive form of region segmentation is global thresholding.

Global thresholding involves the installment of a single threshold limit, whereby all pixel values

above the set limit produce a 1 value in the binary output image and all other pixel values

produce a 0 value output. This method performs well when confronted with an image with a

bimodal distribution of intensity values, where one mode represents the object of interest (or

Page 94: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

77

foreground) and the other mode represents background. A threshold bisecting these two modes

should perform well in segmenting the object of interest, assuming minimal overlap between

modes. The appropriate threshold can be determined empirically, and can be expected to

continue to perform well if the foreground and background intensity distributions remain the

same in all subsequent images. In practice, however, this is rarely the case. Efficient global

thresholding, therefore, generally relies on automatic selection of an optimal threshold for each

image processed [134].

2.4.5.2 Otsu’s Method

Otsu’s thresholding, a clustering method, is an example of one of the more popular forms

of global thresholding. In Otsu’s method, we iterate through all possible thresholds, and select

the threshold which minimizes intra-class variance, or maximizes inter-class variance (equivalent

effects). Otsu’s method moves along each point in the x-axis of a histogram representing the

distribution of intensity values in an image. At each point, a global threshold is set, dividing the

histogram into 2 classes (above and below the threshold). Each class can be considered a cluster

of values. The goal of Otsu thresholding is to define the threshold which produces clusters that

are as tightly clustered as possible (minimizing intra-class variance) and as far as possible from

the other cluster (maximizing inter-class variance). Adjusting the threshold in one direction of

the other will reduce the spread of one cluster, while increasing the spread of the other. The

intra-class variance is calculated at each point as the weighted sum of the variance of each of the

two classes (the combined spread of both clusters), with weights defined as class probability

computed from the histogram binning and variance defined as the squared deviation from the

Page 95: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

78

mean. The point of minimal intra-class variance gives the threshold which maximizes the

tightness of clustering and, conversely, maximizes the separation between clusters [144].

Figure 2.19: Otsu thresholding applied to optical coherence tomography

B-scan of human kidney. (a) Histogram of initial intensity values plotted

across 0-255 intensity range (b) Histogram of initial intensity values plotted

across 0-255 intensity range with red vertical line indicating the Otsu defined

threshold (c) Binary output of Otsu thresholding.

2.4.5.3 Local Thresholding

Global thresholding fails when illumination or signal in an image is unevenly distributed.

In cases where the intensity of the region of interest and/or of the background varies throughout

an image, a single global threshold will either overestimate or underestimate the foreground in

different parts of the image. In these cases, local thresholding is the preferred strategy. Local

thresholding is similar in theory to global thresholding, but involves partitioning the image into

tiles and determining unique thresholds depending on the intensity distribution within each tile.

Page 96: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

79

Thresholds within each window can be set by any specified method, including Otsu’s. Similar to

window sizes of low and high-pass filters, tile size can be user-defined and tends to impact

output in a tradeoff fashion. A smaller tile size will produce a threshold based on the intensity

distribution of a limited sampling area, so may falsely identify noise in homogenous areas.

Larger tile sizes limit identification of noise, but move closer towards the problems associated

with global threholding. In addition, as tiles increase in size, the boundaries of segmented

regions tend to increase in tandem. Selection of a tile below or above an optimal size may

produce underestimation or overestimation, respectively, of segmented regions of interest [134],

[145].

Figure 2.20: Local adaptive thresholding applied to optical coherence

tomography B-scan of human kidney. (a) Binary output of local adaptive

thresholding with a 15x15 window.

2.4.6 Active Contour (Snakes)

Active contour models, or “snakes”, are a common strategy employed in region and edge

detection when the approximate shape of the region of interest is known beforehand. Snakes can

be considered a form of curve-level segmentation, as opposed to pixel-level segmentation seen in

thresholding. Snakes are essentially a circular or linear shape, placed by a user near an object or

edge of interest, which expands or contracts to fit the region or edge. They move towards high

Page 97: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

80

gradient magnitudes, but attempt to fit image contours smoothly rather than in a jagged pixel-

wise configuration which may traverse more high gradient values.

In the context of active contour models, a snake can be considered a deformable spline (a

function defined by a series of polynomials) constrained by a set of rules governing its flexibility

and affinity towards edges. Snakes operate iteratively to conform to image edges while achieving

a minimal energy state. Energy in active contour models can be decomposed into internal and

external forces. Internal forces represent a user-defined assumption of curve flexibility. This is

independent of image features, and defines only how tolerant the snake will be in terms of

overall curvature and local blebbing. A low internal energy state reflects a relaxed shape (e.g.,

circular form of a rubber band at rest), while a high internal energy state reflects the amount of

energy required to maintain the distortion of the native shape. External forces depend on image

intensity (e.g., image gradient represented by first order derivatives), and are perceived

negatively when summing energy during minimization steps (i.e., E = Einternal – Eexternal). A low

external energy state reflects a curve fit which overlaps with high gradient magnitudes, while a

high external energy state reflects a curve fit which overlaps with low gradient magnitudes.

At each iteration, snakes tighten their fit to the contours of the region or edge of interest

by following the current of the gradient vector flow; a gradient vector field, representing gradient

strengths and orientations as vectors, pulls portions of the snake towards the direction of higher

gradient magnitudes. At each step, the energy of the internal and external energy are summed

until either the minimum energy is reached (further iterations do not reduce total energy) or until

the user-defined iteration limit is reached. The resulting minimum energy snake should fit the

contours of the targeted region or edge as tightly as possible without deviating too far from a low

Page 98: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

81

energy shape. In design of an active contour model, there is an obvious tradeoff between

tightness of fit and smoothness. The more inflexible the snake, the less tightly the snake will be

fit to edges. The more flexible the snake, the more the resulting segmentation will deviate from

the assumed shape. Consequently, the user must balance the weight of internal and external

energies depending on the suspected irregularity of the target shape or edge. This segmentation

method is effective and relatively immune to noise, but is computationally exhaustive and

requires user input for each region or edge of interest [146].

Figure 2.21: Active contour model applied to optical coherence

tomography B-scan of human kidney. Green indicates the active contour

segmentation of the renal capsule following 200 iterations. Pink indicates the

active contour segmentation of the kidney cortex following 200 iterations.

2.4.7 Graph Cuts

Graph cuts are another popular segmentation method relying on the principal of energy

minimization used by active contour models. Graph cut segmentation borrows tools from graph

theory to segment foreground from background in an image by assigning energy (or weight) to

pixel connections and finding the least expensive cut of all connections to partition the image. In

graph cut segmentation, a 2-dimensional or 3-dimensional image is treated as a graph, with each

pixel corresponding to a vertices (or node) and each node connected to adjacent nodes by links

(or edges). Edges are generally assigned weights based on differences in intensity (gradient), but

Page 99: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

82

some graph cut methods include color and texture components in the weighting scheme. Weights

are intended to promote adhesion of pixels with similar features; edges between nodes with

similar intensity receive greater weights, increasing the cost if these edges were to be cut.

Additional sets of edges with corresponding weights are defined between each node in

the image and two user-specified nodes belonging to the foreground and background (“source”

and “sink” respectively). After the user identifies the source and sink in the original image, edge

weights between each of these nodes and all other nodes in the image are calculated based on the

probability that each node belongs to the foreground or background. This probability value is

based primarily on intensity similarity and distance between nodes. Nodes closer to and

exhibiting similar intensity profiles as the source node, for example, receive a strongly weighted

edge link to the source and a lesser weighted edge link to the sink.

Graph cuts segment foreground from background by seeking out the path through node

edges which isolate the foreground while cutting through the minimum sum edge weight. To

achieve this segmentation, cuts must occur both between edges connecting adjacent nodes and

between the edges connecting nodes to either the source or sink. The cumulative weight of all

edge cuts produces the total energy of the graph cut, and the minimum energy cut is selected for

segmentation. On its own, this method would isolate very small regions of background or

foreground (the sum weight of the cut is proportional to the total number of edges in the cut). To

circumvent this bias, graph cuts utilize a “normalized cut” which penalizes weight sums from

large or small segmentations. Graph cut segmentation is widely used, particularly in

segmentation of 3-dimensional volumes. Graph cuts are computationally efficient relative to

Page 100: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

83

active contour models, but again come with caveat that user intervention is required (in selection

of the source and sink nodes) [147]–[149].

Figure 2.22: Graph cut applied to optical coherence tomography B-scan

of human kidney. Green indicates the graph cut segmentation of the renal

capsule. Pink indicates the graph cut segmentation of the kidney cortex. Blue

indicates the graph cut segmentation of PCT lumen.

2.4.8 Segmentation in Optical Coherence Tomography

The segmentation of OCT images carries with it unique medical image processing

hurdles. OCT images are riddled with speckle-noise, granular noise inherent to OCT which can

obscure image quality and deteriorate edge strength. Additionally, since OCT signal is derived

from the light absorption and scattering characteristics of tissue, intensity of the signal will

decline irrespective of tissue properties as penetration into the tissue increases (as less and less

light reaches the tissue due to absorption by superficial tissue). This effect forms a downward

intensity gradient along each A-scan, even in homogenous tissue. Similarly, imaging depth of

OCT is considerably reduced relative to most other imaging modalities. To acquire the maximum

amount of data, the field of view of most OCT imaging protocols extends beyond the maximum

penetrating depth of the OCT system. Consequently, most OCT images include a decreasing

intensity gradient corresponding to a diminishing signal, and a point in each A-scan where signal

Page 101: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

84

is replaced entirely by noise in all additional pixels. Finally, high absorption of light by

hemoglobin can induce optical shadowing beneath blood vessels [150].

OCT retinal imaging is routinely captured in 3D volumes, producing massive amounts of

imaging data. To enable ophthalmologists to process this data in a timely and reliable manner,

segmentation algorithms have begun to receive considerable attention. A flood of segmentation

algorithms has emerged over the last 20 years to process retinal OCT imaging data. To deal with

speckle-noise, the majority of algorithms include pre-processing with some form of median filter

or an anisotropic low-pass filter (weighted to smooth horizontally without minimal corruption

vertically of retinal layers) [151]. The first OCT segmentation algorithm, proposed by Hee et al.,

was published in 1995. This method utilized a custom 1D edge detection filter, which iterated

over each A-scan and isolated the two strongest edges [152]. Yang et al. introduced a 2D

approach, utilizing a custom Canny edge detector to detect 9 retinal layers [153]. Since then,

active contour models have emerged as the leading choice for 2D edge detection methods [154],

[155]. Active contour models have proven superior to other 2D edge detectors in both accuracy

and handling of noise. Adaptive thresholding was introduced by Duan et al. for the segmentation

of vessel lumens in OCT scans [156]. Graph cuts have gained popularity, both in 2D and 3D, and

have found use in both layer segmentation and region segmentation of macular cysts [157]–

[159]. In recent years, attention has shifted to OCT retinal segmentation algorithms which utilize

some form of machine learning.

Page 102: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

85

2.5 Summary

OCT occupies a special niche on the spectrum of non-invasive imaging modalities. OCT

offers very high resolution compared to all other non-invasive modalities, with the exception of

confocal microscopy. While confocal microscopy has been used successfully to image non-

human kidneys, its poor penetration disqualifies it from use in imaging human kidneys which

contain a relatively thick renal capsule. MRI has been attempted in assessment of ischemic

damage and post-transplant pathology, however, its resolution is too poor to discern any fine

tissue structure and its utility prior to transplant is largely unexplored. CT and ultrasound have

been employed in guiding post-transplant kidney biopsies, however, these again lack the

resolution to discern the kind of fine tissue structure which are evident in traditional biopsies and

established as predictive of post-transplant function. OCT provides penetration adequate to

visualize into the kidney cortex and enough resolving power to be able to discriminate the kind

of anatomical features which are routinely evaluated in kidney transplant biopsies.

Contrast stretching and histogram equalization both prove useful in pre-processing of OCT

images of human kidneys. These contrast enhancement techniques provide normalization of

intensities across a wide range of images, and compensate for the signal dissipation at increasing

imaging depths. Gaussian smoothing is similarly necessary as a pre-processing step for analysis

of OCT imaging of human kidneys. Gaussian smoothing operations proved effective in removal

of speckle noise, and smoothing of discontinuities in edge or region intensity. This enabled

uncorrupted edge and region detection steps.

Page 103: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

86

Due to the large size of OCT image volumes, more sophisticated image segmentation

methods which are common to OCT analysis (active contour models, graph cuts) were not

feasible. Both methods generally require substantial user input, which would be exhaustive in

larger image sets and also introduces the possibility of user bias. Similarly, both methods (active

contour models especially) are computationally exhaustive and so would take considerable time

to process imaging data, limiting their practical use in clinical decision making. Canny edge

detection and adaptive thresholding are relatively computationally inexpensive methods of layer

and region segmentation, respectively. Canny edge detection is robust and optimally suited for

segmentation of OCT kidney imaging data where stretches of edge lengths vary widely in

strength and encounter numerous discontinuities. Other first and second derivative edge detectors

lack the precision and consistency in the face of imaging artifacts, discontinuous edges, and

noisy features. Adaptive thresholding is similarly robust, and effective across the wide range of

image gradients and varying intensity profiles which present in OCT kidney imaging data.

Page 104: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

87

CHAPTER 3: IMAGE CAPTURE AND ANALYSIS PIPELINE

3.1 Introduction

We open this chapter with a survey of the OCT system specifications, and go on to

explain the motivation and process behind selection of OCT system settings. Next, we introduce

the protocol exercised for imaging human donor kidney in the operating room prior to and

following transplantation. We then introduce the framework for manual segmentation of kidney

features in OCT B-scans, and go on to navigate the automatic segmentation pipeline highlighting

pre-processing, layer-segmentation, and region segmentation steps. Next, we assess performance

of the segmentation algorithm as compared to segmentation by manual raters. Finally, we

introduce measurement methods for segmented kidney features. We conclude this chapter with a

brief discussion on the advantages of high speed segmentation and quantification of kidney

structures in OCT for assessing graft viability.

3.2 System Specifications

3.2.1 930 vs. 1325nm

The majority of scans in this study (92.5%) were performed with a 1325 nm center

wavelength SD-OCT imaging system (Telesto-II, Thorlabs Inc.), with an incident power of 2.5

mW. The Telesto OCT system was equipped with a 36 mm focal length (LSM03, Thorlabs Inc.)

objective, providing a lateral resolution of 13 µm and an axial resolution of 5.5 µm in air. The

remaining scans were performed with a 930 nm center wavelength SD-OCT imaging system

(Ganymede-II, Thorlabs, Inc.) equipped with a LK20 objective, a lateral resolution of 8µm and

Page 105: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

88

an axial resolution of 3µm in air. Scans were captured at a rate of 28 kHz (video capture rate), a

sensitivity of 103dB, and an incident power of approximately 2.5µW. OCT scans were captured

using Thorlabs native imaging software, ThorImage 4.1, operated on a 3.6 gHz Dell desktop with

a dual CPU intel processor and 16GB RAM. MATLAB image processing was also conducted on

this desktop.

The higher wavelength Ganymede system provides higher resolution than the Telesto

system at the cost of reduced penetration (Figure 3.1). The Ganymede and Telesto systems were

used in tandem in the initial protocol-optimization phase of this study to determine the

practicality and effectiveness of each system in the imaging of human kidneys. To select the

optimal system for the duration of the study, the cost and benefit of resolution and penetration

differences between the two systems were investigated.

Figure 3.1: 930nm and 1325nm B-scans of the human kidney. (a)

Representative B-scan performed by the 930nm Ganymede imaging system

on an ex-vivo human kidney prior to transplant. (b) Representative B-scan

captured minutes later, performed by the 1325nm Telesto II imaging system

on the same kidney as is represented in (a).

Page 106: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

89

Manual segmentation was performed independently by 4 raters on a randomly selected

subset of images from kidneys imaged by both the Ganymede and Telesto systems. Raters were

asked to segment the interface between the renal capsule and the cortex (green in Figure 3.2), and

to segment the depth at which the signal to noise ratio became insufficient to discern any

anatomical landmarks (i.e. lumen of PCTs, glomeruli, blood vessels) (blue in Figure 3.2).

Figure 3.2: Manually segmented 930nm and 1325nm B-scans of the

human kidney. Representative B-scans from the 930nm Ganymede (a) and

1325nm Telesto II (b) with the renal-capsule interface manually segmented in

green and the lowest point of perceptible signal manually segmented in blue.

The distance between the capsule-cortex interface and the signal-noise cutoff was

measured for each A-scan and averaged across all A-scans in each B-scan to produce a measure

of average penetration into the cortex for each image. Penetration depths were averaged for each

kidney imaged by both Thorlabs systems. The Telesto-II system achieved a mean penetration

into the cortex of 278.91µm with a standard deviation of (±) 61µm while the Ganymede-II

system achieved a penetration of 165.66µm ± 38µm into the cortex. The 40.6% reduction in

penetration into the kidney cortex by the Ganymede system was considered a prohibitive

limitation; a wide range of renal capsule and adipose thicknesses between and within kidneys

Page 107: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

90

suggested that the Ganymede system may fail to penetrate into the cortex of kidneys with thicker

capsules or higher degrees of adipose. The improved resolution of the Ganymede system

revealed no additional anatomical features (capillaries, interstitial space) and did not demonstrate

sufficient improvement on the resolution of the PCTs to outweigh the limited penetration.

3.2.2 2D vs. 3D (OCT-B versus OCT-C cans)

While the capture of 3D volumes is within the capabilities of both the Telesto and

Ganymede systems, this proved impractical due to the motion artifact suffered by handheld

scanning. Multiple speeds were initially attempted during the protocol-optimization phase of the

study, with lower speeds providing higher resolution and higher speeds offering the appeal of

potential 3D volume capture. 28, 48, and 76 kHz scans were performed on a preserved kidney by

a lab member operating a handheld probe. Resolution of the 76 kHz scans proved to be too poor

to be able to accurately discriminate PCT features. The 48 kHz mode provided considerably

better resolution than the 76 kHz, and may have been sufficient quality for analysis, but was still

limited in its ability to produce reliable 3D volumes during handheld imaging (i.e., motion

artifact induced by hand motion distorted the 3D volumes). 28 kHz was ultimately selected as the

final imaging speed as 3D capture did not appear to be a feasible option, and 28 kHz provided

the highest resolution in B-scans.

3.2.3 Field of View

Field of view (FOV) was similarly modified throughout the protocol-optimization portion

of the study. Initial settings included a large FOV, spanning >8mm in the transverse direction (x-

axis) and nearly 3mm in the axial direction (z-axis). These settings were chosen to maximize the

Page 108: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

91

area of kidney imaged but often occupied so much random-access memory (RAM) as to crash

the data acquisition software. Similarly, successful capture at these settings resulted in image file

sizes (>5 GB) which could not be opened without a custom Python addition to the Thorlabs

export code, and which decelerated processing substantially. The FOV was reduced to a 4.9mm

x-axis and 1.9mm z-axis (after adjusting for a refractive index of 1.3). This reduced RAM usage

and file size, preventing further system crashes and enabling practical use of the data files.

3.2.4 Averaging

A-scan averaging was similarly increased to 2 to halve the capture rate and similarly

reduce RAM usage and file size (B-scan averaging was avoided as the motion tied to hand-held

imaging exceeded the speed of the mechanical movement between B-scans and so effected

image quality negatively). A-scan averaging had negligible effect on image quality.

3.2.5 Scale

The minimum provided scale of 2.73µm/pixel was selected for the X-axis to maximize

resolution in this dimension. 3.54µm/pixel was selected as the Z-scale in air, equating to

2.73µm/pixel after adjusting for the refractive index of 1.3 (the 1.3 refractive index was an

estimate provided early in the study by collaborators, post-study evaluation determined the

refractive index of formalin preserved kidney tissue to be closer to 1.4). The synchronization of

the X and Z scales prior to capture enabled export to an analyzable format, without interpolation

or any change to the raw data producing feature loss.

Page 109: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

92

3.3 Operating Room Imaging Protocol

3.3.1 Basic Setup

A technician in sterile surgical attire operated a handheld scanner, draped in a sterile

sleeve with a layer of sterile 3M Tegaderm Transparent Adherent Film Dressing affixed to the

focal spacer of the OCT probe (Figure 3.3). A second technician in non-sterile scrubs operated

the ThorImage software, and maneuvered the cart holding the OCT system when necessary.

Image files were saved in the native ThorImage “.oct” format. Prior to processing, images from

original “.oct” files were opened as 32 bit floating tiffs in MATLAB along with accompanying

meta data (x and z scales, FOV, wavelength).

Page 110: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

93

Figure 3.3: Technician in sterile surgical attire operates a probe draped

in a sterile sleeve to image a kidney ex-vivo (flushed with preservation

solution and resting in a bowl of ice on the OR back-table).

The renal capsule (Figure 3.4a, b), and adipose tissue present on the kidney surface (Figure

3.4c, d) varied widely between kidnys and between regions on the same kidney. Absorption of the

OCT light by these tissues attenuated the signal, reducing penetration into the kidney cortex

relative to capsule or adipose thickness. This made a global imaging protocol infeasible; thicker

portions of the capsule and areas of high adipose impeded OCT penetration into the cortex.

Figure 3.4: Cropped portions of B-scans of donor kidneys with varying

capsule and cortex thickness. B-scans of kidneys with a thin renal capsule

(a), thick renal capsule (b), small degree of adipose present above the renal

Page 111: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

94

capsule (c) and higher degree of adipose present (d). Yellow arrows indicate

the thickness of the renal capsule (a, b) and adipose tissue (c, d).

To quantify the variation and range in in capsule and adipose thickness, manual raters were

asked to measure capsule and adipose thickness at a randomized location on the X-axis in a

randomized set of 1,000 B-scans. The renal capsule ranged in thickness from 44 µm to just over 1

mm with an average thickness of 189.5 ± 108.7 µm. Kidneys often had little or no adipose present

on their surface but in some instances had adipose that exceeded the penetration depth of the OCT

system (>1.9mm). Average adipose thickness across all 1,000 B-scans was 67.2 ± 90.6 µm.

Consequently, technicians were instructed to survey the kidney and image regions where

adipose and the renal capsule both appeared thin (in the preserved kidney, thicker portions of the

capsule and adipose appear grey and white respectively while areas with thin capsule and

minimal adipose appear redder). In these regions, the OCT imaging was more likely to penetrate

further into the kidney cortex and increase the quantifiable area. Within these regions, the

technician was instructed to search for locations where the tubular lumen appeared most dilated

and the number of visible tubular lumen was highest. When a location matching these criteria

had been found, the technician was instructed to continue imaging in that same approximate

location.

The technician was instructed to attempt to keep the interface between the renal capsule

and cortex horizontal across the x-axis of the resulting B-scan, and at approximately 0.5mm

below the top of the B-scan. This method was intended to ensure that the cortex was placed in

the area of best focus with the given OCT settings. Live video-rate capture was advantageous

given this protocol as it allowed for continuous capture during the survey process, ensuring the

Page 112: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

95

maximum number of images was obtained. The capsule-cortex interface, however, was often

significantly above or below the 0.5mm target, resulting in out of focus cortex and strong

variations in the intensity of the cortex (the cortex was brighter and tubule lumen were less

apparent when the capsule-cortex interface was significantly above the 0.5 mm mark; the cortex

was darker and tubule lumen again less apparent when the capsule-cortex interface was

significantly below the 0.5 mm mark).

While technicians were able to maneuver the ex-vivo kidney prior to transplant and

position the OCT probe against the kidney in any location, following transplant and reperfusion

the in-vivo kidney was immobile and buried within the body cavity of the transplant recipient.

Depending on the placement of the kidney and the size of the body cavity, this limited the area

which the probe was able to contact from a few square inches to approximately 1/3rd of the

kidney surface area.

As an undesired consequence of the technicians’ instructions and the 2D imaging feed,

there was no indication of where on the kidney each image was captured, how many regions of

the kidney were imaged, or how many images were redundant/duplicates of the same region.

Regions imaged following reperfusion could not be paired with regions captured prior to

transplant for direct comparison. Video-rate capture of the survey process similarly resulted in a

high number of images captured during manual manipulation of the probe. These included

images where the probe was not in contact with kidney, or where the probe was only in partial

contact with the kidney. The kidney often only partially spanned the x-axis of resulting B-scans,

and was often diagonally oriented along the x-axis. Many images contained large amounts of

adipose and occasionally only adipose.

Page 113: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

96

3.3.2 Timing of Pre and Post Scans, Multiple Scans

Pre-implantation image sets were obtained ex-vivo following back-table preparation of

the kidney graft and again in-vivo immediately (13 ± 4 minutes) following reperfusion of the

transplanted kidney. Machine perfused kidneys were removed from the perfusion pumps prior to

imaging (i.e. were static at the time image sets were obtained). The time between removal from

perfusion pumps and imaging was not recorded but may have been an important variable; PCT

morphology may change following removal from the pump in a time-dependent manner.

Removal from the perfusion pump was generally conducted immediately prior to preparation of

the kidney graft, suggesting the time between removal and imaging was minimal (involving

primarily the time required to prepare the kidney, roughly ~30min depending on the surgeon,

anatomy of the graft, and the effect of the procurement process on the graft).

Time of access was an additional variable component of the protocol. Prior to transplant,

technicians were granted access to the kidney from the period following the transplant surgeons’

“prep-work” to “off-ice” time when the kidney was removed from the cold saline bath and began

placement into the recipient. For cadaver transplants, generally only a single a transplant team

was involved and the time between “prep-work” and “off-ice” time was extended by the

transplant team’s preparation of the recipient. This time window was generally sufficient for

even a lengthy imaging protocol. For live donor transplants, generally two transplant teams were

active with the donor and recipient in adjacent operating rooms. To minimize ischemic time,

procurement of the kidney graft from the donor is timed to occur only when the recipient is ready

to receive the kidney. In these cases, the window between “prep-work” and “off-ice” time was

minimal and technicians had 1 to 2 minutes to perform pre-implantation imaging.

Page 114: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

97

Following reperfusion, some bleeding from the re-anastomosed vasculature was

commonplace. Surgeons were required to address this hastily and before allowing technicians

access to the kidney for imaging. This introduced another variable period: the time between

reperfusion and post-reperfusion imaging. In rabbit studies of ischemic PCTs, swelling

dissipated rapidly following reperfusion [160]. This, however, followed only a short period of

warm ischemia. It is unclear how quickly swelling of the PCTs would dissipate in a human

model, following prolonged cold ischemia in preservation solution. It is possible that the

dissipation of swelling under these circumstances is a more gradual or heterogeneous in which

case the time period between reperfusion and imaging may be an important variable.

Post-reperfusion imaging time was limited in both cadaver and live donor transplants.

While the more damaging period of ischemia had passed by this point, imaging following

reperfusion directly interfered with the progress of the surgery. Technicians were limited to

approximately 1 to 2 minutes for post-reperfusion imaging. Resulting image sets under the

established protocol contained a variable number of images, ranging from roughly 200 to nearly

1,500 images for each pre-implantation and post-reperfusion scan.

3.4 Manual Segmentation

Images were analyzed manually to provide a standard to evaluate performance of the

automatic segmentation and also to produce thresholds for inclusion/exclusion of automatically

segmented PCT lumen. Manual categorization and segmentation of images was performed in

ImageJ (NIH) by 4 trained raters. Pre-implantation and post-reperfusion image sets from the first

150 patients were anonymized, randomly divided into 4 groups, and split between the 4 trained

Page 115: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

98

raters. Following manual analysis, 20 percent of the manually segmented images were reassigned

to different raters to produce measures of inter-rater variation (Figure 3.5).

Manual segmentation was performed on 5 randomly selected images from each image set.

Raters segmented the interface between the renal capsule and the cortex (upper red and blue lines

in Figure 3.5). Raters also segmented the full volume of quantifiable cortex (the area of cortex

beneath the capsule where the signal appeared sufficient to discriminate anatomical features) (area

between upper and lower red and blue lines in Figure 3.5). Raters then segmented all regions which

appeared to be cross-sections of PCT lumen, using the ImageJ “Versatile Wand” plugin [161] (red

and blue selections in Figure 3.5 with cyan indicating overlap). If a randomly selected image

contained no quantifiable cortex, the image was skipped and the reason for exclusion was tallied

as either “empty” with no contact between the probe and kidney, “high reflection,” or “high

adipose.”

Figure 3.5: Inter-rater segmentation overlay. Representative B-scan

independently segmented by 2 manual raters. Selections by the first rater

are indicated in red while selections by the second rater are indicated in blue.

Cyan indicates an overlap in selection by both raters

Manual raters were instructed to avoid segmenting structures which they could not

reliably determine to be the lumen of PCTs (i.e., blood vessels, cysts, glomeruli). This is

Page 116: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

99

admittedly a challenge, particularly with respect to blood vessels, when raters are limited to a 2D

view of the kidney cortex. In the attached en-face view of a 3D reconstruction of an OCT scan

performed on a preserved human kidney, vessel lumen (labeled in red in Figure 3.6) and PCT

lumen can be more reliably distinguished.

Figure 3.6: 3D and 2D representation of optical coherence tomography

imaging of vessels in the human kidney. (a) Reconstruction of a 3D OCT

scan performed on a preserved human kidney, with suspected vessels

highlighted in red (b) OCT B-scan with suspected vessels (middle, right) (c)

OCT B-scan with suspected vessel (left).

PCT features were defined, by a kidney anatomy and histology expert involved in the

study (Dr. Peter Andrews), as having high tortuosity and a lumen diameter within an

approximate range. “Straight and elongated” were features used not independently, but in

combination with lumen diameter to discriminate blood vessels. While some cross-sections of

PCT lumen create straight and elongated features, these are most often shorter or curved due to

their high tortuosity and are lower in luminal diameter relative to vessel lumen.

Page 117: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

100

3.5 Automatic Segmentation

3.5.1 Automatic Analysis Pipeline

Automatic segmentation was executed in MATLAB R2017b (Mathworks, Inc., Natick,

MA, USA). To remove user bias and to improve feasibility of clinical application, automatic

segmentation and analysis was performed on the original full 2D video image sets and not

manually selected subsets of images.

In addition to the issues in image sets incurred by the imaging protocol, speckle noise

intrinsic to OCT coupled with imaging artifacts derived from anatomical heterogeneity and OCT

system issues confounded automatic segmentation. Most B-scans captured contained varying

degrees of reflection, which presented as bright vertical stripes. Reflection varied between

different images captured in the same image set and varied significantly between scans, likely

due to issues with the OCT system (bending of the fiber-optic cable during manipulation of the

OCT probe, or a loose or overly tight connection of the fiber-optic cable to the OCT base station)

or characteristics of the sample (light scattering characteristics of objects in the FOV may have

caused reflection). Conversely, shadowing was a routine artifact. Shadowing generally resulted

from obstruction of A-scans into the kidney cortex by an abnormal feature present on the surface

of the renal capsule. Large globules of fat often limited penetration into the cortex along

corresponding A-scans but not adjacent A-scans which were not intercepted by fat globules.

Resulting B-scans had a discontinuous renal capsule edge and heterogeneous penetration into the

cortex. Similarly, variations in adipose and the thickness of the renal capsule created an uneven

penetration into the cortex both between images and along the x-axis of each image. Sub-

capsular features with low refractive indices (vessel and PCT lumen, capsular space of Bowman,

Page 118: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

101

cysts) limited penetration less than features which produced positive space and so led to an

increase in total depth of penetration into the cortex at these locations, increasing heterogeneity

of penetration depth across the x-axis.

To circumvent the litany of imaging hurdles presented by the protocol, imaging artifacts,

and heterogeneity in kidney appearance, a pipeline was devised for processing of image sets. To

expedite analysis and prevent error, it was necessary to remove images from processing which

contained no quantifiable cortex. Features were extracted and compiled from images skipped and

marked during manual analysis. These features were utilized to classify and exclude empty, high

reflection, or high adipose images prior to performing more computationally expensive sections

of the algorithm (Figure 3.7, blue tier 2).

Following moderate pre-processing, layer segmentation was performed to segment the

interface between the renal capsule and the kidney cortex, the area of high signal quantifiable

cortex was identified and segmented, and potential lumen cross-sections were segmented and

classified as PCT lumen or non-PCT lumen cross-sections (Figure 3.7, blue tier 3). Finally, the

density of lumen area (based on the criteria assigned to technicians for regions to target and

hover over) was calculated for each image, and the single B-scan with the maximum density

value was used for analysis (Figure 3.7, blue tier 4).

Page 119: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

102

Figure 3.7: Automated image analysis pipeline.

3.5.2 Empty B-Scan Detection

While a threshold of total intensity values would be an intuitive and high-speed approach

to detection of empty B-scans, variations between empty images in background intensity, imaging

artifacts and hyper-reflectivity of Tegaderm disallowed this strategy. Empty images were therefore

identified by their average standard deviation in intensity values across the z-axis. For each B-

scan, the standard deviation of intensity values across each A-scan was taken and all A-scan

standard deviations for that B-scan averaged. This process was repeated for all images marked

during manual analysis as “empty” (Figure 3.8a), and for all images which had cortex present and

were manually segmented (Figure 3.8b). Comparison between these two groups demonstrated that

a mean A-scan standard deviation of 47 or less correlated highly with images categorized as

“empty” while a mean A-scan standard deviation above 47 correlated well with images which

contained kidney (Figure 3.8c). A standard deviation cutoff of 47 identified empty images with a

sensitivity of 83.28% and a specificity of 98.91%.

Full Image Set

Quantifiable Cortex

Capsule-Cortex Interface

Segmentation

Cortex Segmentation

Lumen Segmentation and

Classification

Final Measurements Extracted from

Highest Density B-Scan

Empty B-Scan

High Reflection

High Adipose

Page 120: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

103

Figure 3.8: Empty image detection. (a) Section of a B-scan with no kidney.

The white horizontal lines in the bottom third of the image result from the

Tegaderm. The vertical arrow represents the location of a single A-scan with a

corresponding standard deviation in intensity values of 34.3. The average of

all A-scans across the image is 37.8. (b) Section of a B-scan with kidney. The

vertical arrow represents the location of a single A-scan with a corresponding

standard deviation in intensity values of 56.1. The average of all A-scans

across the image is 58.4. (c) Histogram representing the average standard

deviation of all images manually marked as empty (blue) and all images

which contained quantifiable cortex and were manually segmented (red).

3.5.3 Reflection Detection

Bright vertical stripes due to strong reflection were a frequent imaging artifact which

interfered with several segmentation steps and in some instances rendered images impossible to

analyze (Figure 3.9a). To isolate and quantify these stripes of reflection, a horizontal filter was

applied to each image to provide an estimate of the image without the reflection (Figure 3.9b).

Reflection stripes were then defined as A-scans from the original image whose average intensity

exceeded a global threshold above the corresponding A-scans from the filtered image (Figure

3.9c). Images where reflection stripes exceeded 30% of the total number of A-scans were excluded

from analysis. A similar strategy was employed to isolate stripes of shadowing to aid in capsule

segmentation and PCT lumen selection.

Page 121: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

104

Figure 3.9: Reflection detection. (a) Original section of B-scan with high

reflection. (b) Section of the same high reflection B-scan following

application of a horizontal blurring filter. (c) Binary mask with all white

portions representing all A-scans in 6a whose mean intensity value exceeded

15 above the mean intensity value of the corresponding A-scan in 6b. (49% of

A-scans in this example qualify as reflection stripes)

3.5.4 High Adipose Detection

The amount of adipose tissue on the surface of the kidney was widely variable between

kidneys and between regions within the same kidney. Images which contained enough adipose

tissue to interfere with segmentation were infrequent in most scan sets but when present often

contained features which were falsely identified as PCT lumen (Figure 3.10). Since the intent was

to analyze regions of the highest area of PCT lumen, in image sets where the overall area of PCT

lumen was low, falsely segmented adipose was prioritized and significantly affected results.

Figure 3.10: Cortex and adipose. (a-b) Sections of B-scans with kidney

cortex and PCT lumen. (c-d). Sections of B-scans with high degrees of

adipose and circular features which may be mistaken for PCT lumen.

Page 122: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

105

In the majority of images where adipose interfered with segmentation, the segmentation of

the renal capsule was interrupted, triggering an error and exclusion. In a small but significant

number of images, the segmentation process finished uninterrupted but falsely identified adipose

as cortex. To detect these instances, a complex decision tree was generated with MATLAB’s

“Classification Learner App”. Two image sets were compiled containing falsely segmented high-

adipose images and correctly segmented images respectively. Features were extracted from the

images in each set. Features included variations in intensity values in what was interpreted as

cortex as well as the dimensions, orientation, eccentricity and size of what was interpreted as PCT

lumen. Training was performed with 10-fold cross validation and yielded a sensitivity of 97.5%

and a specificity of 98.6%.

3.5.5 Segmentation of the Renal Capsule-Kidney Cortex Interface

Segmentation of the interface between the renal capsule and cortex is a necessary step

preceding segmentation of the kidney cortex and PCT lumen. Defining this interface prevents

anything above it (capsule, adipose, background) from being falsely identified as cortex or PCT

lumen. In OCT scans, the renal capsule had consistently higher intensity than the cortex beneath

it. The shift in intensity provided a border which edge detection was able to identify.

The kidney was most often flush to the OCT probe during imaging and so in most B-scans

spanned the full x-axis. The capsule-cortex interface was therefore identified by targeting strong

edges which spanned the majority of the x-axis. Breaks in edge continuity by stripes of reflection

and shadowing were filled in with values from adjacent A-scans. Images similarly underwent a

horizontally weighted Gaussian blur (kernel=2.73x45.78 µm) to unify the length of the interface

Page 123: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

106

(Figure 3.11a). A relatively weak Canny edge detection (threshold=0.22:0.66, σ=1.83) was used

to detect and remove the Tegaderm edge and all edges above it from selection (blue line Figure

3.11b). More sensitive Canny edge detection (threshold=0.13:0.28, σ=10.99) was used to identify

edges which would correspond with the capsule-cortex interface (red and yellow lines in Figure

3.11b). From the identified edges, the lowermost detected edge which spanned at least half of the

x-axis and contained higher intensity values above the edge than below (i.e. high intensity capsule

above lower intensity cortex) was selected as the capsule-cortex interface (yellow line in Figure

3.11b).

Figure 3.11: Edge detection for renal capsule. (a) Section of B-scan

following reflection and shadow stripe filling, and Gaussian blurring. (b)

Blurred B-scan section with overlay of output from weak Canny edge

detection (blue) and higher sensitivity Canny edge detection (red and yellow

lines). The yellow line indicates the selected capsule-cortex interface from the

higher sensitivity edge detection output.

3.5.6 Segmentation of Quantifiable Kidney Cortex

Segmentation of the area of quantifiable kidney cortex was a necessary step for assessing

the degree of swelling of the PCTs. Variations in capsule thickness, adipose, and OCT performance

led to widely variable penetration into the kidney cortex both between kidneys and between

different regions in the same kidney. To accurately segment the quantifiable cortex, it was

Page 124: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

107

necessary to identify features which could help discriminate between noise and strong-signal

regions of cortex.

The z-axis depth of 1.9 mm should at no point penetrate past the renal cortex, which has

an average thickness of just less than ~14 mm [162]. The cortex is densely populated with PCTs.

If swelling is minimal, the lumen of the PCTs should be visible throughout a cross-section of the

cortex with their visibility restricted only by the limitations of the OCT system’s penetration. In

images where PCT lumen were readily visible, the volume of quantifiable cortex could be inferred

as the area immediately surrounding lumen with distinct edges, with anything beneath that point

defined as background/noise beyond the penetration of the OCT system (Figure 3.12a). However,

in images where PCT were fully swollen and their lumen fully occluded by their swollen

epithelium, there were often no visible anatomical landmarks to help distinguish strong-signal

cortex from noise (Figure 3.12b). Quantifiable cortex in these images was challenging even for

trained raters to identify.

Intensity alone was likewise not a consistent marker of quantifiable cortex as the average

intensity of the cortex varied widely between scans (Figure 3.12c, d). Similarly, the intensity

gradient marking the transition between signal and noise was widely variable between scans. Some

scans had a rapidly diminishing intensity as the penetration increased beyond where signal was

present (Figure 3.12c), while in other scans intensity was roughly homogenous between the signal

and noise (Figure 3.12d). Automating segmentation to accurately identify quantifiable cortex in

images with and without visible PCT lumen required a weighted combination of PCT lumen edge

strength, texture estimates, and intensity values.

Page 125: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

108

Figure 3.12: Heterogeneity of cortex appearance. (a) Section of a B-scan

with visible PCT lumen. (b) Section of a B-scan with no visible PCT lumen

and no anatomical landmarks. (c) Section of a B-scan with rapidly

diminishing intensity values as the FOV moves past the OCT system’s

penetration into the cortex. (d) Section of a B-scan with little reduction in

intensity values as the FOV moves past the OCT system’s penetration into the

cortex.

Maps of lumen edge strength were generated by a local standard deviation filter passed

over the original B-scan with a contrast adjusted output (kernel=15x25 µm) (Figure 3.13b). Texture

was estimated with a second vertically weighted standard deviation filter (kernel=45x2.73 µm),

which more clearly highlighted transition from signal to noise in regions where no or little lumens

were present (Figure 3.13c). Intensity values were drawn from the original B-scan and were

weighted by variables reflecting the contrast between the capsule and superficial cortex, the

contrast between cortex and lumen, and the intensity gradient beneath the capsule-cortex interface.

In each instance, higher contrast or degree of gradient increased the weighting of the B-scan

intensity values. Weighted intensity values were combined with the map of lumen edge strength

and texture to yield a greyscale image (Figure 3.13d) from which quantifiable cortex could be

inferred by thresholding (area between yellow lines in Figure 3.13d).

Page 126: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

109

Figure 3.13: Cortex segmentation. (a) Original section of a B-scan. (b) Map

of lumen edge strength with brighter regions corresponding to stronger lumen

edges. (c) Texture estimate generated by a vertically weighted standard

deviation filter and contrast adjustment. (d) Greyscale output of the weighted

combination of a-c. The area between the green line and bottom yellow line

represents the output of the thresholding of the image. The area between the

top yellow line (derived from the capsule-cortex interface segmentation step

(section 2.6.4) and the bottom yellow line represents the final segmented area

of quantifiable cortex.

The lumen edge strength map (1st standard deviation filter) was generated for lumen

selection and re-incorporated into cortex segmentation for computational expediency. In images

where PCT lumen were readily visible, lumen cross-sections were a good indication of

quantifiable cortex. Lumen edge strength in these images deteriorates as the A-scans move

beyond the penetrating depth of the OCT system. The contribution of this standard deviation

filter therefore supports quantifiable cortex segmentation primarily in images where lumen are

plentiful. The texture estimate (2nd standard deviation filter) utilizes a more vertically weighted

filter, which does a better job highlighting the transition from signal to noise (i.e. discriminating

between kidney cortex and background) in images where there are no lumen present.

Page 127: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

110

3.5.7 Segmentation of Proximal Convoluted Tubule Lumen (Region of Interest Map for

Automatic Selection)

Prior to PCT lumen selection, a map of all potential PCT lumen was generated with a

combination of local adaptive thresholding and global thresholding around an empirically

determined level. The local adaptive thresholding binarized the original B-Scan tile by tile within

an approximately 70x70 µm window, with a threshold defined by the values within each tile

(Figure 3.14b). The global thresholding was performed on a contrast-enhanced version of the

original image where contrast was enhanced with MATLAB’s histeq function in a similar tiled

fashion based on the range of values in each tile (Figure 3.14c).

Figure 3.14: Lumen segmentation. (a) Contrast enhanced section from B-

scan (same section as used in Figure 3.13) following adaptive histogram

equalization. (b) Binary output of adaptive thresholding performed on original

B-scan. (c) Binary output of global thresholding performed on the contrast-

stretched image (Figure 3.13a). (d) ROI map generated after combining b and

c.

The adaptive thresholding was especially proficient in locating potential regions of interest

(ROIs) throughout the image but was indiscriminate in identifying ROIs and routinely located

them throughout regions of noise. The global thresholding was less comprehensive in its

identification of ROIs but was better able to discriminate between signal and noise (i.e. identified

high-noise regions as a single large ROI). The logical sum of the two thresholded images produced

Page 128: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

111

a comprehensive binary mapping of ROIs with a signal to noise weighting component Figure

3.14d). ROIs outside of the segmented quantifiable cortex were removed from the ROI map.

Binarization steps were optimized empirically to produce automatic ROI segmentation which

closely mimicked the segmentation of manual raters. Denoising and smoothing steps coupled with

moderately generous thresholds for binarization may have produced over-estimation of kidney

feature size but were necessary to ensure proper segmentation (i.e. not bisecting single PCT cross-

sections or producing other artifacts).

Selection of PCT lumen was the most subjective of the manual segmentation processes and

varied considerably between raters. Manual raters were instructed to segment regions within the

quantifiable cortex which they could, with confidence, identify as cross-sections of PCT lumen.

Criteria for selection included size consistent with PCT lumen, and well-defined lumen edges such

that the selection could be reliably distinguished from imaging artifacts or noise. Raters were

instructed not to segment ROIs which could be confidently distinguished from PCTs as glomeruli

(characterized by their ~200 µm diameter and capillary tuft), blood vessels (characterized by large

diameter lumen and length relative to PCT lumens), or cysts (characterized by their >200 µm

diameter, and irregular shape). Features from manual selections were summed and employed to

define thresholds for inclusion/exclusion of automatic selections.

For selection of PCT lumen from the ROI map, a classification model was generated using

features extracted from manual segmentation. A set of features including edge-strength, diameter,

and depth beneath the capsule-cortex interface were extracted from automatic ROI selections

which coincided with PCT lumen selections made during manual analysis. These features were

similarly extracted from automatic ROI selections which manual raters did not select as PCT

Page 129: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

112

lumen. These groups of features were assigned a “hit” or “miss” label respectively and were used

to train a fit binary classification tree using the MATLAB fitctree function. A simple decision tree,

with sensitivity and specificity comparable to more complex models, was selected to ensure

robustness of the classifier. The classification tree was able to accurately select PCT lumen from

the ROI map with a sensitivity of 85.58% and a specificity of 89.04%. The classification tree was

employed following generation of the ROI map, wherein ROIs whose features yielded a “hit” were

included in analysis and ROIs whose features yielded a “miss” were excluded.

Additional criteria were set for exclusion of false ROIs and imaging artifacts. Vertical

stripes of shadowing and areas between adjacent stripes of reflection were routinely falsely

identified as PCT lumen (Figure 3.15a). The shadow and reflection masks generated during capsule

segmentation were utilized here to remove these ROIs from selection at points where these masks

overlapped with selected ROIs (Figure 3.15b). Similarly, separation between the renal capsule and

cortex, while infrequent, created ROIs which were frequently identified as PCT lumen (Figure

3.15c). These ROIs were identified by their proximity to the capsule-cortex interface and their

horizontally elongated appearance, and were excluded from analysis (Figure 3.15d).

Figure 3.15: False regions of interest removal. (a) Section of a B-scan with

adjacent stripes of reflection (surrounding left arrow) and shadowing (right

arrow) which produce false ROIs. (b) B-scan section from 3.15a with yellow

representing the corresponding ROI map generated for this image. The

vertical red stripes represent detected stripes of reflection while the vertical

Page 130: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

113

blue stripe represents detected stripes of shadowing. (c) Section of a B-scan

with separation between capsule and cortex. The arrow indicates pockets at

the point of separation which produce false ROIs. (d) B-scan section from

3.15c with yellow representing the corresponding ROI map. The arrow

indicates a region (red) where separation of the capsule and cortex produces

false ROIs.

3.6 Comparison of Automatic and Manual Segmentation

Automatic segmentation performed on images which were also manually segmented (~1,500

images) had a capsule-cortex interface with a mean absolute error (MAE) of 15.0 ± 10.7 µm (5.2

± 3.7 pixels) as compared to the manual segmentations (top yellow, blue and red lines in Figure

3.16b for automatic and the 2 manual raters respectively). Multiple raters performing manual

segmentation on the same images deviated by an average of 11.5 ± 5.9 µm (4.0 ± 2.0 pixels).

Automatic segmentation performed on images which were also manually segmented produced a

quantifiable cortex boundary (line across x-axis highlighting the point at which signal transitions

into noise, represented by the bottom yellow, blue and red lines in Figure 3.16b for automatic and

the 2 manual raters respectively) with a MAE of 45.0 ± 11.23 µm (4.0 ± 2.0 pixels) as compared

with manual segmentations. Multiple raters performing manual segmentation on the same images

deviated by an average of 59.0 ± 29.3 µm (20.8 ± 10.7 pixels).

Sørensen-Dice similarity coefficient scores were calculated to demonstrate the degree of

agreement (agreement between methods as to what area was segmented as quantifiable cortex and

what area was excluded; a Dice score of 0 would indicate no agreement whereas a Dice score of 1

would indicate perfect agreement) between manual and automatic selections of cortex volume

(area between the segmented capsule-cortex interface and the segmented quantifiable cortex

boundary). Automatically segmented cortex volumes compared to manually segmented cortex

Page 131: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

114

produced a Dice score of 0.84 ± 0.05. Comparison between manual raters’ segmentations produced

a Dice score of 0.81 ± 0.06.

Figure 3.16: Manual versus automatic segmentation overlay. (a) Original

B-scan (same as used in Figure 3.5) and (b) B-scan following segmentation

automatically and by 2 manual raters. Segmentation of the capsule-cortex

interface is represented by the top yellow, red and blue lines as segmented by

the algorithm and 2 manual raters respectively. Segmentation of the

quantifiable cortex boundary is represented by the bottom yellow, red and

blue lines as segmented by the algorithm and 2 manual raters respectively.

Automatic PCT lumen selections are represented in green if they overlap with

either of the manual rater’s selections and yellow if they do not overlap with

manual segmentation. Manual PCT lumen selections are represented in cyan

if they overlap with 2nd rater’s selections and red or blue for each rater if there

is no overlap.

To assess reproducibility among manual raters, raters were reassigned 25 B-scans,

randomly selected from B-scans which they had previously segmented. MAE was calculated, for

segmentation of the capsule-cortex interface, between each rater’s two segmentations for each B-

scan, and ranged from 9 to 15 µm between raters. Dice scores were similarly calculated between

Page 132: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

115

each rater’s two segmentations of quantifiable kidney cortex and ranged from 0.77 to 0.9, with

most raters achieving >0.8. Cohen’s kappa coefficients were calculated between PCT lumen

selections in both sets of segmented images and demonstrated fair to moderate agreement, with a

range in scores between 0.38 and 0.6. Kappa coefficients improved dramatically to a range of

scores between 0.55 to 0.72 when assessing only images with at least moderate (>5%) density

(Table 3.1).

Table 3.1: Intra-rater reproducibility and algorithm performance.

Reproducibility measurements for manual raters (left) reassigned 25 B-scans

each from their original sets. MAE, Dice coefficients, and Cohen’s kappa

coefficients are calculated for reproducibility in capsule-cortex interface,

quantifiable cortex, and PCT lumen selections respectively. Kappa scores are

also shown for only B-scans where density measurements were >5% (i.e.

there was not a low population of tubule lumen). Comparison between manual

raters’ initial segmentations of the 25 reassigned images and automatic

segmentation performed on those same images is also shown (right).

3.6.1 Measurement Extraction

3.6.1.1 Density Measurements

PCT swelling was rarely homogenous within a kidney. Swelling (as evidenced by a

reduction in visible PCT lumen size) often varied within a kidney with some PCT lumen being

fully occluded by swelling, while others had little reduction in lumen diameter. Since fully

occluded PCT lumens were not visible in the OCT image sets, measurements of PCT morphology

in these instances would be biased by only including less swollen PCT lumen. To supplement

tubular measurements with a measure which accounts for the influence of fully occluded PCT

lumen, a “density” measure was devised which calculates the total area of PCT lumen divided by

MAE Dice Kappa Kappa at >5% MAE Dice Kappa Kappa at >5%

Rater 1 10.6 0.90 0.38 0.58 13.2 0.89 0.17 0.50

Rater 2 9.2 0.85 0.47 0.72 12.6 0.83 0.23 0.65

Rater 3 12.0 0.82 0.60 0.62 13.2 0.87 0.21 0.52

Rater 4 15.2 0.77 0.38 0.55 16.4 0.79 0.13 0.35

Intra-Rater Reproducibility Performance against Automatic Segmentation

Page 133: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

116

the total area of quantifiable cortex. This method does not neglect the swelling of PCTs whose

lumen is fully occluded, but instead provides an estimate of the ratio of the total area of PCT lumen

to quantifiable cortex for each B-scan. A high diameter measurement may, for example, be taken

together with a low density measurement suggesting a small population of dilated PCTs within a

B-scan showing mostly occluded lumen. The density measurement also provides a number value

for the criteria technicians were instructed to pursue (technicians were instructed to preferentially

image regions with a higher total area of PCT lumen).

One limitation of the 2D imaging protocol is that B-scans intersect the PCTs randomly

(horizontal red line in Figure 3.17b) and do not necessarily create cross-sections orthogonal to the

direction of the tubule (blue plane in Figure 3.17b). This creates elongated and irregularly shaped

cross-sections (red shape in Figure 3.17c) which may misrepresent the cross-sectional area of PCT

lumen (blue shape in Figure 3.17c). This impacts the density measurement, with non-orthogonal

cross-sections contributing a greater amount to the total lumen area than the corresponding true

orthogonal cross-section (red and blue shapes respectively in Figure 3.17c). To adjust for this bias,

a set of features including circularity, extent, and eccentricity were compiled for every B-scan

cross-section in the 3D scan performed on a preserved kidney (red line and red shape in Figure

3.17b and Figure 3.17c respectively). The true area of each cross-section was acquired by capturing

a plane (blue plane in Figure 3.17b) at the same location (yellow arrow in Figure 3.17b) orthogonal

to the orientation of that section (40 µm section length) of the tubule.

Page 134: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

117

Figure 3.17: 3D lumen reconstruction with B-scan and orthogonal

sectioning. (a) En face view of automatically segmented PCT lumen in a

reconstructed 3D scan. Each tubule was extracted for comparison of B-scan

cross-section features to features of cross-sections taken orthogonal to the

orientation of the PCT at the same locations. The tubule segment highlighted

in green is represented in 3.17b. (b) The red line represents the location of the

B-scan while the blue plane represents the plane orthogonal to the orientation

of that section of the PCT lumen segment (orientation from position 20 µm

earlier in the segment to 20 µm further). The arrow indicates the point on the

tubule where the cross-sections in 3.17c are captured. This process was

repeated at every point along the length of the tubule. (c) Resulting B-scan

and orthogonal cross-sections from 3.17b are represented in red and blue

respectively.

The B-scan cross-section features were fed as inputs into MATLAB’s “Regression Learner

App” with the percent reduction in area from the B-scan cross-section to the true cross-section as

the response. A linear regression model was trained with 10-fold cross-validation to predict the

percent reduction in area required to transform an elongated or irregularly shaped cross-section

into the area of the corresponding orthogonal cross-section. The model yielded a root-mean-square

error (RMSE) of 0.15 and an R-Squared value of 0.69. The linear regression model was employed

to correct the area of elongated and irregularly shaped cross-sections to the area of the

corresponding true cross-sections. A notable limitation of this correction method, however, is that

only one kidney was used for training of the model. In addition, this kidney was preserved in a

formaldehyde solution and so may not accurately represent PCT morphology of a kidney used for

transplant. Similarly, feature evaluation of the orthogonal cross-sections revealed that these

Page 135: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

118

sections were, on average, moderately elliptical (eccentricity of 0.67 ± 0.15); orthogonal cross-

sections contained, on average, a minor axis to major axis length ratio of 3:4. Consequently, the

linear regression model, depending on input features, may produce area estimations of non-circular

orthogonal cross-sections. While orthogonal sectioning of tubules in kidneys preserved for

transplant likely do not consistently produce perfectly circular lumen cross-sections due to

anatomical heterogeneity and storage effects, it should be considered that the formaldehyde

preservation of the kidney used in the linear regression model may have altered circularity of

tubular lumen.

3.6.1.2 Diameter Measurements

The diameter of lumen in PCT cross-sections was measured for all cross-sections in each

B-scan. As the epithelium of the PCTs swells, the visible lumen should reduce. Conversely, as the

epithelium is flattened or simplified, the visible lumen should increase. Diameter of the PCT lumen

should therefore maintain an inverse relationship to the degree of swelling, and a direct relationship

to the degree of epithelial flattening/simplification.

Diameter measurements are similarly impacted by the limitations of the 2D imaging

protocol, with elongated non-orthogonal sections (red in Figure 3.18) potentially misrepresenting

true lumen diameter. To circumvent this issue, diameter was defined as the “minor axis length”

(shortest diameter which passes through the center of the ROI). This definition ensures that the

elongated axis of tangential sections does not bias the diameter measurement, however, this may

result in under-representation of the true diameter if the imaging plane does not cut through the

tubular center axis. Consequently, an additional diameter measurement, derived from the corrected

Page 136: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

119

area, was used. This measure calculated diameter from the linear regression corrected area using

the equation for calculating the area of a circle (𝑨 = 𝝅𝒓𝟐).

To assess accuracy, a 50 µm capillary phantom was embedded in an agar solution which

mimicked the scattering properties of kidney tissue. OCT scans were performed on the phantoms

at three locations, and ROI maps were generated by the described method. Diameter of the interior

of the capillary phantoms was calculated by the two methods described in this section and

produced diameters of 45.7±2.9 µm and 50.3±3.1µm as measured by minor axis length and from

corrected area respectively.

3.6.1.3 Inter-Lumen Measurements

The minimum distance between edges of adjacent lumen was measured between all adjacent

PCT lumen cross-sections in each B-scan (green in Figure 3.18). Adjacency of ROIs was defined

as when centroids were within 110 µm of each other (determined empirically as the maximum

distance before tubule lumen outside of immediate adjacency were included) (red circle in Figure

3.18). This inter-lumen distance was considered a measurement of the combined thickness of the

epithelium of two adjacent PCTs and any interstitial space. As the epithelium swells, the inter-

lumen distance should increase. Conversely, as the epithelium is flattened or simplified, the inter-

lumen distance should reduce.

Page 137: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

120

Figure 3.18: Depiction of methodology for inter-lumen and inter-centroid

measurements. The red circle represents a 110 µm radius around the center

ROI of “adjacent” ROIs. Distances between lumen edges and centroids are

represented in green and blue respectively.

3.6.1.4 Inter-Centroid Measurements

The distance between centroids of adjacent PCT lumen was similarly measured between

all adjacent PCT lumen cross-sections in each B-scan (blue in Figure 3.18). This was considered a

measurement of the combined lumen, epithelium, and interstitial space. The inter-centroid distance

may be mostly unaffected by PCT swelling and epithelial flattening as changes to epithelial

thickness and lumen diameter are inversely related and may balance. The inter-centroid distance

may therefor reflect changes to the interstitial space.

3.6.2 B-Scan Selection and Measurement Compilation

Measurements were compiled for each B-scan in each image set. As the 2D imaging

protocol produced numerous duplicate or redundant images, only one B-scan was selected from

each image set for analysis. As imaging protocol was to survey regions with the greatest area of

visible tubule lumen (i.e. highest PCT lumen density), B-scan results were sorted by density and

the maximum density B-scan was selected for inclusion in results. Measurements from these

selected B-scans were averaged to yield values for pre-implantation and post-reperfusion scans for

each kidney. Results were averaged for each recovery group in each transplant group and

represented in box and whisker plots.

Page 138: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

121

In addition to analysis of correlation between measurements from selected B-scans and

binary recovery group categories (IGF/DGF), the relationship between measurements and

decline in patient’s serum creatinine levels (which should decline rapidly and to a level <3.0

mg/dL if a transplanted kidney is well functioning) following transplant was investigated [163].

Linear mixed effects models were fitted to regress the longitudinal measures of serum creatinine

from day 0 to day 5 on each patient to account for the within-subject variation by assuming an

first order auto-regressive structure with homogenous variances covariance structure and

allowing for random intercepts for between-subject variation. The baseline creatinine measure,

time, and interactions between time and each measurement were also included in the models.

Models were fitted following our initial hypotheses that flattened PCT epithelium and dilated

lumen would represent pathology, and consequently higher inter-lumen distance measurements,

lower diameter measurements, and lower density measurements (which we initially predicted

would echo diameter measurement trends) would correlate with a faster recovery (steeper decline

in creatinine). Higher inter-centroid distances were hypothesized to represent pathology (as

indicative of interstitial inflammation), and consequently lower inter-centroid distances would

correlate with a faster recovery (steeper decline in serum creatinine).

3.7 Summary

One complication with OCT analysis of human kidneys is the very high number of

images captured that need to be analyzed; pathologists could be presented with potentially

thousands of images for pre-implantation and post-reperfusion analysis. Filtering through these

sets is not only time consuming, but introduces a point where considerable variability or bias

between raters and pathologists may be introduced. For OCT to be used effectively in a clinical

Page 139: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

122

setting, image analysis must be conducted quickly, reliably, and without bias. Automated

segmentation achieves these goals and can provide rapid and accurate assessments of the kidney

[164], [165]. Automated selection and analysis of images ensures both reproducibility and speed.

The reproducibility of automated analysis should promote confidence in the measurements

derived from OCT. The speed of automated analysis, coupled with the real-time imaging speed

(relative to the several hours required to prepare biopsy histology), should position OCT as

entirely practical in an OR setting.

The benefits of automated segmentation in assessing transplant kidneys have already

been explored in the context of renal biopsy analysis. Wide variability between pathologists in

interpreting and scoring kidney biopsies has led to the introduction of numerous automated and

semi-automated systems for quantifying histology [166]–[168]. These systems ensure

reproducibility and have the added benefit of high speed analysis. These benefits are of particular

value at points when a decision to accept or reject a kidney is pressing, and when the only

pathologists on call are not expertly trained. Biopsy histology, however, appears vastly different

from the anatomy present in OCT imagery. The majority of automated systems for quantification

of biopsy histology rely on color features and other markers not present in OCT, leaving these

methods unsuitable for application to OCT imaging of the kidney. The majority of research in

automated segmentation of OCT imaging is focused on segmentation of features in the eye,

namely segmentation of the retinal layers [169],[170]. These algorithms target vastly different

anatomy than is seen in the kidney cortex and encounter different image processing hurdles due

to vastly different imaging protocols, and so are similarly unsuitable for segmentation of OCT

kidney imagery.

Page 140: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

123

The segmentation algorithm proposed in this chapter is uniquely designed to segment

kidney features captured by OCT. This algorithm was designed to provide accurate, and high

speed segmentation and measurement of kidney microanatomy in a fully automated fashion. The

reproducibility, speed, and accuracy of the proposed algorithm have facilitated the analysis of

vast amounts of OCT imaging of human kidneys for the purpose of research.

Page 141: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

124

CHAPTER 4: CLINICAL RESULTS

4.1 Introduction

We open this chapter with an overview of the categorization schema used for partitioning

a heterogeneous patient population into smaller groups. We then investigate, in each sub-

grouping of patient populations, the correlation between measurements derived from OCT

imaging of donor kidneys with the recovery of the respective recipients. Next, we evaluate

redundancy of measurements, and isolate the measures most predictive of recovery within each

patient grouping. We conclude this chapter with a brief discussion on the specific pathology that

may be represented by the described measurement trends.

4.2 Evaluating Donors

4.2.1 Live and Deceased Donor Kidney Transplantation

The majority of kidneys transplanted in the United States are from deceased donors. The

advantages of living donor kidney transplants (LDKTs) over deceased donor kidney transplants

(DDKTs) are well known. Recipients who receive a kidney from living donors will recover

faster and their kidney grafts will survive roughly twice as long. This is due to a variety of

factors. Major contributors to this disparity are the improved health of living donors over

deceased donors (prior to death), and the reduced ischemic period of LDKTs. While many

DDKTs are procured from otherwise healthy donors who suffered premature death, many are

procured from elderly donors or donors with various comorbidities which contributed to the

Page 142: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

125

cause of death and may impact kidney viability. Similarly, LDKTs are generally scheduled far in

advance, which affords transplant surgeons the opportunity to optimize transplant conditions.

Most LDKTs are performed with donor and recipient in adjacent rooms, meaning ischemic time

incurred between procurement and transplant into the recipient is greatly reduced (approximately

1 hour for LDKTs versus up to 30 hours for DDKTs) [171]–[173].

4.2.2 Static Cold Storage and Hypothermic Machine Perfusion in Kidney Transplantation

Reducing the temperature of storage for kidneys during ischemic time slows enzymatic

degradation and depletion of metabolic stores [174]. The Van Hoff rule dictates that the majority

of cellular enzymes respond to a 10-degree Celsius drop in temperature with a 2-fold drop in

activity [175]. A kidney exposed to warm ischemia shows some detriment in function after only

a 5-minute interval, and is damaged to the point where it becomes unviable after approximately

60 minutes [176]. Simply lowering the storage temperature to 4-degrees Celsius, even without

preservation solution, limits PCT metabolic activity to 10-12% its normal physiologic activity

and increases the time the kidney can remain viable to 12 or 13 hours [43]. Static cold storage

(SCS) of donor kidneys, however, utilizes preservation solutions which prolong viability of SCS

kidneys significantly.

While hypothermic storage drastically reduces the rate of energy consumption and

enzymatic degradation, these processes still continue at a slowed pace; ATP and ADP reserves

continue to become depleted at a rate which exceeds their ischemically stunted production.

Ischemic damage therefore remains, even under hypothermic conditions, a duration-dependent

condition. Unavoidable logistics like tissue matching and transport mean that the cold ischemic

Page 143: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

126

time can extend for long periods of time. The United Network of Organ Sharing (UNOS) registry

cite an average hypothermic storage time of 21 hours. The US Renal Data System Registry

suggests that each 6 hours of cold ischemia correspond with a 23% increase in risk of DGF

following transplant, although some labs suggest a more moderate 8% increase in DGF for every

6 hours of cold ischemia [43], [177]. Thus the duration of hypothermic storage and cold ischemia

is widely considered the greatest contributor to injury upon reperfusion (IRI) and subsequent

DGF [176]. In contrast to this trend, Xia et al. found no significant difference in survival time of

kidneys or their inclination towards DGF and a period of prolonged cold ischemia. This study

had a small sampling size however and a direct link between cold ischemic time and incidence of

DGF and subsequent reduction in survival does seem to be the consensus [178].

Enzymes respond heterogeneously to changes in temperature. Hypothermic preservation

maintains a temperature between 0 and 10 degrees, generally 4 degrees Celsius. In this

temperature range, Na/K/ATPase activity is almost non-existent (at 5 degrees Celsius,

Na/K/ATPase operates at 1% of its normothermic capacity). Hypothermic conditions therefor

induce similar effects as hypoxia on the Na/K/ATPase and maintenance of the sodium

electrochemical gradient. Similarly, by the time hypothermic storage is implemented, enough

warm ischemia has likely occurred to deplete the kidney cortex’s oxygen reserves and shift the

PCTs to an anaerobic metabolic state.

Respiration requires the translocation of adenine nucleotides across the mitochondrial

membrane by the adenine-nucleotide-translocator. While synthesis of adenine nucleotides

continues under hypothermic conditions, the translocating enzyme ceases function. Similarly,

transport of NADH across the mitochondrial membrane through the malate-aspartate shuttle is

Page 144: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

127

debilitated by a cold-induced reduction in enzyme function. Hypothermic paralysis of these

transporters therefor further stunts energy production [38], [179]. The lack of energy driving the

Na/K/ATPase and the cold’s paralyzing effects result in a failure to efflux sodium from the cell;

as the gradient is disrupted and sodium continues to diffuse down its electrochemical gradient

into the cell, sodium accumulates intracellularly and water follows osmotically resulting in cell

swelling [179], [180].

Hypothermic preservation also serves to destabilize the PCT cell membrane. Reduction in

temperature drives the phospholipids of the cell membrane from a liquid crystalline state to a

highly ordered hexagonal lattice. This phase change increases the activation energy of

membrane-bound enzymes, dramatically reducing their efficiency [181]. The altered structure

and enzymatic function of the membrane alter the cell’s permeability and ionic composition

respectively, leaving the cell less able to extrude sodium and more susceptible to its intrusion.

Hypothermic conditions also slow the cell’s defenses against free radicals to a greater

extent than they slow the reactions which produce those free radicals. This discrepancy leads to

an accumulation of free radicals within the cell. These highly reactive species react with

membrane lipids, further disturbing the integrity of the membrane and contributing to membrane

permeability [182].

Continuous hypothermic machine perfusion (HMP) was shown in the late 60’s by Belzer

et al. to preserve the viability of an ex vivo dog kidney for up to 72 hours [183]. Today,

continuous hypothermic storage occurs either with use of the LifePort Organ Recovery System

(ORS) or the Waters Medical System (WMS). Multiple studies have demonstrated the ability of

Page 145: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

128

these systems to reduce the degree of IRI as evidenced by reduced instances of DGF given the

same duration of cold ischemia as static cold storage grafts [184], [185]. The required

equipment, facilities, trained personnel and cost have limited this potentially superior method of

preservation to cases with higher risk/extended criteria kidneys.

HMP is typically performed at a higher temperature than in static cold perfusion. At 10

degrees Celsius, it is believed that ATP generation and Na/K/ATPase efficiency, while reduced,

are sufficient to maintain ionic gradients and support cell viability and cell volume to some

degree if provided the necessary substrates. As opposed to SCS, HMP generally uses

“extracellular” preservation solution which bears more similarity to plasma than the intracellular

environment. Oxygen and metabolic substrates are perfused through the kidney at a steady low

pressure (between 40 and 60 mmHg) or in a pulsatile fashion. While some groups maintain an

advantage in pulsatile over steady flow, others suggest the two modes of perfusion share no

significant difference in outcome. Similarly, certain groups promote a higher rate of flow during

pulsatile perfusion [183][186]. Oxygen administration during HMP is similarly debated: some

claim introduction of oxygen into the perfusate, either by oxygenation or by passive equilibration

with the air, as instrumental in maintenance of compromised kidneys, while others maintain that

its’ inclusion in protocol has no bearing on outcome or viability of the allograft [183].

In theory, the delivered substrates should be steadily converted to ATP by metabolic

mechanisms which are less inhibited by the higher storage temperatures. This ATP should in turn

power the Na/K/ATPase which is again less inhibited by the higher storage temperatures,

actively extruding sodium from the cell and maintaining proper cell volume of the PCTs. The

higher ATP supply and lower intracellular sodium concentrations should help to sustain lower

Page 146: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

129

intracellular calcium levels by active extrusion and extrusion by the sodium-calcium antiporter

respectively.

In addition to delivery of substrates and oxygen, this system has the added benefit of

clearance of metabolic waste products. Clearance of metabolic waste by the active flow of the

perfusate should prevent accumulation of reactive oxygen species and acidic components within

the cell. HMP therefore should act to maintain low intracellular calcium concentrations, a stable

pH, clearance of ROS [187].

Recent studies have also focused on continuous perfusion performed at normothermic

and sub-normothermic temperatures. Perfusion at temperatures between 25 and 37 degrees

Celsius have been shown to maintain graft viability. Normothermic and sub-normothermic

preservation strategies have the potential advantages of maintaining near-physiologic metabolic

function. Under a healthy metabolic state these kidneys could retain most functional and

structural characteristics and would avoid the damaging effects of IRI [145]. In the porcine

model, Kerkhove et al. successfully maintained organ viability for over 24 hours at sub-

normothermic temperatures-although the 15 degree Celsius model was more successful than the

21 and 28 degree preservation strategies [188].

4.2.3 Standard and Expanded Criteria Donors in Kidney Transplantation

In 2002 the term extended/expanded criteria donors (ECD) was introduced to identify

higher risk kidneys (use of “extended” and “expanded” varies between publications). These more

marginal ECD kidneys generally perform more poorly and have a shorter period of allograft

survival than kidneys procured from standard donors. As such, they are reserved for recipients

Page 147: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

130

where the reduced performance and shorter graft life would still culminate in a better quality of

life and improved prognosis over the alternative: remaining on dialysis and waiting for a more

ideal offer. This term was initially derived from a binary scoring system which identified all

donors 60 or older, and all donors between 50 and 59 who satisfied at least two other criteria

associated with poor kidney function as ECD. These criteria included a history of hypertension, a

cerebrovascular accident as the cause of death, and a serum creatinine of greater than 1.5 mg/dL.

Kidneys which satisfied either of these two definitions were considered to hold a 70% greater

change of allograft failure than kidneys procured from standard criteria donors (SCD). Both age

and these additional criteria were considered markers of reduced nephron mass, or a reduction in

the number of functional units of the kidney [189].

Age alone is not a reliable indicator of kidney function and viability. The effect of age on

the kidney can be heterogeneous. This is one shortcoming of the 2002 binary ECD scoring

system; kidneys from donors 60 or older may in some instances outperform kidneys from

younger donors and evidence of this is discounted with the 60 year ECD cutoff [190]. The

inclusion of additional factors in the evaluation criteria can help identify these instances. Since

the 2002 introduction of the binary ECD/SCD scoring, various groups have put forward more

comprehensive methodologies for assessing donor characteristics. Following a study of nearly

35,000 deceased donor transplants conducted over a 5 year period, Nyberg et al. introduced a

system which considered donor age, history of hypertension, creatinine clearance, cause of death,

and number of HLA-mismatches (human leukocyte antigen matching to determine suitable

donor-recipient pairings). Resulting scores ranged from 0-39 with scores above 20 indicating

Page 148: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

131

ECD [191], [192]. Schold et al. added CMV serology (to determine cytomegalovirus infection),

donor ethnicity and cold ischemia as donor variables, and assigned a score from 1 to 4 [193].

In 2009, Rao et al. introduced a system of kidney donor assessment which encompassed

10 donor characteristics. A regression model was constructed which traced and quantified the

association between a litany of donor characteristics and duration of graft survival. Training data

for this model drew from just under 70,000 transplants conducted between 1995 and 2005 and

included only first-time recipients and only deceased donor single-kidney transplants. Donor

characteristics which demonstrated no significant link to graft survival were discarded from the

model (e.g. gender, history of smoking). The ten remaining features included were age, height,

weight, ethnicity,, history of hypertension or diabetes, cause of death, serum creatinine, hepatitis

C and donation after cardiac death. Age was the most heavily weighted feature of the model with

each year after 50 increasing the predicted risk of graft failure by 1%. The output from this

model, titled the kidney donor risk index (KDRI), was a score from 0 to 100 representing the risk

of graft failure in an average adult recipient. A low KDRI would indicate a low risk of graft

failure while a higher KDRI would indicate a higher probability of graft failure.

In December of 2014, the Organ Procurement and Transplantation Network (OPTN)

implemented a new kidney allocation system (KAS) with the intent of optimizing use of the

donor pool and increasing consideration and use of less ideal but still viable kidneys. In addition

to adjusting the system of triage for ordering the transplant list, the KAS introduced a

standardized system for assessing kidney donors and scoring donor characteristics. The kidney

donor profile index (KDPI) was proposed as the standard for quantifying and interpreting donor

features in their relation to kidney viability. The limitations of previous methods of scoring

Page 149: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

132

donor factors were believe to lead to greater aversion by transplant surgeons, and ultimately to a

greater number of unnecessary discards. The KDPI was introduced as a more comprehensive

scoring system that had greater correlation with graft outcome. The KDPI was intended to reduce

discards by offering surgeons a more detailed look not just at the quality of donor, but of the

quality of the donor in relation to the rest of the donor pool.

The KDPI is a derivation of Rao et al.’s KDRI. KDPI is similarly a scoring from 0 to 100

but rather than a risk of graft failure it reflects a mapping of the KDRI for that kidney against all

cadaver donor kidneys procured the previous year. In essence, the KDPI is a predictive ordering

of the donor pool from least to most likely to experience graft failure. A KDPI of 85%, for

example, would indicate that the kidney is associated with a KDRI higher than 85% of recovered

kidneys from the year prior. A higher KDPI therefor reflects not a direct high risk of graft failure

but rather a high risk of graft failure relative to the composition of the donor pool. Conversely, a

low KDPI suggests that a kidney is less likely than most of the donor pool to experience graft

failure. Under the KAS system, a donor with a KDPI of 85% or greater was considered as ECD,

while a donor with a KDPI of less than 85% was considered SCD.

4.2.4 Immediate and Delayed Graft Function in Kidney Transplant Recovery

The increasing use of higher risk, more ischemically damaged kidneys to combat the

daunting waiting list has led to an increase in incidence of poor graft function and graft failure.

While more kidneys are being transplanted overall, a higher percentage of kidneys transplanted

are compromised. In spite of advancements in treatment, the number of recipients experiencing

poor graft function has risen approximately 20% every decade since 1985. This rise in

Page 150: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

133

complications has grown in parallel with increased usage of ECD kidneys, with up to 50% of

kidneys from ECD donors experiencing delayed graft function (DGF) as compared to 4-10%

incidence of DGF in kidneys procured from live donors [194]–[196].

DGF is diagnosed clinically when a patient is required to undergo dialysis following

transplant (either prior to discharge or within 1 week following transplant). Half of those

diagnosed with DGF will begin to see a return in kidney function 10 days following transplant, a

third will see function begin to return 10-20 days following transplant, 10-15% do not see

function until after 20 days following transplant and 2-15% will experience primary non-function

or chronic graft dysfunction [1]. Speed of recovery is considered an indication of the expected

time of graft survival; a quick return to normal function following transplant is generally

associated with a longer lifetime of the transplanted kidney whereas a prolonged recovery period

generally corresponds with a shorter lifetime of the transplant [196]. Patients diagnosed with

DGF demonstrate a significantly higher incidence of acute rejection, a 40% reduction in lifetime

of the transplant and a 5-year graft survival rate of only 34% [197]. Following expiration of the

graft or rejection, these patients are returned to the kidney transplant waiting list if they remain

healthy enough for an additional transplant.

4.3 Patient Demographics

This study was approved by the Georgetown University and the University of Maryland

Institutional Review Boards (Study number: IRB#2010-396). Written informed consent was

obtained prior to enrollment. Patients eligible for this study included any kidney transplant

recipient 18 years or older at the MedStar Georgetown Transplant Institute.

Page 151: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

134

Patient demographics were obtained at the time of consent. The patient pool was composed

of approximately 60% male and 40% female recipients. Mean age at transplant was 52 ± 12.5

years. Mean BMI of recipients was 28.4 ± 4.7. 61% of patients in this study were African

American, 24% were Caucasian, 8% were Hispanic, and 7% were Asian.

Of the 169 kidneys imaged and included in this study, 66 were from LDKTs and 103 were

from DDKTs. All LDKTs were preserved by SCS. Of the 103 DDKTs, 88 were preserved by SCS

and 15 were preserved by HMP. 4 of the kidneys in the SCS group were part of a multi-organ

transplant (kidney/pancreas). Of the 88 DDKT-SCS kidneys, 26 had a KDPI of 85 or more and

were subcategorized as ECD kidneys. The remaining 62 SCS kidneys were subcategorized as SCD

kidneys [3]. Of the 15 kidneys in the HMP group, 2 kidneys qualified as ECD, and the remaining

13 were subcategorized as SCD kidneys (Figure 4.1). Patients whose data were excluded from the

analysis included those involved in parallel studies for anti-DGF clinical trials (1 patient) and those

where image quality of the OCT image sets was compromised (2 patients).

Graft function following transplant was categorized as either immediate or delayed. DGF

was designated when a transplant recipient was required to undergo dialysis within the first

seven days following transplant [198] or when otherwise specified as DGF in clinical notes. All

cases where transplant recipients did not require dialysis prior to discharge were considered IGF.

Recovery groupings for each transplant group were as follows: LDKT (65 IGF, 1 DGF), SCS

SCD (51 IGF, 11 DGF), SCS ECD (18 IGF, 8 DGF), HMP SCD (8 IGF, 5 DGF), and HMP

ECD (1 IGF, 1 DGF) (Figure 4.1).

Page 152: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

135

Figure 4.1: Hierarchy classification of transplant groups. All transplants

(blue tier 1) divided into live and deceased donor kidney transplants (blue tier

2). DDKTs are further divided into subgroups based on storage method (blue

tier 3). DDKTs stored by SCS and HMP are further divided into subgroups

based on risk of graft failure (blue tier 4). Each end-tier transplant group is

divided into recovery groups based on requirement of dialysis (green and red).

4.4 Density by Area Results

4.4.1 Density by Area Results Stratified by Transplant Group (IGF and DGF Combined)

Distinctions between measurements from the ECD subgroup of DDKT kidneys stored by

HMP and other transplant groups were not investigated due to limited sampling of ECD kidneys

in the DDKT-HMP group (n=2).

To correct for multiple comparisons between transplant groups, p-values from Student’s t-

tests are supplemented with p-values after adjustment with false discovery rate (FDR). Multiple

comparison correction was performed in R (programming language) with the p.adjust() command

using the FDR method for measurements across transplant groups (i.e., LDKT, DDKT-SCS SCD,

DDKT-SCS ECD, and DDKT-HMP SCD measurements).

Page 153: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

136

Prior to implantation (left in Figure 4.2), kidneys from the LDKT transplant group

demonstrated higher (p<0.001, FDR adjusted p-value<0.001) PCT lumen density than DDKT

kidneys stored by SCS. This difference may be considered a consequence of the markedly different

transplant conditions, namely a considerably reduced ischemic time (mean of 1.47 ± 0.61 hours

for LDKT versus 13.49 ± 7.06 hours for DDKT-SCS SCD and ECD subgroups). The SCD

subgroup of DDKT kidneys stored by HMP had a higher (p<0.001, FDR adjusted p-value<0.001)

pre-implantation density than all other transplant groups. The high HMP density may be a result

of artificial dilation of the PCT lumen by the machine-perfusion process. The LDKT group, and

the DDKT-SCS SCD and ECD subgroups all experienced an increase in density between pre-

implantation and post-reperfusion scans. This is consistent with prior studies demonstrating a

dramatic reduction in swelling of ischemic PCTs (which would present as an increase to total

lumen area) following reperfusion [199], [200]. In contrast to all other groups, the HMP group

experienced a reduction in density following reperfusion, suggesting either some dissipation of the

artificial dilation or induction of swelling. Post-reperfusion density (right in Figure 4.2) was similar

between LDKT and the DDKT-SCS SCD and ECD subgroups. Post-reperfusion density in the

HMP group remained higher (p<0.05, FDR adjusted p-value=0.078 when compared to SCS-ECD

and FDR adjusted p-value =0.1136 when compared to SCS-SCD) than in both DDKT-SCS

subgroups, and moderately higher than in the LDKT transplant group (p=0.09, FDR adjusted p-

value=0.25). The high post-reperfusion density suggests some persistence of the effects of the

artificial dilation.

Page 154: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

137

Figure 4.2: Box and whisker plots of density measurements. Density

measurements (represented in lumen percentage of cortex, i.e., density

ratioE02) calculated with original lumen area (a) and with lumen area

corrected by linear regression (b) for pre-implantation (left) and post--

reperfusion (right) scans for the LDKT group, and the DDKT subgroups: SCD

kidneys stored by SCS, ECD kidneys stored by SCS, and SCD kidneys stored

by HMP. Each transplant group is further divided into recovery groups which

experienced either IGF (green) or DGF (red) following transplant. Mean

density values for each recovery group are included in the attached table with

p-values (from Student’s t-test), and values adjusted for false discovery rate

(FDR) between transplant groups, representing significance of difference

between recovery groups for each transplant group. The mean percent change

(increase or decrease) to density following reperfusion is included at the

bottom of each table for both recovery groups in each transplant group.

Page 155: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

138

4.4.2 Density by Area Results Stratified by Recovery Group (IGF vs. DGF)

Distinctions between IGF and DGF recovery group measurements in the LDKT transplant

group were not investigated due to limited sampling of DGF kidneys (n=1). Similarly, distinctions

between IGF and DGF recovery group measurements in the ECD subgroup of DDKT kidneys

stored by HMP were not investigated due to limited sampling (n=1 for IGF, n=1 for DGF).

In all transplant groups, density values were similar between IGF and DGF recovery groups

(green and red respectively in Figure 4.2) prior to implantation. Following transplant and

reperfusion, density measurements for the DDKT kidneys stored by SCS increased in both SCD

and ECD subgroups for both IGF and DGF recovery groups. In the HMP group, the IGF recovery

group experienced a <1% change in density while the DGF recovery group experienced a 23%

reduction in density following reperfusion. In the SCD subgroup of DDKT kidneys stored by SCS,

post-reperfusion density was similar between IGF and DGF recovery groups. In the ECD

subgroup, however, post-reperfusion density in the IGF recovery group was lower (p<0.05, FDR

adjusted p-value=0.03) than that of the DGF group. Conversely, in the HMP group, post-

reperfusion density in the IGF recovery group was higher (p=0.28, FDR adjusted p-value=0.30 for

original density, and p<0.05, FDR adjusted p-value=0.06 for corrected) than in the DGF recovery

group.

4.4.3 Density Results by Association with Post-Transplant Creatinine Decline

Following our initial hypothesis that lower PCT lumen density would correlate with a faster

recovery following transplant (i.e. density is positively correlated with creatinine values and lower

density is correlated with a steeper decline in creatinine (i.e., has a negative interaction effect with

Page 156: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

139

time), linear mixed effect models were fitted for each DDKT transplant group. The pre-

implantation fitted model for the SCS-SCD group did not support the hypothesis (p=0.89),

however the post-reperfusion SCS-SCD model trended towards support of the hypothesis

(p=0.09). Both pre-implantation and post-reperfusion fitted models for the SCS-ECD group

similarly did not support the hypothesis (p=0.74, and p=0.15 respectively). Finally, the pre-

implantation model for the HMP-SCD group did support the hypothesis (p<0.01), as did the post-

reperfusion model (p<0.001).

4.5 Diameter Results

4.5.1 Diameter Results Stratified by Transplant Group (IGF and DGF Combined)

Diameter measurements were relatively consistent between minor axis length and

corrected area methods of measurement. Diameter calculated from corrected area was, however,

moderately but consistently higher than diameter calculated as the minor axis length. This effect

is likely due to the linear regression model’s predictions of instances of moderately elliptical

orthogonal cross-sections, which the minor axis length would underestimate.

Prior to implantation (left in Figure 4.3), kidneys from the LDKT transplant group

demonstrated moderately higher PCT lumen diameter than DDKT kidneys stored by SCS. DDKT

kidneys stored by HMP had higher (p<0.001, FDR adjusted p-value<0.005) pre-implantation

diameter than all other transplant groups. All groups experienced an increase in diameter between

pre-implantation and post-reperfusion scans. The LDKT and DDKT-HMP groups both

experienced a modest 5% increase, while DDKT-SCS SCD and ECD subgroups both experienced

a larger increase in diameter (18%, and 13% respectively). Post-reperfusion diameter (right in

Page 157: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

140

Figure 4.3) was similar between the LDKT transplant group and the ECD subgroup of DDKT

kidneys stored by SCS. Post-reperfusion diameter in the SCD subgroup of DDKT kidneys stored

by SCS was moderately higher (p=0.08, FDR adjusted p-value=0.10) than in the ECD subgroup

and the LDKT transplant group (p<0.05, FDR adjusted p-value=0.06). Post-reperfusion diameter

in the HMP group was higher than in all other groups (p<0.005/FDR adjusted p-value<0.005,

p<0.05/FDR adjusted p-value=0.06, p<0.005/FDR adjusted p-value<0.005 for LDKT, DDKT-

SCD, and DDKT-ECD respectively).

Page 158: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

141

Figure 4.3: Box and whisker plots of diameter measurements. Measurements

calculated by minor axis length (a) and from lumen area corrected by linear regression (b)

for pre-implantation (left) and post-reperfusion (right) scans for the LDKT group, and the

DDKT subgroups: SCD kidneys stored by SCS, ECD kidneys stored by SCS, and SCD

kidneys stored by HMP. Each transplant group is further divided into recovery groups

which experienced either IGF (green) or DGF (red) following transplant. Mean diameter

values for each recovery group are included in the attached table with p-values (from

Student’s t-test) and values adjusted for FDR between transplant groups, representing

significance of difference between recovery groups for each transplant group. The mean

percent change (increase or decrease) to diameter following reperfusion is included at the

bottom of each table for both recovery groups in each transplant group.

Page 159: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

142

4.5.2 Diameter Results Stratified by Recovery Group (IGF vs. DGF)

In the SCD subgroup of DDKT kidneys stored by SCS, diameter measurements were

similar between IGF and DGF recovery groups (green and red respectively in Figure 4.3) prior to

implantation. In the ECD subgroup of DDKT kidneys stored by SCS, pre-implantation diameter

measurements were lower (p<0.05, FDR adjusted p-value=0.06) in the IGF than in the DGF

recovery group. In the SCD subgroup of DDKT kidneys stored by HMP, pre-implantation diameter

measurements were similar between IGF and DGF recovery groups. Following reperfusion,

diameter measurements for all recovery groups in all transplant groups increased. Within the SCD

subgroup of DDKT kidneys stored by SCS and the HMP group, increases were similar between

IGF and DGF recovery groups. In the ECD subgroup, diameter of the IGF recovery group

increased 10% while diameter in the DGF group increased 17%. Post-reperfusion diameter in the

SCD subgroup of kidneys stored by SCS was similar between IGF and DGF recovery groups.

Within the ECD subgroup, diameter in the IGF recovery group remained lower (p<0.005, FDR

adjusted p-value<0.005) than in the DGF group. In the HMP transplant group, IGF diameter was

moderately lower than in the DGF group (p=0.34).

4.5.3 Diameter Results by Association with Post-Transplant Creatinine Decline

Following our initial hypothesis that lower PCT lumen diameter would correlate with a faster

recovery following transplant (i.e. diameter is positively correlated with creatinine values and

lower diameter is correlated with a steeper decline in creatinine (i.e., has a negative interaction

effect with time)), linear mixed effect models were fitted for each DDKT transplant group. The

pre-implantation fitted model for the SCS-SCD group did not support the hypothesis (p=0.54),

Page 160: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

143

however the post-reperfusion SCS-SCD model did support the hypothesis (p<0.05). The pre-

implantation fitted model for the SCS-ECD group similarly did not support the hypothesis

(p=0.96), and the post-reperfusion SCS-ECD model did support the hypothesis (p<0.05). Finally,

the pre-implantation model for the HMP-SCD group did support the hypothesis (p<0.05), while

the post-reperfusion model did not (p=0.56).

4.6 Inter-Centroid Results

4.6.1 Inter-Centroid Results Stratified by Transplant Group (IGF and DGF Combined)

Prior to implantation (left in Figure 4.4), kidneys from the LDKT transplant group and

DDKT kidneys stored by SCS (both SCD and ECD) all exhibited a similar inter-centroid distance.

DDKT kidneys stored by HMP had a higher (p<0.05, FDR adjusted p-value<0.05) pre-

implantation inter-centroid distance than all other transplant groups. All groups experienced a

modest 1-4% increase in inter-centroid distance between pre-implantation and post-reperfusion

scans. Post-reperfusion (right in Figure 4.4) inter-centroid distance in the LDKT transplant group,

and DDKT-SCS subgroups was similar. Post-reperfusion inter-centroid distance in the HMP group

remained higher (p<0.005, FDR adjusted p-value=0.06) than the LDKT group and moderately

higher than the DDKT-SCS SCD and ECD subgroups (p=0.09/FDR adjusted p-value=0.14, and

p<0.05/FDR adjusted p-value=0.08 respectively).

Page 161: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

144

Figure 4.4: Box and whisker plots of inter-centroid measurements.

Measurements for pre-implantation (left) and post-reperfusion (right) scans

for the LDKT group, and the DDKT subgroups: SCD kidneys stored by SCS,

ECD kidneys stored by SCS, and SCD kidneys stored by HMP. Each

transplant group is further divided into recovery groups which experienced

either IGF (green) or DGF (red) following transplant. Mean inter-centroid

distance values for each recovery group are included in the attached table with

p-values (from Student’s t-test) and values adjusted for FDR between

transplant groups, representing significance of difference between recovery

groups for each transplant group. The mean percent change (increase or

decrease) to inter-centroid distance following reperfusion is included at the

bottom of each table for both recovery groups in each transplant group.

4.6.2 Inter-Centroid Results Stratified by Recovery Group (IGF vs. DGF)

Prior to implantation, inter-centroid distance was similar between the IGF and DGF

recovery groups in all transplant groups. Following reperfusion, inter-centroid distances increased

in all transplant groups for both IGF and DGF recovery groups. In the SCD subgroup of DDKT

kidneys stored by SCS, IGF and DGF recovery groups (green and red respectively in Figure 4.4)

experienced a similar increase following reperfusion. In the ECD subgroup of DDKT kidneys

stored by SCS, and in the SCS subgroup of DDKT kidneys stored by HMP, the IGF recovery

Page 162: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

145

groups experienced a smaller increase in inter-centroid distance following reperfusion than the

DGF groups. In the SCD subgroup of DDKT kidneys stored by SCS, post-reperfusion inter-

centroid distance measurements were similar between IGF and DGF groups. In the ECD subgroup,

inter-centroid distance was moderately lower (p=0.09) in the IGF recovery group than in the DGF

group. Post-reperfusion inter-centroid distance for the HMP group was lower (p<0.05) in the IGF

recovery group than in the DGF group.

4.6.3 Inter-Centroid Results by Association with Post-Transplant Creatinine Decline

Following our hypothesis that lower inter-centroid distance would correlate with a faster

recovery following transplant (i.e. inter-centroid distance is positively correlated with creatinine

values and lower inter-centroid distance is correlated with a steeper decline in creatinine (i.e., has

a negative interaction effect with time)), linear mixed effect models were fitted for each DDKT

transplant group. Both the pre-implantation and post-reperfusion fitted models for the SCS-SCD

group did not support the hypothesis (p=0.14, and p=0.17 respectively). Both the pre-implantation

and post-reperfusion fitted models for the SCS-ECD group did not support the hypothesis (p=0.28,

and p=0.72 respectively). Finally, the pre-implantation model for the HMP-SCD group did not

support the hypothesis (p=0.37), however the post-implantation model trended towards moderate

support of the hypothesis (p=0.07).

4.7 Inter-Lumen Results

4.7.1 Inter-Lumen Results Stratified by Transplant Group (IGF and DGF Combined)

Prior to implantation (left in Figure 4.5), the LDKT group exhibited larger (p<0.05, FDR

adjusted p-value=0.05) inter-lumen distance than the SCD and ECD subgroups of DDKT kidneys

Page 163: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

146

stored by SCS. The SCD subgroup of DDKT kidneys stored by HMP exhibited an inter-lumen

distance similar to the LDKT group. Following reperfusion, inter-lumen distance decreased

slightly in the LDKT transplant group, the SCD subgroup of DDKT kidneys stored by SCS, and

the SCD subgroup of DDKT kidneys stored by HMP. In the ECD subgroup of DDKT kidneys

stored by SCS, inter-lumen distance increased slightly following reperfusion. Post-reperfusion

(right in Figure 4.5) inter-lumen distance was higher (p<0.05, FDR adjusted p-value =<0.05) in

the LDKT transplant group than in the SCD subgroup of DDKT kidneys stored by SCS, and the

SCD subgroup of DDKT kidneys stored by HMP.

Figure 4.5: Box and whisker plots of inter-lumen measurements. Measurements

for pre-implantation (left) and post-reperfusion (right) scans for the LDKT group

(green), and the DDKT subgroups: SCD kidneys stored by SCS, ECD kidneys stored

by SCS, and SCD kidneys stored by HMP. Each transplant group is further divided

into recovery groups which experienced either IGF (green) or DGF (red) following

transplant. Mean inter-lumen distance values for each recovery group are included in

the attached table with p-values (from Student’s t-test) and values adjusted for FDR

between transplant groups, representing significance of difference between recovery

groups for each transplant group. The percent change (increase or decrease) to inter-

lumen distance following reperfusion is included at the bottom of each table for both

recovery groups in each transplant group.

Page 164: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

147

4.7.2 Inter-Lumen Results Stratified by Recovery Group (IGF vs. DGF)

Prior to implantation, inter-lumen distance was similar between the IGF and DGF recovery

groups in all transplant groups. Following reperfusion, inter-lumen distances in all transplant

groups decreased by less in the IGF recovery groups than in DGF groups (green and red

respectively in Figure 4.5). Post-reperfusion inter-lumen distance in the SCD subgroup of DDKT

kidneys stored by SCS was similar between IGF and DGF recovery groups. In the ECD subgroup,

post-reperfusion inter-lumen distance was moderately higher (p=0.06, FDR adjusted p-

value=0.09) in the IGF recovery group than in the DGF group. In the HMP group, post-reperfusion

inter-lumen distance was higher (p<0.05 FDR adjusted p-value<0.05) in the IGF recovery group

than in the DGF group.

4.7.3 Inter-Lumen Results by Association with Post-Transplant Creatinine Decline

Following our initial hypothesis that smaller inter-lumen distance would correlate with a

faster recovery following transplant (i.e. inter-lumen distance is negatively correlated with

creatinine values and higher inter-lumen distance is correlated with a steeper decline in creatinine

(i.e., has a negative interaction effect with time)), linear mixed effect models were fitted for each

DDKT transplant group. The pre-implantation fitted model for the SCS-SCD group did not support

the hypothesis (p=0.24), however the post-reperfusion SCS-SCD model showed strong support of

the hypothesis (p<0.001). The pre-implantation model for the SCS-ECD group did not support the

hypothesis (p=0.78), however the post-reperfusion model did support the hypothesis (p<0.05).

Finally, both the pre-implantation and post-reperfusion models for the HMP-SCD group showed

strong support for the hypothesis (p<0.0005, and p<0.005 respectively).

Page 165: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

148

4.8 Parsimony of Image Measurements

To assess relevance and redundancy of measurements, the compiled measurements from

each transplant group were included in the pool of candidate predictor variables in lasso penalized

regression models, with the post-transplant function (IGF coded as 1 vs. DGF coded as 0) as the

binary outcome variable. Two sets of penalized logistic regression models were run for each

transplant group: one included pre-implantation measurements only in the candidate pool to

identify the most relevant of pre-implantation measurements to post-transplant function (i.e.,

measurements which could affect allocation or discard), and the other included all the pre-

implantation and post-reperfusion measurements in the pool to determine the most relevant

measurements to post-transplant function (i.e., measurements which could affect post-operative

care). The number of selected measurements was determined by minimizing the averaged 3-fold

cross-validation error. Selected measurements and their impact are listed in Table 4.1.

Table 4.1: Measurements selected by lasso penalized regression modeling

as the most relevant to post-transplant function. Selected measurements

from only pre-implantation measurements (top), and from the combined pre-

implantation and post-reperfusion measurements (bottom) were selected.

Page 166: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

149

In the ECD subgroup of DDKT kidneys stored by SCS, the penalized model indicated pre-

implantation diameter was most relevant, among pre-implantation measurements, to post-

transplant function. Pre-implantation diameter had a negative impact on post-transplant function

in this instance, suggesting that larger lumen diameter is the most predictive of assessed

measurements for development of DGF in this transplant subgroup. When including both pre-

implantation and post-reperfusion measurements, the regression model indicated post-reperfusion

diameter and post-reperfusion density as the two variables, among all measurements, that were

most relevant to post-transplant function. Both have negative impact on the outcome, suggesting

that larger post-reperfusion lumen diameter and higher post-reperfusion lumen density are the most

predictive of assessed measurements for development of DGF in this transplant subgroup.

In the SCD subgroup of DDKT kidneys stored by SCS, the penalized model indicated pre-

implantation inter-centroid distance was most relevant, among pre-implantation measurements, to

post-transplant function. Pre-implantation inter-centroid distance had a negative impact on post-

transplant function in this instance, suggesting that larger inter-centroid distance is the most

predictive of assessed measurements for development of DGF in this transplant subgroup. When

including both pre-implantation and post-reperfusion measurements, the regression model

indicated pre-implantation inter-centroid distance and post-reperfusion density as the two

variables, among all measurements, that were most relevant to post-transplant function. Inter-

centroid distance and density had negative and positive impacts on outcome, respectively,

suggesting that larger pre-implantation inter-centroid distance and lower post-reperfusion lumen

density are the most predictive of assessed measurements for development of DGF in this

transplant subgroup.

Page 167: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

150

In the SCD subgroup of DDKT kidneys stored by HMP, the penalized model indicated pre-

implantation diameter was most relevant, among pre-implantation measurements, to post-

transplant function. Pre-implantation diameter had a negative impact on post-transplant function

in this instance, suggesting that larger diameter is the most predictive of assessed measurements

for development of DGF in this transplant subgroup. When including both pre-implantation and

post-reperfusion measurements, the regression model indicated post-reperfusion inter-lumen

distance and post-reperfusion density as the two variables, among all measurements, that were

most relevant to post-transplant function. Both have negative impact on the outcome, suggesting

that smaller post-reperfusion inter-lumen distance and lower post-reperfusion lumen density are

the most predictive of assessed measurements for development of DGF in this transplant subgroup.

4.9 Summary

Fibrosis in donor kidneys may compromise graft viability, and is routinely evaluated in

pre-implantation kidney biopsies [201]–[203]. Partial EMT may play a role in the progression of

fibrosis. This process has the effect of flattening PCT epithelial cells, and may produce an

increased lumen diameter in affected tubules [204], [205]. Similarly, fibrosis contributes to tubular

atrophy, and in turn, compensatory hypertrophy of surviving PCTs [206], [207]. The lumen of

hypertrophied tubules is also frequently dilated to accommodate their increased role [208]. The

effects of fibrosis therefore may be visible in OCT imaging, evidenced by the dilation of tubular

lumen.

Acute tubular injury (ATI) in donor kidneys may similarly compromise graft viability. ATI

can induce simplification of the tubular epithelium [62]. Shedding of the PCTs’ microvillus brush

border and sloughing of tubular epithelial cells into the lumen may also present as a dilation of the

Page 168: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

151

tubular lumen in OCT scans. In addition, as blood flow is restored following reperfusion, sloughed

epithelial cells may obstruct flow and increase proximal tubule pressure dramatically; heightened

pressure may produce substantial dilation of the tubular lumen presented in post-reperfusion OCT

scans and potentially pre-implantation OCT scans of kidneys preserved by HMP [209]. The short-

term effects of ATI therefor may be visible in OCT imaging, evidenced by the dilation of visible

tubular lumen.

Swelling of the PCT epithelium, induced by ischemic damage, may similarly represent the

effects or symptoms of ATI [62]. Epithelial swelling occludes the luminal space, resulting in a

reduced diameter and an increased inter-lumen distance. If PCT swelling reduces the tubular lumen

beyond the resolution of the OCT system, diameter and inter-lumen measurements would not

reflect the contribution of more swollen PCTs. Density measurements, however, would illustrate

this effect.

In the SCD subgroup of DDKT kidneys stored by SCS, there were no strong differences in

measurements between IGF and DGF recovery groups. In the ECD subgroup—those most at risk

for poor post-transplant function, and most subject to discard—measures of PCT lumen density

and diameter, acquired both prior to implantation and following reperfusion, were lower in the

IGF than in the DGF recovery group. The IGF recovery group similarly demonstrated a larger

inter-lumen distance measurement following reperfusion than the DGF group. Taken together,

these measurements suggest a flattening of the PCT epithelium and consequent dilation of tubular

lumen in ECD kidneys which go on to experience DGF. This may be a symptom of pre-existing

pathology (fibrosis) or ATI. It is unclear why this pattern does not present in the SCD subgroup.

Following reperfusion, density and diameter measurements in both the SCD and ECD

subgroups of DDKT kidneys stored by SCS experienced increases in both IGF and DGF recovery

Page 169: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

152

groups. This may reflect dissipation of epithelial swelling as the kidney moves away from an

ischemic state. This may also result from the effect of flow rate of filtrate on luminal diameter

[210]. Increased distinction between IGF and DGF recovery group measurements following

reperfusion may be due to pre-existing pathology being revealed by the dissipation of swelling

(e.g. dilated lumen of hypertrophied tubules may become more evident when epithelial swelling

subsides). More likely, this is the result of the reperfusion process inducing further shedding of the

microvillus brush border and/or further epithelial sloughing. Similarly, sloughed tubular epithelial

cells which may have fully occluded the lumen during static-storage may be cleared following

reperfusion, revealing further luminal dilation.

In the ECD subgroup, but not the SCD subgroup, of DDKT kidneys stored by SCS, the

DGF recovery group experienced an increase in inter-centroid distance following reperfusion,

while the IGF group did not. This may reflect infiltration of inflammatory cells into the interstitial

space, and subsequent interstitial edema [211]. This would be consistent with the ATI theory and

would suggest symptoms of IRI in the DGF group.

In the SCD subgroup of DDKT kidneys stored by HMP, diameter, and inter-lumen

measurements for DGF kidneys echo the trends apparent in the ECD subgroup of DDKT kidneys

stored by SCS (i.e. increased lumen diameter and reduced inter-lumen distance). This suggests

that, in HMP preserved kidneys, ATI or pre-existing pathology may also present as dilated tubular

lumen with simplified or flattened tubular epithelium. Inter-centroid measurements similarly echo

trends apparent in the ECD transplant group. Following reperfusion, the DGF recovery group

experienced an increase in inter-centroid distance and subsequently exhibited a higher inter-

centroid measurement than the IGF recovery group. This again may suggest interstitial edema

following reperfusion.

Page 170: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

153

Surprisingly, HMP kidneys in the DGF recovery group experienced a dramatic reduction

in density following reperfusion, while the IGF group experienced little change. The resulting IGF

density was higher than the density in the DGF group. Higher diameter and lower inter-lumen

distances in the post-reperfusion DGF group would normally correlate with higher density

measurements. One explanation for this contradictory result is that some PCT lumens in the HMP-

DGF group had become fully occluded following reperfusion, excluding these PCTs from diameter

and inter-lumen measurement, but still detracting from luminal area in the density measurement.

Page 171: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

154

CONCLUSION

There is a dire need in the transplant community for new measures of kidney viability. To

support the growing need for kidneys, higher risk kidneys must be considered for transplant. To

efficiently utilize this deeper end of the donor pool, surgeons must be able to confidently predict

kidneys’ potential function and longevity following transplant.

OCT provides a non-invasive method for obtaining optical cross-sections of the superficial

kidney cortex [212], [213]. These images reveal the microanatomy of the PCTs, which comprise

the majority of the superficial kidney cortex. Swelling of the epithelium of the PCTs may be

identified in OCT images by reduction in visible tubular lumen area, and may be considered a

symptom of ischemic insult [199], [214]; as PCT epithelium swells, lumen space is occluded.

Conversely, dilation of the tubular lumen may represent tubular simplification as a symptom of

ATI (i.e., shedding of the PCT microvillus brush border and/or epithelial sloughing); loss of

epithelial cells or microvilli should increase lumen space. Alternatively, dilation of the tubular

lumen could be considered a symptom of pre-existing pathology (i.e., as partial EMT, hypertrophy,

or potentially tubular atrophy).

Quantification of the degree of swelling, simplification, or fibrotic symptoms in OCT

images may provide a valuable addition to current measures of kidney viability. If the degree of

ischemic damage could be accurately determined prior to transplant from quantification of

swelling or tubular simplification, the degree of IRI which would ensue following reperfusion

could potentially be ascertained. An accurate measure and quantification of the accumulation of

factors from ischemic damage that would contribute to IRI could enable a widening of the donor

Page 172: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

155

pool through inclusion of kidneys which based purely on the duration of cold and warm ischemia

would normally be discarded for an assumed degree of ischemic damage but which may still be

viable due to an increased resistance to ischemic insult. Quantification of fibrotic dilation in

OCT imaging of kidneys may similarly provide a valuable addition to current viability measures.

Biopsy procedures are not only invasive, but localized and destructive to the biopsied sample.

OCT’s non-invasive nature affords it the opportunity to visualize fibrotic symptoms throughout

the kidney, and without damage to the tissue being directly investigated. Identification by OCT

of pre-existing pathology, or ischemic damage/ATI prior to transplant, may inform transplant

surgeons’ decision to accept or reject a kidney for transplant, or may affect allocation of the

graft. In this regard, OCT may supplement or guide biopsies and offer a more global view of the

pathology and its distribution.

OCT also has potential utility following transplant, where a more accurate prediction of

post-transplant function could influence post-operative care. DGF is an established risk factor for

survival of a transplanted kidney [215]. If DGF can be predicted immediately following

transplant, early post-operative biopsies to investigate poor function can be avoided. Early

diagnosis of DGF can similarly inform the development of immunosuppressive treatments,

where evidence of potential DGF would provide indications for a less nephrotoxic, Calcineurin-

sparing regiment [216]. An accurate prediction of DGF would also promote the usage of any of a

number of anti-DGF medications currently in development, should they be approved.

OCT imaging performed following transplant has the added benefit of providing a view

of kidney microanatomy following dissipation of swelling, following clearance of debris,

following reperfusion-induced damage, and under the normal pressure of luminal flow.

Page 173: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

156

Dissipation of swelling can provide a benchmark from which to grade the degree of swelling

seen prior to implantation. The dissipation of swelling can also reveal tubular dilation associated

with fibrotic symptoms which may have been masked by luminal occlusion by swollen

epithelium. Reperfusion may clear debris which has occluded the luminal space, again revealing

potential fibrotic dilation, and also revealing tubular simplification which may have been

disguised by debris occlusion of the lumen (i.e., sloughed epithelial cells and microvilli in static

lumen may increase the refractive index of the luminal space, preventing the lumen from being

discriminable from cortex in OCT imaging). The reperfusion process may induce further tubular

simplification as a product of IRI, leading to further tubular dilation following transplant.

Finally, the intra-tubular pressure from flow through the lumen following reperfusion may reveal

patterns not seen in the static kidney; fibrotic tubules, for example, may be less flexible to the

effects of flow rate and so respond differently to the increased flow.

This study shows that dilation of tubular lumen and simplification of tubular epithelium

of the PCTs can be assessed by OCT, and that these measurements correlate with post-transplant

function in some transplant groups. These factors may represent symptoms of pre-existing

pathology or ATI. The variability between manual raters in this study demonstrates the necessity

for consistency and reproducibility in analysis. The fully automated analysis pipeline presented

in this thesis and used in this study removes the elements of user bias and subjective

segmentation. Similarly, manual segmentation is considerably too slow a process when advising

a surgeon on the time-sensitive decision to accept or reject a kidney for transplant. Fully

automated segmentation and analysis provides a high-speed solution to obtaining accurate

predictive measures.

Page 174: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

157

This study assessed the potential utility of OCT imaging in predicting post-transplant

function. While results are promising, inclusion of additional variables (KDPI, ischemic times,

biopsy scoring, etc.) into one prediction model may provide a more comprehensive view of kidney

viability. Similarly, global OCT imaging and capture of 3D volumes would provide a more

detailed view of the distribution of PCT morphology, and may aid in prediction of post-transplant

function. 3D volumes would similarly enable adoption of previously developed OCT segmentation

strategies, for example the Hessian filter approach by Yousefi et al. and single-scattering model

with segment-joining algorithm by Gong et al. [217], [218].

One limitation of this study is the imaging protocol, which heavily weighted the

composition of image sets towards regions of the kidney where tubule lumens were most visible

and dilated. While this protocol may highlight focal points of pathology, it does not provide a

global distribution of PCT features. Global imaging sampling multiple areas of the kidney may

reveal a more heterogeneous pattern of swelling and dilation, with some areas exhibiting tubular

lumen dilated by fibrosis or ATI, and other areas exhibiting significant swelling.

In future studies, a more systematic and global imaging strategy may yield further

insights. While the selection of a single B-scan for each image set removes issues of redundancy,

it also severely limits the total area being investigated. In future studies, a 3D imaging protocol

would eliminate this issue, allowing all imaging data to be evaluated and a larger volume of

kidney to be assessed. Similarly, 3D imaging would enable orientation of tubular features in a

3D space and would provide more accurate measurements. While the linear regression model

utilized in this study attempts to correct for this issue, training data for the model is extracted

only from a single preserved kidney and may not be applicable to all kidneys.

Page 175: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

158

BIBLIOGRAPHY

[1] N. Perico, P. Ruggenenti, M. Scalamogna, and G. Remuzzi, “Tackling the Shortage of

Donor Kidneys: How to Use the Best that We Have,” Am. J. Nephrol., vol. 23, no. 4, pp.

245–259, 2003.

[2] A. Hart et al., “OPTN/SRTR 2016 Annual Data Report: Kidney,” Am. J. Transplant., vol.

18, pp. 18–113, Jan. 2018.

[3] A. O. Ojo et al., “Survival in recipients of marginal cadaveric donor kidneys compared

with other recipients and wait-listed transplant candidates.,” J. Am. Soc. Nephrol., vol. 12,

no. 3, pp. 589–97, Mar. 2001.

[4] P. Nagaraja et al., “Impact of Expanded Criteria Variables on Outcomes of Kidney

Transplantation from Donors After Cardiac Death,” Transplantation, vol. 99, no. 1, pp.

226–231, Jan. 2015.

[5] J. Pascual, J. Zamora, and J. D. Pirsch, “A Systematic Review of Kidney Transplantation

From Expanded Criteria Donors,” Am. J. Kidney Dis., vol. 52, no. 3, pp. 553–586, Sep.

2008.

[6] Y. Tomita, T. Tojimbara, K. Iwadoh, I. Nakajima, and S. Fuchinoue, “Long-Term

Outcomes in Kidney Transplantation From Expanded-Criteria Donors After Circulatory

Death,” Transplant. Proc., vol. 49, no. 1, pp. 45–48, Jan. 2017.

[7] S. Bae, A. B. Massie, X. Luo, S. Anjum, N. M. Desai, and D. L. Segev, “Changes in

Discard Rate After the Introduction of the Kidney Donor Profile Index (KDPI),” Am. J.

Transplant., vol. 16, no. 7, pp. 2202–2207, Jul. 2016.

[8] A. O. Ojo et al., “Survival in recipients of marginal cadaveric donor kidneys compared

with other recipients and wait-listed transplant candidates.,” J. Am. Soc. Nephrol., vol. 12,

no. 3, pp. 589–97, Mar. 2001.

[9] M. G. Snoeijs et al., “Kidneys from donors after cardiac death provide survival benefit.,”

J. Am. Soc. Nephrol., vol. 21, no. 6, pp. 1015–21, Jun. 2010.

[10] R. L. Heilman, E. P. Green, K. S. Reddy, A. Moss, and B. Kaplan, “Potential Impact of

Risk and Loss Aversion on the Process of Accepting Kidneys for Transplantation.,”

Transplantation, vol. 101, no. 7, pp. 1514–1517, Jul. 2017.

[11] M. D. Doshi et al., “Utility of Applying Quality Assessment Tools for Kidneys With

KDPI ≥80.,” Transplantation, vol. 101, no. 6, pp. 1125–1133, Jun. 2017.

[12] G. J. Tortora and B. Derrickson, Principles of anatomy and physiology. Hoboken, NJ :

John Wiley & Sons, 2010.

[13] A. a McDonough, “Mechanisms of proximal tubule sodium transport regulation that link

extracellular fluid volume and blood pressure.,” Am. J. Physiol. Regul. Integr. Comp.

Physiol., vol. 298, no. 4, pp. R851–R861, 2010.

[14] A. Leaf, “Cell swelling: a factor in ischemic tissue injury,” Circulation, vol. 48, no. 3, pp.

455–458, 1973.

[15] M. J. Taylor, “Biology of Cell Survival in the Cold,” in Advances in Biopreservation,

Page 176: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

159

2006, pp. 15–62.

[16] J. L. Zhuo and X. C. Li, “Proximal Nephron,” Compr. Physiol., vol. 3, no. 3, pp. 1079–

1123, 2013.

[17] M. Ikeda, W. Prachasilchai, M. J. Burne-Taney, H. Rabb, and N. Yokota-Ikeda, “Ischemic

acute tubular necrosis models and drug discovery: a focus on cellular inflammation,” Drug

Discovery Today. 2006.

[18] B. J. McNamara, B. Diouf, R. N. Douglas-Denton, M. D. Hughson, W. E. Hoy, and J. F.

Bertram, “A comparison of nephron number, glomerular volume and kidney weight in

Senegalese Africans and African Americans,” Nephrol. Dial. Transplant., 2010.

[19] L. J. Mandel, “Metabolic substrates, cellular energy production, and the regulation of

proximal tubular transport,” Annu. Rev. Physiol., vol. 47, no. 1, pp. 85–101, 1985.

[20] B. D. Ross, J. Espinal, and P. Silva, “Glucose metabolism in renal tubular function,”

Kidney Int., vol. 29, pp. 54–67, 1986.

[21] D. Bon, Chatauret, S. Guraud, and T. Hauet, “New strategies to optimize kidney recovery

and preservation in transplantation,” Nephrology, vol. 8, pp. 339–347, 2012.

[22] K. K. Tennankore, S. J. Kim, I. P. J. Alwayn, and B. A. Kiberd, “Prolonged warm

ischemia time is associated with graft failure and mortality after kidney transplantation.,”

Kidney Int., vol. 89, no. 3, pp. 648–58, Mar. 2016.

[23] J. Gill, C. Rose, J. Lesage, Y. Joffres, J. Gill, and K. O’Connor, “Use and Outcomes of

Kidneys from Donation after Circulatory Death Donors in the United States,” J. Am. Soc.

Nephrol., vol. 28, no. 12, pp. 3647–3657, Dec. 2017.

[24] F. O. Belzer, B. S. Ashby, J. S. Huang, and J. E. Dunphy, “Etiology of rising perfusion

pressure in isolated organ perfusion.,” Ann. Surg., 1968.

[25] D. Butter G, A. Saunder, D. Marsh, and J. Southard, “Comparison of solutions for

preservation of the rabbit liver as tested by isolated perfusion,” Transpl. Int., vol. 8, no. 6,

pp. 466–471, 1995.

[26] C. Y. Lee and M. J. Mangino, “Preservation methods for kidney and liver,”

Organogenesis. 2009.

[27] N. Ahmad, L. Hostert, J. R. Pratt, K. J. Billar, D. J. Potts, and J. P. A. Lodge, “A

pathophysiologic study of the kidney tubule to optimize organ preservation solutions,”

Kidney Int., 2004.

[28] J. J. Grantham, C. M. Lowe, M. Dellasega, and B. R. Cole, “Effect of hypotonic medium

on K and Na content of proximal renal tubules,” Am. J. Physiol. Physiol., 1977.

[29] A. Sardini, J. S. Amey, K.-H. Weylandt, M. Nobles, M. A. Valverde, and C. F. Higgins,

“Cell volume regulation and swelling-activated chloride channels,” Biochim. Biophys.

Acta (BBA)-Biomembranes, vol. 1618, no. 2, pp. 153–162, 2003.

[30] R. F. Parsons and J. V. Guarrera, “Preservation solutions for static cold storage of

abdominal allografts: which is best?,” Transplant. Curr. Opin., vol. 19, no. 2, pp. 100–107,

2014.

Page 177: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

160

[31] A. Debout et al., “Each additional hour of cold ischemia time significantly increases the

risk of graft failure and mortality following renal transplantation,” Kidney Int., vol. 87, no.

2, pp. 343–349, Feb. 2015.

[32] E. R. Pieter Hoogland, T. C. van Smaalen, M. H. L. Christiaans, and L. W. E. van Heurn,

“Kidneys from uncontrolled donors after cardiac death: which kidneys do worse?,”

Transpl. Int., vol. 26, no. 5, pp. 477–484, May 2013.

[33] N. O. S. Camara, W. W. Williams Jr, and A. Pacheco-Silva, “Proximal tubular

dysfunction as an indicator of chronic graft dysfunction,” Brazilian J. Med. Biol. Res., vol.

42, pp. 229–236, 2009.

[34] M. A. Venkatachalam, D. B. Bernard, J. F. Donohoe, and N. G. Levinsky, “Ischemic

damage and repair in the rat proximal tubule: Differences among the S1, S2 and S3

segments,” Kidney Int., vol. 14, pp. 31–49, 1978.

[35] J. M. Weinberg, M. A. Venkatachalam, N. Roeser, and I. Nissim, “Mitochondrial

dysfunction during hypoxia/reoxygenation and its correction by anaerobic metabolism of

citric acid cycle intermediates,” PNAS, vol. 97, no. 6, pp. 2826–2831, 2000.

[36] D. J. Parekh et al., “Tolerance of the Human Kidney to Isolated Controlled Ischemia,” J.

Am. Soc. Nephrol., vol. 24, no. 3, pp. 506–517, Mar. 2013.

[37] J. M. Weinberg et al., “Anaerobic and aerobic pathways for salvage of proximal tubules

from hypoxia-induced mitochondrial injury.,” Am. J. Physiol. Renal Physiol., 2000.

[38] D. K. Das, “Metabolic and morphological alterations induced by ischemia and reperfusion

during cardiac surgery,” Pathophysiol. Reperfus. Inj., pp. 265–267, 1993.

[39] C. A. Schumacher, A. Baartscheer, R. Coronel, and J. W. T. Fiolet, “Energy-dependent

transport of calcium to the extracellular space during acute ischemia of the rat heart,” J.

Mol. Cell. Cardiol., 1998.

[40] E. K. Hoffmann and L. O. Simonsen, “Membrane mechanisms in volume and pH

regulation in vertebrate cells,” Physiol. Rev., vol. 69, no. 2, pp. 315–382, Apr. 1989.

[41] M. Büsing, U. T. Hopt, W. Schareck, M. Ernst, and K. Morgenroth, “Ultrastructural

changes of different preserved human kidney allografts before and after reperfusion.,”

Transplant. Proc., vol. 22, no. 2, pp. 448–9, Apr. 1990.

[42] P. M. Andrews, B. S. Khirabadi, and B. C. Bengs, “Using tandem scanning confocal

microscopy to predict the status of donor kidneys.,” Nephron, vol. 91, no. 1, pp. 148–55,

May 2002.

[43] N. Chatauret, L. Badet, B. Barrou, and T. Hauet, “Ischemia-reperfusion: From cell biology

to acute kidney injury,” Progrès en Urol., vol. 24, pp. S4–S12, Jun. 2014.

[44] M. Kosieradzki and W. Rowiński, “Ischemia/Reperfusion Injury in Kidney

Transplantation: Mechanisms and Prevention,” Transplant. Proc., vol. 40, no. 10, pp.

3279–3288, Dec. 2008.

[45] S. K. Malek, “Procurement biopsies in kidneys retrieved for transplantation.,” Clin. J. Am.

Soc. Nephrol., vol. 9, no. 3, pp. 443–4, Mar. 2014.

Page 178: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

161

[46] D. Patschan, S. Patschan, and G. A. Müller, “Inflammation and Microvasculopathy in

Renal Ischemia Reperfusion Injury,” J. Transplant., 2012.

[47] J. M. Thurman, “Triggers of inflammation after renal ischemia/reperfusion,” Clinical

Immunology. 2007.

[48] C. Li and R. M. Jackson, “Reactive species mechanisms of cellular hypoxia-

reoxygenation injury,” AJP Cell Physiol., 2002.

[49] G. T. Vatassery, “Impairment of brain mitochondrial oxidative phosphorylation

accompanying vitamin E oxidation induced by iron or reactive nitrogen species: A

selective review,” Neurochemical Research. 2004.

[50] J. A. Lee and D. G. Allen, “Changes in intracellular free calcium concentration during

long exposures to simulated ischemia in isolated mammalian ventricular muscle,” Circ.

Res., 1992.

[51] S. Javadov, J. C. Hunter, G. Barreto-Torres, and R. Parodi-Rullan, “Targeting the

mitochondrial permeability transition: Cardiac ischemia-reperfusion versus

carcinogenesis,” Cellular Physiology and Biochemistry. 2011.

[52] S. G. Kasimsetty and D. B. McKay, “Ischemia as a factor affecting innate immune

responses in kidney transplantation,” Current Opinion in Nephrology and Hypertension.

2016.

[53] E. Lechevallier et al., “Posttransplantation acute tubular necrosis: risk factors and

implications for graft survival.,” Am. J. Kidney Dis., vol. 32, no. 6, pp. 984–91, Dec.

1998.

[54] G. Gobé, D. Willgoss, N. Hogg, E. Schoch, and Z. Endre, “Cell survival or death in renal

tubular epithelium after ischemia-reperfusion injury.,” Kidney Int., vol. 56, no. 4, pp.

1299–304, Oct. 1999.

[55] B. Schröppel and C. Legendre, “Delayed kidney graft function: from mechanism to

translation.,” Kidney Int., vol. 86, no. 2, pp. 251–8, Aug. 2014.

[56] L. W. Gaber, L. W. Moore, R. R. Alloway, M. H. Amiri, S. R. Vera, and A. O. Gaber,

“Glomerulosclerosis as a determinant of posttransplant function of older donor renal

allografts.,” Transplantation, vol. 60, no. 4, pp. 334–9, Aug. 1995.

[57] J. Dominguez et al., “Benefits of Early Biopsy on the Outcome of Kidney

Transplantation,” Transplant. Proc., vol. 37, no. 8, pp. 3361–3363, Oct. 2005.

[58] G. W. Moeckel, “Pathologic Perspectives on Acute Tubular Injury Assessment in the

Kidney Biopsy,” Semin. Nephrol., vol. 38, no. 1, pp. 21–30, Jan. 2018.

[59] a M. Sheridan and J. V Bonventre, “Cell biology and molecular mechanisms of injury in

ischemic acute renal failure.,” Curr. Opin. Nephrol. Hypertens., 2000.

[60] E. M. Ongeri, O. Anyanwu, W. B. Reeves, and J. S. Bond, “Villin and actin in the mouse

kidney brush-border membrane bind to and are degraded by meprins, an interaction that

contributes to injury in ischemia-reperfusion,” AJP Ren. Physiol., 2011.

[61] B. A. Molitoris, A. Geerdes, and J. R. McIntosh, “Dissociation and redistribution of

Page 179: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

162

Na+,K+-ATPase from its surface membrane actin cytoskeletal complex during cellular

ATP depletion,” J. Clin. Invest., 1991.

[62] G. W. Moeckel, “Pathologic Perspectives on Acute Tubular Injury Assessment in the

Kidney Biopsy,” Seminars in Nephrology. 2018.

[63] P. Boor, T. Ostendorf, and J. Floege, “Renal fibrosis: novel insights into mechanisms and

therapeutic targets,” Nat. Rev. Nephrol., vol. 6, no. 11, pp. 643–656, Nov. 2010.

[64] L. C. Racusen, K. Solez, and R. Colvin, “Fibrosis and atrophy in the renal allograft:

interim report and new directions.,” Am. J. Transplant, vol. 2, no. 3, pp. 203–6, Mar.

2002.

[65] R. A. Risdon, J. C. Sloper, and H. E. De Wardener, “Relationship between renal function

and histological changes found in renal-biopsy specimens from patients with persistent

glomerular nephritis.,” Lancet (London, England), vol. 2, no. 7564, pp. 363–6, Aug. 1968.

[66] B. Kaissling and M. Hir, “The renal cortical interstitium: morphological and functional

aspects,” Histochem. Cell Biol., vol. 130, no. 2, pp. 247–262, Aug. 2008.

[67] A. B. Farris and R. B. Colvin, “Renal interstitial fibrosis: mechanisms and evaluation.,”

Curr. Opin. Nephrol. Hypertens., vol. 21, no. 3, pp. 289–300, May 2012.

[68] R. B. Colvin, “Eye of the needle.,” Am. J. Transplant, vol. 7, no. 2, pp. 267–8, Feb. 2007.

[69] Y. Ovadya and V. Krizhanovsky, “A new Twist in kidney fibrosis,” Nat. Med., vol. 21,

no. 9, pp. 975–977, Sep. 2015.

[70] S. Lovisa, M. Zeisberg, and R. Kalluri, “Partial Epithelial-to-Mesenchymal Transition and

Other New Mechanisms of Kidney Fibrosis.,” Trends Endocrinol. Metab., vol. 27, no. 10,

pp. 681–695, Oct. 2016.

[71] M. L. Nicholson, E. Bailey, S. Williams, K. P. Harris, and P. N. Furness, “Computerized

histomorphometric assessment of protocol renal transplant biopsy specimens for surrogate

markers of chronic rejection.,” Transplantation, vol. 68, no. 2, pp. 236–41, Jul. 1999.

[72] W. Kriz and M. Lehir, “Pathways to nephron loss starting from glomerular diseases—

Insights from animal models,” Kidney Int., vol. 67, no. 2, pp. 404–419, Feb. 2005.

[73] R. L. Chevalier, “The proximal tubule is the primary target of injury and progression of

kidney disease: role of the glomerulotubular junction.,” Am. J. Physiol. Renal Physiol.,

vol. 311, no. 1, pp. F145-61, 2016.

[74] A. Servais et al., “Interstitial fibrosis evolution on early sequential screening renal

allograft biopsies using quantitative image analysis.,” Am. J. Transplant, vol. 11, no. 7,

pp. 1456–63, Jul. 2011.

[75] S. Bunnag et al., “Molecular correlates of renal function in kidney transplant biopsies.,” J.

Am. Soc. Nephrol., vol. 20, no. 5, pp. 1149–60, May 2009.

[76] D. Serón et al., “Protocol renal allograft biopsies and the design of clinical trials aimed to

prevent or treat chronic allograft nephropathy.,” Transplantation, vol. 69, no. 9, pp. 1849–

55, May 2000.

[77] P. Cravedi and P. S. Heeger, “Complement as a multifaceted modulator of kidney

Page 180: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

163

transplant injury,” J. Clin. Invest., vol. 124, no. 6, pp. 2348–2354, Jun. 2014.

[78] J. Floege et al., “Age-related glomerulosclerosis and interstitial fibrosis in Milan

normotensive rats: a podocyte disease.,” Kidney Int., vol. 51, no. 1, pp. 230–43, Jan. 1997.

[79] A. B. Fogo, “Causes and pathogenesis of focal segmental glomerulosclerosis,” Nat. Rev.

Nephrol., vol. 11, no. 2, pp. 76–87, Feb. 2015.

[80] E. Ważna et al., “Arteriolar Hyalinization in Implantation Kidney Biopsies as a Predictor

of Graft Function,” Transplant. Proc., vol. 41, no. 8, pp. 2975–2977, Oct. 2009.

[81] A. T. Woestenburg, G. A. Verpooten, D. K. Ysebaert, E. A. Van Marck, D. Verbeelen,

and J.-L. Bosmans, “Fibrous Intimal Thickening at Implantation Adversely Affects Long-

Term Kidney Allograft Function,” Transplantation, vol. 87, no. 1, pp. 72–78, Jan. 2009.

[82] C. M. Burchfiel, R. E. Tracy, P. H. Chyou, and J. P. Strong, “Cardiovascular risk factors

and hyalinization of renal arterioles at autopsy. The Honolulu Heart Program.,”

Arterioscler. Thromb. Vasc. Biol., vol. 17, no. 4, pp. 760–8, Apr. 1997.

[83] M. Matignon et al., “Arteriolar hyalinization predicts delayed graft function in deceased

donor renal transplantation.,” Transplantation, vol. 86, no. 7, pp. 1002–5, Oct. 2008.

[84] K. Kohagura et al., “An association between uric acid levels and renal arteriolopathy in

chronic kidney disease: a biopsy-based study,” Hypertens. Res., vol. 36, no. 1, pp. 43–49,

Jan. 2013.

[85] C. Kaplan, B. Pasternack, H. Shah, and G. Gallo, “Age-related incidence of sclerotic

glomeruli in human kidneys.,” Am. J. Pathol., vol. 80, no. 2, pp. 227–34, Aug. 1975.

[86] M. S. Hommos, R. J. Glassock, and A. D. Rule, “Structural and Functional Changes in

Human Kidneys with Healthy Aging,” J. Am. Soc. Nephrol., vol. 28, no. 10, pp. 2838–

2844, Oct. 2017.

[87] L. W. Gaber, L. W. Moore, R. R. Alloway, M. H. Amiri, S. R. Vera, and A. O. Gaber,

“Glomerulosclerosis as a determinant of posttransplant function of older donor renal

allografts.,” Transplantation, vol. 60, no. 4, pp. 334–9, Aug. 1995.

[88] C. J. Wang, J. B. Wetmore, G. S. Crary, and B. L. Kasiske, “The Donor Kidney Biopsy

and Its Implications in Predicting Graft Outcomes: A Systematic Review,” Am. J.

Transplant., vol. 15, no. 7, pp. 1903–1914, Jul. 2015.

[89] I. Jochmans and J. Pirenne, “Graft quality assessment in kidney transplantation: not an

exact science yet!,” Curr. Opin. Organ Transplant., vol. 16, no. 2, pp. 174–179, Apr. 2011.

[90] A. B. Fogo, “Causes and pathogenesis of focal segmental glomerulosclerosis,” Nat. Rev.

Nephrol., vol. 11, no. 2, pp. 76–87, Feb. 2015.

[91] V. D. D’Agati, A. B. Fogo, J. A. Bruijn, and J. C. Jennette, “Pathologic classification of

focal segmental glomerulosclerosis: a working proposal.,” Am. J. Kidney Dis., vol. 43, no.

2, pp. 368–82, Feb. 2004.

[92] M. D. Hughson, K. Johnson, R. J. Young, W. E. Hoy, and J. F. Bertram, “Glomerular size

and glomerulosclerosis: relationships to disease categories, glomerular solidification, and

ischemic obsolescence.,” Am. J. Kidney Dis., vol. 39, no. 4, pp. 679–88, Apr. 2002.

Page 181: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

164

[93] A. K. Israni et al., “New National Allocation Policy for Deceased Donor Kidneys in the

United States and Possible Effect on Patient Outcomes,” J. Am. Soc. Nephrol., vol. 25, no.

8, pp. 1842–1848, Aug. 2014.

[94] R. S. Sung et al., “Determinants of Discard of Expanded Criteria Donor Kidneys: Impact

of Biopsy and Machine Perfusion,” Am. J. Transplant., vol. 8, no. 4, pp. 783–792, Apr.

2008.

[95] E. Tavares da Silva et al., “Pretransplant Biopsy in Expanded Criteria Donors: Do We

Really Need It?,” Transplant. Proc., vol. 46, no. 10, pp. 3330–3334, Dec. 2014.

[96] M. Naesens, “Zero-Time Renal Transplant Biopsies,” Transplantation, vol. 100, no. 7, pp.

1425–1439, Jul. 2016.

[97] B. J. Nankivell and J. R. Chapman, “The Significance of Subclinical Rejection and the

Value of Protocol Biopsies,” Am. J. Transplant., vol. 6, no. 9, pp. 2006–2012, Sep. 2006.

[98] M. Mengel et al., “Protocol Biopsies in Renal Transplantation: Insights into Patient

Management and Pathogenesis,” Am. J. Transplant., vol. 7, no. 3, pp. 512–517, Mar.

2007.

[99] K. Solez et al., “Banff 07 Classification of Renal Allograft Pathology: Updates and Future

Directions,” Am. J. Transplant., vol. 8, no. 4, pp. 753–760, Apr. 2008.

[100] D. D. H. Koo, K. I. Welsh, J. A. Roake, P. J. Morris, and S. V. Fuggle,

“Ischemia/Reperfusion Injury in Human Kidney Transplantation,” Am. J. Pathol., vol.

153, no. 2, pp. 557–566, Aug. 1998.

[101] M. Haas, L. E. Ratner, and R. A. Montgomery, “C4d staining of perioperative renal

transplant biopsies.,” Transplantation, vol. 74, no. 5, pp. 711–7, Sep. 2002.

[102] J. Kanellis et al., “JNK signalling in human and experimental renal ischaemia/reperfusion

injury,” Nephrol. Dial. Transplant., vol. 25, no. 9, pp. 2898–2908, Sep. 2010.

[103] G. Remuzzi et al., “Early experience with dual kidney transplantation in adults using

expanded donor criteria. Double Kidney Transplant Group (DKG).,” J. Am. Soc.

Nephrol., vol. 10, no. 12, pp. 2591–8, Dec. 1999.

[104] R. B. Munivenkatappa et al., “The Maryland Aggregate Pathology Index: A Deceased

Donor Kidney Biopsy Scoring System for Predicting Graft Failure,” Am. J. Transplant.,

vol. 8, no. 11, pp. 2316–2324, Nov. 2008.

[105] L. C. Racusen et al., “The Banff 97 working classification of renal allograft pathology,”

Kidney Int., vol. 55, no. 2, pp. 713–723, Feb. 1999.

[106] D. M. Bhowmik, A. K. Dinda, P. Mahanta, and S. K. Agarwal, “The evolution of the

Banff classification schema for diagnosing renal allograft rejection and its implications for

clinicians.,” Indian J. Nephrol., vol. 20, no. 1, pp. 2–8, Jan. 2010.

[107] H. Liapis et al., “Banff Histopathological Consensus Criteria for Preimplantation Kidney

Biopsies,” Am. J. Transplant., vol. 17, no. 1, pp. 140–150, Jan. 2017.

[108] A. E. Suttie, J. R. Leininger, A. E. Bradley, and G. A. Boorman, Boorman’s pathology of

the rat : reference and atlas. .

Page 182: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

165

[109] S. P. Laucks and M. S. F. McLachlan, “Aging and simple cysts of the kidney,” Br. J.

Radiol., vol. 54, no. 637, pp. 12–14, Jan. 1981.

[110] P. Veroux et al., “Living transplantation using a kidney with a large cyst as curative

treatment of donor’s hypertension,” Nephrol. Dial. Transplant., vol. 17, no. 12, pp. 2258–

2260, Dec. 2002.

[111] D. Grotemeyer et al., “Renal Cysts in Living Donor Kidney Transplantation: Long-Term

Follow-up in 25 Patients,” Transplant. Proc., vol. 41, no. 10, pp. 4047–4051, Dec. 2009.

[112] H. Myrvang, “Native renal cysts linked to RCC in transplanted kidneys,” Nat. Rev.

Nephrol., vol. 7, no. 3, pp. 125–125, Mar. 2011.

[113] S. M. Flechner, J. Kobashigawa, and G. Klintmalm, “Calcineurin inhibitor-sparing

regimens in solid organ transplantation: focus on improving renal function and

nephrotoxicity.,” Clin. Transplant., vol. 22, no. 1, pp. 1–15, Jun. 2007.

[114] “Introduction to OCT - OBEL.” [Online]. Available:

http://obel.ee.uwa.edu.au/research/fundamentals/introduction-oct/. [Accessed: 06-Mar-

2019].

[115] K. Hueper et al., “T2 relaxation time and apparent diffusion coefficient for noninvasive

assessment of renal pathology after acute kidney injury in mice: Comparison with

histopathology,” Invest. Radiol., 2013.

[116] P. Steiger, S. Barbieri, A. Kruse, M. Ith, and H. C. Thoeny, “Selection for biopsy of

kidney transplant patients by diffusion-weighted MRI,” Eur. Radiol., vol. 27, no. 10, pp.

4336–4344, Oct. 2017.

[117] B. T. Kudryk, C. R. Martinez, S. Gunasekeran, and G. Ramirez, “CT-guided renal biopsy

using a coaxial technique and an automated biopsy gun.,” South. Med. J., vol. 88, no. 5,

pp. 543–6, May 1995.

[118] M. D. Patel, S. W. Young, J. Scott Kriegshauser, and N. Dahiya, “Ultrasound-guided renal

transplant biopsy: practical and pragmatic considerations,” Abdom. Radiol., vol. 43, no.

10, pp. 2597–2603, Oct. 2018.

[119] M. L. Onozato, P. M. Andrews, Q. Li, J. Jiang, A. Cable, and Y. Chen, “Optical coherence

tomography of human kidney.,” J. Urol., 2010.

[120] W. Drexler, M. Liu, A. Kumar, T. Kamali, A. Unterhuber, and R. A. Leitgeb, “Optical

coherence tomography today: speed, contrast, and multimodality,” J. Biomed. Opt., vol.

19, no. 7, p. 071412, Jul. 2014.

[121] D. Huang et al., “Optical coherence tomography,” Science (80-. )., 1991.

[122] B. H. Bunch and A. Hellemans, The history of science and technology : a browser’s guide

to the great discoveries, inventions, and the people who made them, from the dawn of time

to today. Houghton Mifflin, 2004.

[123] “Tracking Solar Flares.” [Online]. Available: http://solar-

center.stanford.edu/SID/activities/ionosphere.html. [Accessed: 05-Mar-2019].

[124] J. F. de Boer, R. Leitgeb, and M. Wojtkowski, “Twenty-five years of optical coherence

Page 183: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

166

tomography: the paradigm shift in sensitivity and speed provided by Fourier domain OCT

[Invited],” Biomed. Opt. Express, vol. 8, no. 7, p. 3248, Jul. 2017.

[125] M. Adhi and J. S. Duker, “Optical coherence tomography – current and future

applications,” Curr. Opin. Ophthalmol., vol. 24, no. 3, pp. 213–221, May 2013.

[126] A. C. Sull et al., “Comparison of spectral/Fourier domain optical coherence tomography

instruments for assessment of normal macular thickness,” Retina, 2010.

[127] C. V. Regatieri, L. Branchini, J. Carmody, J. G. Fujimoto, and J. S. Duker, “Choroidal

thickness in patients with diabetic retinopathy analyzed by spectral-domain optical

coherence tomography,” Retina, 2012.

[128] Y. Imamura, T. Fujiwara, R. Margolis, and R. F. Spaide, “Enhanced depth imaging optical

coherence tomography of the choroid in central serous chorioretinopathy,” Retina, 2009.

[129] Y. Wang, Y. Zhang, Z. Yao, R. Zhao, and F. Zhou, “Machine learning based detection of

age-related macular degeneration (AMD) and diabetic macular edema (DME) from optical

coherence tomography (OCT) images,” Biomed. Opt. Express, vol. 7, no. 12, p. 4928,

Dec. 2016.

[130] E. Sattler, R. Kästle, and J. Welzel, “Optical coherence tomography in dermatology,” J.

Biomed. Opt., vol. 18, no. 6, p. 061224, Jan. 2013.

[131] Y.-S. Hsieh et al., “Dental optical coherence tomography.,” Sensors (Basel)., vol. 13, no.

7, pp. 8928–49, Jul. 2013.

[132] M. J. Gora, M. J. Suter, G. J. Tearney, and X. Li, “Endoscopic optical coherence

tomography: technologies and clinical applications [Invited].,” Biomed. Opt. Express, vol.

8, no. 5, pp. 2405–2444, May 2017.

[133] D. Huang et al., “Optical coherence tomography.,” Science, vol. 254, no. 5035. pp. 1178–

81, 1991.

[134] R. C. Gonzalez and R. E. (Richard E. Woods, Digital image processing. Prentice Hall,

2008.

[135] M. S. Nixon and A. S. Aguado, “Feature extraction by shape matching,” in Feature

Extraction and Image Processing, 2013.

[136] J. M. S. Prewitt, “Object enhancement and extraction,” Pict. Process. Psychopictorics,

1970.

[137] I. Sobel, “NEIGHBORHOOD CODING OF BINARY IMAGES FOR FAST CONTOUR

FOLLOWING AND GENERAL BINARY ARRAY PROCESSING.,” Comput Graph.

Image Process, 1978.

[138] M. Reuter, S. Biasotti, D. Giorgi, and G. Patanè, “Discrete Laplace–Beltrami operators for

shape analysis and segmentation,” Comput. Graph., vol. 33, no. 3, pp. 381–390, Jun.

2009.

[139] T. Lindeberg, “Scale Selection Properties of Generalized Scale-Space Interest Point

Detectors,” J. Math. Imaging Vis., vol. 46, no. 2, pp. 177–210, Jun. 2013.

[140] Y. C. Hum, K. W. Lai, and M. I. Mohamad Salim, “Multiobjectives bihistogram

Page 184: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

167

equalization for image contrast enhancement,” Complexity, vol. 20, no. 2, pp. 22–36, Oct.

2014.

[141] P. E. Trahanias and A. N. Venetsanopoulos, “Color image enhancement through 3-D

histogram equalization,” in Proceedings., 11th IAPR International Conference on Pattern

Recognition. Vol. IV. Conference D: Architectures for Vision and Pattern Recognition,

pp. 545–548.

[142] S. J. Bedros, “Edge Detection.” [Online]. Available:

http://www.me.umn.edu/courses/me5286/vision/VisionNotes/2017/ME5286-Lecture7-

2017-EdgeDetection2.pdf. [Accessed: 03-Jun-2019].

[143] J. Canny, “A Computational Approach to Edge Detection,” IEEE Trans. Pattern Anal.

Mach. Intell., vol. PAMI-8, no. 6, pp. 679–698, Nov. 1986.

[144] N. Otsu, “A Threshold Selection Method from Gray-Level Histograms,” IEEE Trans.

Syst. Man. Cybern., vol. 9, no. 1, pp. 62–66, Jan. 1979.

[145] M. A. Ayala-Garcia, M. A. P. Hernandez, E. J. Ramirez-Barba, and B. G. Yebra,

“Preservation of Renal Allografts for Transplantation,” Ren. Transplant., pp. 1–17, 2012.

[146] M. Kass, A. Witkin, and D. Terzopoulos, “Snakes: Active contour models,” Int. J.

Comput. Vis., vol. 1, no. 4, pp. 321–331, Jan. 1988.

[147] A. H. Seheult, D. . Greig, and B. T. Porteous;, “Exact Maximum A Posteriori Estimation

For Binary Images,” J. R. Stat. Soc., 1989.

[148] E. H. Adelson and J. R. Bergen, “The plenoptic function and the elements of early

vision.,” in Computational models of visual processing, 1991.

[149] Y. Boykov, O. Veksler, and R. Zabih, “Fast approximate energy minimization via graph

cuts,” IEEE Trans. Pattern Anal. Mach. Intell., 2001.

[150] R. Kafieh, H. Rabbani, and S. Kermani, “A review of algorithms for segmentation of

optical coherence tomography from retina.,” J. Med. Signals Sens., vol. 3, no. 1, pp. 45–

60, Jan. 2013.

[151] D. C. DeBuc, “A review of algorithms for segmentation of retinal image data using optical

coherence tomography,” Image Segmentation, 2011.

[152] M. R. Lee et al., “Optical coherence tomography for ophthalmic imaging - new technique

delivers micron-scale resolution.,” IEEE Eng. Med. Biol., 1995.

[153] Q. Yang et al., “Automated layer segmentation of macular OCT images using dual-scale

gradient information,” Opt. Express, 2010.

[154] D. C. Fernández, N. Villate, C. A. Puliafito, and P. J. Rosenfeld, “Comparing total

macular volume changes measured by Optical Coherence Tomography with retinal lesion

volume estimated by active contours. ,” Invest. Ophthalmol. Vis. Sci., 2004.

[155] A. Yazdanpanah, G. Hamarneh, B. Smith, and M. Sarunic, “Intra-retinal layer

segmentation in optical coherence tomography using an active contour approach,” in

Lecture Notes in Computer Science (including subseries Lecture Notes in Artificial

Intelligence and Lecture Notes in Bioinformatics), 2009.

Page 185: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

168

[156] S. Yousefi, T. Liu, and R. K. Wang, “Segmentation and quantification of blood vessels for

OCT-based micro-angiograms using hybrid shape/intensity compounding.,” Microvasc.

Res., vol. 97, pp. 37–46, Jan. 2015.

[157] M. D. Abràmoff et al., “Automated segmentation of the cup and rim from spectral domain

OCT of the optic nerve head,” Investig. Ophthalmol. Vis. Sci., 2009.

[158] Yuri Boykov and Gareth Funka-Lea, “Graph Cuts and Efficient N-D Image

Segmentation,” Int. J. Comput. Vis., 2006.

[159] R. Kardon, Xiaodong Wu, M. Sonka, M. D. Abramoff, M. K. Garvin, and S. R. Russell,

“Intraretinal Layer Segmentation of Macular Optical Coherence Tomography Images

Using Optimal 3-D Graph Search,” IEEE Trans. Med. Imaging, 2008.

[160] P. M. Andrews, B. S. Khirabadi, and B. C. Bengs, “Using tandem scanning confocal

microscopy to predict the status of donor kidneys.,” Nephron, vol. 91, no. 1, pp. 148–55,

May 2002.

[161] M. Schmid, “Versatile Wand Tool.” ImageJ, 2015.

[162] A. Lubas et al., “Ultrasonic evaluation of renal cortex arterial area enables differentiation

between hypertensive and glomerulonephritis-related chronic kidney disease,” Int. Urol.

Nephrol., vol. 49, no. 9, pp. 1627–1635, 2017.

[163] H. Boom, M. J. K. Mallat, J. W. De Fijter, A. H. Zwinderman, and L. C. Paul, “Delayed

graft function influences renal function, but not survival,” Kidney Int., vol. 58, no. 2, pp.

859–866, 2000.

[164] B. Wang et al., “Optical coherence tomography and computer-aided diagnosis of a murine

model of chronic kidney disease,” J. Biomed. Opt., vol. 22, no. 12, pp. 1–11, 2017.

[165] Q. Li et al., “Automated quantification of microstructural dimensions of the human kidney

using optical coherence tomography (OCT).,” Opt. Express, vol. 17, no. 18, pp. 16000–16,

Aug. 2009.

[166] W. K. Tey, Y. C. Kuang, M. P.-L. Ooi, and J. J. Khoo, “Automated quantification of renal

interstitial fibrosis for computer-aided diagnosis: A comprehensive tissue structure

segmentation method,” Comput. Methods Programs Biomed., vol. 155, pp. 109–120, Mar.

2018.

[167] J. M. Street et al., “Automated quantification of renal fibrosis with Sirius Red and

polarization contrast microscopy.,” Physiol. Rep., vol. 2, no. 7, Jul. 2014.

[168] P. Sarder, B. G. Ginley, J. E. Tomaszewski, K.-Y. Jen, A. Fogo, and S. Jain,

“Computational analysis of the structural progression of human glomeruli in diabetic

nephropathy,” in Medical Imaging 2018: Digital Pathology, 2018.

[169] P. P. Srinivasan, S. J. Heflin, J. A. Izatt, V. Y. Arshavsky, and S. Farsiu, “Automatic

segmentation of up to ten layer boundaries in SD-OCT images of the mouse retina with

and without missing layers due to pathology,” Biomed. Opt. Express, 2014.

[170] Q. Dai and Y. Sun, “Automated layer segmentation of optical coherence tomography

images,” in Proceedings - 2011 4th International Conference on Biomedical Engineering

and Informatics, BMEI 2011, 2011.

Page 186: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

169

[171] S. Lee et al., “Comparison of Outcomes of Living and Deceased Donor Kidney Grafts

Surviving Longer Than 5 Years in Korea,” Transplant. Proc., vol. 42, no. 3, pp. 775–777,

Apr. 2010.

[172] A. J. Matas et al., “2202 Kidney Transplant Recipients with 10 Years of Graft Function:

What Happens Next?,” Am. J. Transplant., vol. 8, no. 11, pp. 2410–2419, Nov. 2008.

[173] “Organ Donation and Transplantation Statistics | National Kidney Foundation.” [Online].

Available: https://www.kidney.org/news/newsroom/factsheets/Organ-Donation-and-

Transplantation-Stats. [Accessed: 06-Mar-2019].

[174] D. Healy, P. Daly, N. Docherty, and R. Watson, “Heat shock induced protection of renal

proximal tubular epithelial cells fom cold storage and rewarming injury,” J. Am. Soc.

Nephrol., vol. 17, no. 3, pp. 805–812, 2006.

[175] F. Belzer, B. S. Ashby, J. Huang, and J. E. Dunphy, “Etiology of Rising Perfusion

Pressure in Isolated Organ Perfusion,” Ann. Surg., vol. 168, no. 3, pp. 382–390.

[176] L. R. Requiao-Moura, M. de Souza Durao Junior, A. C. C. de Matos, and A. Pacheco-

Silva, “Ischemia and reperfusion injury in renal transplantation: hemodynamic and

immunological paradigms,” Einstein (Sao Paulo), vol. 13, no. 1, pp. 129–135, 2015.

[177] S. L. Linas, D. Whittenburg, P. E. Parsons, and J. E. Repine, “Ischemia increases

neutrophil retention and worsens acute renal failure: Role of oxygen metabolites and

ICAM 1,” Kidney Int., 1995.

[178] Y. Xia, P. Friedmann, C. Cortes, and M. Lubetzky, “Influence of Cold Ischemia Time in

Combination with Donor Acute Kidney Injury on Kidney Transplantation Outcomes,”

Elsevier, vol. 221, no. 2, pp. 532–538, 2015.

[179] J. M. Weinberg, M. A. Venkatachalam, N. Roeser, and I. Nissim, “Anaerobic and aerobic

pathways for salvage of proximal tubules from hypoxia-induced mitochondrial injury,”

Am. J. Ren. Physiol., vol. 279, no. 5, pp. F927–F943, 2000.

[180] A. Leaf, “Cell Swelling,” Circulation, vol. XLVIII, no. 3, pp. 455–458, 1973.

[181] M. J. Taylor, “Biology of Cell Survival in the Cold: The Basis for Biopreservation of

Tissues and Organs,” Adv. Biopreservation, pp. 15–62, 2006.

[182] E. L. Greene and M. S. Paller, “Calcium and free radicals in hypoxia/reoxygenation injury

of renal epithelial cells,” Am. Physiol. Soc., vol. 94, pp. F13–F20, 1994.

[183] S. L. Lindell, H. Muir, J. Brassil, and M. J. Mangino, “Hypothermic Machine Perfusion

Preservation of the DCD Kidney: Machine Effects,” J. Transplant., 2013.

[184] S. A. Hosgood, B. Yang, A. Bagul, I. H. Mohamed, and M. L. Nicholson, “A comparison

of hypothermic machine perfusion versus static cold storage in an experimental model of

renal ischemia reperfusion injury,” Transplantation, 2010.

[185] Z. Z. et al., “Hypothermic machine perfusion improves kidney viability through

alleviation of vasospasm and edema of sertoli cell and renal tubular epithelial cell,”

Transplantation, 2014.

[186] J. V Guarrera et al., “‘When good kidneys pump badly’: outcomes of deceased donor

Page 187: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

170

renal allografts with poor pulsatile perfusion characteristics.,” Transpl. Int., vol. 23, no. 4,

pp. 444–6, Apr. 2010.

[187] D. Bon, N. Chatauret, S. Giraud, R. Thuillier, F. Favreau, and T. Hauet, “New strategies

to optimize kidney recovery and preservation in transplantation,” Nature Reviews

Nephrology. 2012.

[188] M. P. Van De Kerkhove et al., “Subnormothermic preservation maintains viability and

function in a porcine hepatocyte culture model simulating bioreactor transport,” Cell

Transplant., 2006.

[189] F. K. Port, R. M. Merion, E. C. Roys, and R. A. Wolfe, “Trends in Organ Donation and

Transplantation in the United States, 1997–2006,” Am. J. Transplant., vol. 8, no. 4p2, pp.

911–921, Apr. 2008.

[190] A. P. K. Lee and D. Abramowicz, “Is the Kidney Donor Risk Index a step forward in the

assessment of deceased donor kidney quality?,” Nephrol. Dial. Transplant., vol. 30, no. 8,

pp. 1285–1290, Aug. 2015.

[191] S. L. Nyberg et al., “Improved scoring system to assess adult donors for cadaver renal

transplantation.,” Am. J. Transplant, vol. 3, no. 6, pp. 715–21, Jun. 2003.

[192] S. L. Nyberg et al., “Donor scoring system for cadaveric renal transplantation.,” Am. J.

Transplant, vol. 1, no. 2, pp. 162–70, Jul. 2001.

[193] J. D. Schold, B. Kaplan, R. S. Baliga, and H.-U. Meier-Kriesche, “The broad spectrum of

quality in deceased donor kidneys.,” Am. J. Transplant, vol. 5, no. 4 Pt 1, pp. 757–65,

Apr. 2005.

[194] S. Ali and N. S. Sheerin, “Biomarkers of acute injury: predicting the long-term outcome

after transplantation,” Kidney Int., vol. 84, no. 6, pp. 1072–1074, Dec. 2013.

[195] R. F. Saidi et al., “Outcome of kidney transplantation using expanded criteria donors and

donation after cardiac death kidneys: realities and costs.,” Am. J. Transplant, vol. 7, no.

12, pp. 2769–74, Dec. 2007.

[196] A. Siedlecki, W. Irish, and D. C. Brennan, “Delayed graft function in the kidney

transplant.,” Am. J. Transplant, vol. 11, no. 11, pp. 2279–96, Nov. 2011.

[197] W. W.H., L. H.G., and P. R.J., “Brain death induced renal injury,” Curr. Opin. Organ

Transplant., 2011.

[198] S. G. Yarlagadda, S. G. Coca, R. N. Formica, E. D. Poggio, and C. R. Parikh,

“Association between delayed graft function and allograft and patient survival: a

systematic review and meta-analysis,” Nephrol. Dial. Transplant., vol. 24, no. 3, pp.

1039–1047, Oct. 2008.

[199] Z. Li, Q. Tang, L. Jin, P. M. Andrews, and Y. Chen, “Monitoring Kidney Microanatomy

Changes During Ischemia-Reperfusion Process Using Texture Analysis of OCT Images,”

IEEE Photonics J., vol. 9, no. 2, p. 4000110, Apr. 2017.

[200] P. M. Andrews et al., “High-resolution optical coherence tomography imaging of the

living kidney.,” Lab. Investig., vol. 88, no. 4, pp. 441–449, 2008.

Page 188: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

171

[201] R. A. Risdon, J. C. Sloper, and H. E. De Wardener, “Relationship between renal function

and histological changes found in renal-biopsy specimens from patients with persistent

glomerular nephritis.,” Lancet (London, England), vol. 2, no. 7564, pp. 363–6, Aug. 1968.

[202] G. Remuzzi et al., “Early experience with dual kidney transplantation in adults using

expanded donor criteria. Double Kidney Transplant Group (DKG).,” J. Am. Soc.

Nephrol., vol. 10, no. 12, pp. 2591–8, Dec. 1999.

[203] R. B. Munivenkatappa et al., “The Maryland Aggregate Pathology Index: A Deceased

Donor Kidney Biopsy Scoring System for Predicting Graft Failure,” Am. J. Transplant.,

vol. 8, no. 11, pp. 2316–2324, Nov. 2008.

[204] S. Lovisa, M. Zeisberg, and R. Kalluri, “Partial Epithelial-to-Mesenchymal Transition and

Other New Mechanisms of Kidney Fibrosis.,” Trends Endocrinol. Metab., vol. 27, no. 10,

pp. 681–695, Oct. 2016.

[205] Y. Ovadya and V. Krizhanovsky, “A new Twist in kidney fibrosis,” Nat. Med., vol. 21,

no. 9, pp. 975–977, Sep. 2015.

[206] J. R. Schelling, “Tubular atrophy in the pathogenesis of chronic kidney disease

progression.,” Pediatr. Nephrol., vol. 31, no. 5, pp. 693–706, May 2016.

[207] M. L. Nicholson, E. Bailey, S. Williams, K. P. Harris, and P. N. Furness, “Computerized

histomorphometric assessment of protocol renal transplant biopsy specimens for surrogate

markers of chronic rejection.,” Transplantation, vol. 68, no. 2, pp. 236–41, Jul. 1999.

[208] R. L. Chevalier, “The proximal tubule is the primary target of injury and progression of

kidney disease: role of the glomerulotubular junction.,” Am. J. Physiol. Renal Physiol.,

vol. 311, no. 1, pp. F145-61, 2016.

[209] S. Cortell, F. J. Gennari, M. Davidman, W. H. Bossert, and W. B. Schwartz, “A definition

of proximal and distal tubular compliance. Practical and theoretical implications,” J. Clin.

Invest., vol. 52, no. 9, pp. 2330–2339, 1973.

[210] A. B. Maunsbach, G. H. Giebisch, and B. A. Stanton, “Effects of flow rate on proximal

tubule ultrastructure,” Am. J. Physiol. Physiol., vol. 253, no. 3, pp. F582–F587, Sep. 1987.

[211] J. V. Bonventre and L. Yang, “Cellular pathophysiology of ischemic acute kidney injury,”

J. Clin. Invest., vol. 121, no. 11, pp. 4210–4221, 2011.

[212] M. L. Onozato, P. M. Andrews, Q. Li, J. Jiang, A. E. Cable, and Y. Chen, “Optical

coherence tomography of human kidney.,” J. Urol., vol. 183, no. 5, pp. 2090–2094, 2010.

[213] J. Wierwille, P. Andrews, M. Onozato, and Y. Chen, “In Vivo, label-free, three-

dimensional quantitative imaging of kidney microcirculation using Doppler optical

coherence tomograpy (DOCT),” Lab Investig., vol. 91, no. 11, pp. 1596–1604, 2011.

[214] P. M Andrews and Y. Chen, “Using Optical Coherence Tomography (OCT) to Evaluate

Human Donor Kidneys Prior to and Following Transplantation,” J. Nephrol. Ther., vol.

04, no. 01, p. 1000151, Jan. 2014.

[215] W. D. Irish, J. N. Ilsley, M. A. Schnitzler, S. Feng, and D. C. Brennan, “A Risk Prediction

Model for Delayed Graft Function in the Current Era of Deceased Donor Renal

Transplantation,” Am. J. Transplant., vol. 10, no. 10, pp. 2279–2286, Sep. 2010.

Page 189: OPTICAL COHERENCE TOMOGRAPHY OF HUMAN …

172

[216] S. M. Flechner, J. Kobashigawa, and G. Klintmalm, “Calcineurin inhibitor-sparing

regimens in solid organ transplantation: focus on improving renal function and

nephrotoxicity.,” Clin. Transplant., vol. 22, no. 1, pp. 1–15, Jun. 2007.

[217] S. Yousefi, J. Qin, Z. Zhi, and R. K. Wang, “Label-free optical lymphangiography:

development of an automatic segmentation method applied to optical coherence

tomography to visualize lymphatic vessels using Hessian filters,” J. Biomed. Opt., vol. 18,

no. 8, p. 86004, 2013.

[218] P. Gong et al., “In vivo label-free lymphangiography of cutaneous lymphatic vessels in

human burn scars using optical coherence tomography,” Biomed. Opt. Express, vol. 7, no.

12, pp. 4886–4898, 2016.